You are on page 1of 95

pdf version of the entry

Inductive Logic
https://plato.stanford.edu/archives/spr2021/entries/logic-inductive/ Inductive Logic
from the Spring 2021 Edition of the First published Mon Sep 6, 2004; substantive revision Mon Mar 19, 2018

Stanford Encyclopedia An inductive logic is a logic of evidential support. In a deductive logic, the
premises of a valid deductive argument logically entail the conclusion,
of Philosophy where logical entailment means that every logically possible state of
affairs that makes the premises true must make the conclusion true as well.
Thus, the premises of a valid deductive argument provide total support for
the conclusion. An inductive logic extends this idea to weaker arguments.
In a good inductive argument, the truth of the premises provides some
degree of support for the truth of the conclusion, where this degree-of-
Edward N. Zalta Uri Nodelman Colin Allen R. Lanier Anderson
support might be measured via some numerical scale. By analogy with the
Principal Editor Senior Editor Associate Editor Faculty Sponsor
notion of deductive entailment, the notion of inductive degree-of-support
Editorial Board
https://plato.stanford.edu/board.html might mean something like this: among the logically possible states of
affairs that make the premises true, the conclusion must be true in (at
Library of Congress Catalog Data
ISSN: 1095-5054
least) proportion r of them—where r is some numerical measure of the
support strength.
Notice: This PDF version was distributed by request to mem-
bers of the Friends of the SEP Society and by courtesy to SEP If a logic of good inductive arguments is to be of any real value, the
content contributors. It is solely for their fair use. Unauthorized measure of support it articulates should be up to the task. Presumably, the
distribution is prohibited. To learn how to join the Friends of the logic should at least satisfy the following condition:
SEP Society and obtain authorized PDF versions of SEP entries,
please visit https://leibniz.stanford.edu/friends/ . Criterion of Adequacy (CoA):
The logic should make it likely (as a matter of logic) that as evidence
Stanford Encyclopedia of Philosophy accumulates, the total body of true evidence claims will eventually
Copyright c 2021 by the publisher
The Metaphysics Research Lab come to indicate, via the logic’s measure of support, that false
Center for the Study of Language and Information hypotheses are probably false and that true hypotheses are probably
Stanford University, Stanford, CA 94305
true.
Inductive Logic
c 2021 by the author
Copyright The CoA stated here may strike some readers as surprisingly strong. Given
James Hawthorne
a specific logic of evidential support, how might it be shown to satisfy
All rights reserved.
Copyright policy: https://leibniz.stanford.edu/friends/info/copyright/

1
Inductive Logic James Hawthorne

such a condition? Section 4 will show precisely how this condition is evidence accumulates. This result shows that the Criterion of Adequacy is
satisfied by the logic of evidential support articulated in Sections 1 indeed satisfied—that as evidence accumulates, false hypotheses will very
through 3 of this article. probably come to have evidential support values (as measured by their
posterior probabilities) that approach 0; and as this happens, a true
This article will focus on the kind of the approach to inductive logic most hypothesis may very probably acquire evidential support values (as
widely studied by epistemologists and logicians in recent years. This measured by its posterior probability) that approaches 1.
approach employs conditional probability functions to represent measures
of the degree to which evidence statements support hypotheses. 1. Inductive Arguments
Presumably, hypotheses should be empirically evaluated based on what 2. Inductive Logic and Inductive Probabilities
they say (or imply) about the likelihood that evidence claims will be true. 2.1 The Historical Origins of Probabilistic Logic
A straightforward theorem of probability theory, called Bayes’ Theorem, 2.2 Probabilistic Logic: Axioms and Characteristics
articulates the way in which what hypotheses say about the likelihoods of 2.3 Two Conceptions of Inductive Probability
evidence claims influences the degree to which hypotheses are supported 3. The Application of Inductive Probabilities to the Evaluation of
by those evidence claims. Thus, this approach to the logic of evidential Scientific Hypotheses
support is often called a Bayesian Inductive Logic or a Bayesian 3.1 Likelihoods
Confirmation Theory. This article will first provide a detailed explication 3.2 Posterior Probabilities and Prior Probabilities
of a Bayesian approach to inductive logic. It will then examine the extent 3.3 Bayes’ Theorem
to which this logic may pass muster as an adequate logic of evidential 3.4 On Prior Probabilities and Representations of Vague and
support for hypotheses. In particular, we will see how such a logic may be Diverse Plausibility Assessments
shown to satisfy the Criterion of Adequacy stated above. 4. The Likelihood Ratio Convergence Theorem
4.1 The Space of Possible Outcomes of Experiments and
Sections 1 through 3 present all of the main ideas underlying the Observations
(Bayesian) probabilistic logic of evidential support. These three sections 4.2 Probabilistic Independence
should suffice to provide an adequate understanding of the subject. Section 4.3 Likelihood Ratio Convergence when Falsifying Outcomes
5 extends this account to cases where the implications of hypotheses about are Possible
evidence claims (called likelihoods) are vague or imprecise. After reading 4.4 Likelihood Ratio Convergence When No Falsifying
Sections 1 through 3, the reader may safely skip directly to Section 5, Outcomes are Possible
bypassing the rather technical account in Section 4 of how how the CoA is 5. When the Likelihoods are Vague or Diverse
satisfied. List of Supplements
Bibliography
Section 4 is for the more advanced reader who wants an understanding of
Academic Tools
how this logic may bring about convergence to the true hypothesis as
Other Internet Resources

2 Stanford Encyclopedia of Philosophy Spring 2021 Edition 3


Inductive Logic James Hawthorne

Related Entries Semi-formalization Formalization


Premise 1 The frequency (or proportion) of F[A, S] = r
members with attribute A among the
1. Inductive Arguments members of S is r.
Premise 2 S is a random sample of B with respect Rnd[S, B, A]
Let us begin by considering some common kinds of examples of inductive to whether or not its members have A
arguments. Consider the following two arguments: Premise 3 Sample S has exactly n members Size[S] = n
Therefore with degree of support p ======== {p}
Example 1. Every raven in a random sample of 3200 ravens is black. Conclusion The proportion of members of B that F[A, B] = r ± q
This strongly supports the following conclusion: All ravens are black. have attribute A lies between r − q
and r + q (i.e., lies within margin of
Example 2. 62 percent of voters in a random sample of 400 registered error q of r)
voters (polled on February 20, 2004) said that they favor John Kerry
over George W. Bush for President in the 2004 Presidential election. Any inductive logic that treats such arguments should address two
This supports with a probability of at least .95 the following challenges. (1) It should tell us which enumerative inductive arguments
conclusion: Between 57 percent and 67 percent of all registered voters should count as good inductive arguments. In particular, it should tell us
favor Kerry over Bush for President (at or around the time the poll how to determine the appropriate degree p to which such premises
was taken). inductively support the conclusion, for a given margin of error q. (2) It
should demonstrably satisfy the CoA. That is, it should be provable (as a
This kind of argument is often called an induction by enumeration. It is metatheorem) that if a conclusion expressing the approximate proportion
closely related to the technique of statistical estimation. We may represent for an attribute in a population is true, then it is very likely that sufficiently
the logical form of such arguments semi-formally as follows: numerous random samples of the population will provide true premises for
good inductive arguments that confer degrees of support p approaching 1
Premise: In random sample S consisting of n members of population for that true conclusion—where, on pain of triviality, these sufficiently
B, the proportion of members that have attribute A is r. numerous samples are only a tiny fraction of a large population. The
supplement on Enumerative Inductions: Bayesian Estimation and
Therefore, with degree of support p,
Convergence, shows precisely how a a Bayesian account of enumerative
Conclusion: The proportion of all members of B that have attribute A induction may meet these two challenges.
is between r − q and r + q (i.e., lies within margin of error q of r).
Enumerative induction is, however, rather limited in scope. This form of
Let’s lay out this argument more formally. The premise breaks down into induction is only applicable to the support of claims involving simple
three separate statements:[1] universal conditionals (i.e., claims of form ‘All Bs are As’) and claims
about the proportion of an attribute in a population (i.e., claims of form

4 Stanford Encyclopedia of Philosophy Spring 2021 Edition 5


Inductive Logic James Hawthorne

‘the frequency of As among the Bs is r’). But, many important empirical kinds. It almost never involves consideration of a randomly selected
hypotheses are not reducible to this simple form, and the evidence for sequences of past situations when people like the accused committed
these hypotheses is not composed of an enumeration of such instances. similar murders. Or, consider how a doctor diagnoses her patient on the
Consider, for example, the Newtonian Theory of Mechanics: basis of his symptoms. Although the frequency of occurrence of various
diseases when similar symptoms have been present may play a role, this is
All objects remain at rest or in uniform motion unless acted upon by clearly not the whole story. Diagnosticians commonly employ a form of
some external force. An object’s acceleration (i.e., the rate at which its hypothesis evaluation—e.g., would the hypothesis that the patient has a
motion changes from rest or from uniform motion) is in the same brain tumor account for his symptoms?; or are these symptoms more
direction as the force exerted on it; and the rate at which the object likely the result of a minor stroke?; or may some other hypothesis better
accelerates due to a force is equal to the magnitude of the force account for the patient’s symptoms? Thus, a fully adequate account of
divided by the object’s mass. If an object exerts a force on another inductive logic should explicate the logic of hypothesis evaluation,
object, the second object exerts an equal amount of force on the first through which a hypothesis or theory may be tested on the basis of what it
object, but in the opposite direction to the force exerted by the first says (or "predicts") about observable phenomena. In Section 3 we will see
object. how a kind of probabilistic inductive logic called "Bayesian Inference" or
"Bayesian Confirmation Theory" captures such reasoning. The full logical
The evidence for (and against) this theory is not gotten by examining a
structure of such arguments will be spelled out in that section.
randomly selected subset of objects and the forces acting upon them.
Rather, the theory is tested by calculating what this theory says (or
implies) about observable phenomena in a wide variety of specific
2. Inductive Logic and Inductive Probabilities
situations—e.g., ranging from simple collisions between small bodies to
Perhaps the oldest and best understood way of representing partial belief,
the trajectories of planets and comets—and then seeing whether those
uncertain inference, and inductive support is in terms of probability and
phenomena occur in the way that the theory says they will. This approach
the equivalent notion odds. Mathematicians have studied probability for
to testing hypotheses and theories is ubiquitous, and should be captured by
over 350 years, but the concept is certainly much older. In recent times a
an adequate inductive logic.
number of other, related representations of partial belief and uncertain
More generally, for a wide range of cases where inductive reasoning is inference have emerged. Some of these approaches have found useful
important, enumerative induction is inadequate. Rather, the kind of application in computer based artificial intelligence systems that perform
evidential reasoning that judges the likely truth of hypotheses on the basis inductive inferences in expert domains such as medical diagnosis.
of what they say (or imply) about the evidence is more appropriate. Nevertheless, probabilistic representations have predominated in such
Consider the kinds of inferences jury members are supposed to make, application domains. So, in this article we will focus exclusively on
based on the evidence presented at a murder trial. The inference to probabilistic representations of inductive support. A brief comparative
probable guilt or innocence is based on a patchwork of evidence of various description of some of the most prominent alternative representations of

6 Stanford Encyclopedia of Philosophy Spring 2021 Edition 7


Inductive Logic James Hawthorne

uncertainty and support-strength can be found in the supplement Some deductive logic spurred some logicians to attempt to apply a similar
Prominent Approaches to the Representation of Uncertain Inference. approach to inductive reasoning. The idea was to extend the deductive
entailment relation to a notion of probabilistic entailment for cases where
2.1 The Historical Origins of Probabilistic Logic premises provide less than conclusive support for conclusions. These
partial entailments are expressed in terms of conditional probabilities,
The mathematical study of probability originated with Blaise Pascal and probabilities of the form P[C ∣ B] = r (read “the probability of C given B
Pierre de Fermat in the mid-17th century. From that time through the early is r”), where P is a probability function, C is a conclusion sentence, B is a
19th century, as the mathematical theory continued to develop, probability conjunction of premise sentences, and r is the probabilistic degree of
theory was primarily applied to the assessment of risk in games of chance support that premises B provide for conclusion C. Attempts to develop
and to drawing simple statistical inferences about characteristics of large such a logic vary somewhat with regard to the ways in which they attempt
populations—e.g., to compute appropriate life insurance premiums based to emulate the paradigm of formal deductive logic.
on mortality rates. In the early 19th century Pierre de Laplace made further
theoretical advances and showed how to apply probabilistic reasoning to a Some inductive logicians have tried to follow the deductive paradigm by
much wider range of scientific and practical problems. Since that time attempting to specify inductive support probabilities solely in terms of the
probability has become an indispensable tool in the sciences, business, and syntactic structures of premise and conclusion sentences. In deductive
many other areas of modern life. logic the syntactic structure of the sentences involved completely
determines whether premises logically entail a conclusion. So these
Throughout the development of probability theory various researchers inductive logicians have attempted to follow suit. In such a system each
appear to have thought of it as a kind of logic. But the first extended sentence confers a syntactically specified degree of support on each of the
treatment of probability as an explicit part of logic was George Boole’s other sentences of the language. Thus, the inductive probabilities in such a
The Laws of Thought (1854). John Venn followed two decades later with system are logical in the sense that they depend on syntactic structure
an alternative empirical frequentist account of probability in The Logic of alone. This kind of conception was articulated to some extent by John
Chance (1876). Not long after that the whole discipline of logic was Maynard Keynes in his Treatise on Probability (1921). Rudolf Carnap
transformed by new developments in deductive logic. pursued this idea with greater rigor in his Logical Foundations of
Probability (1950) and in several subsequent works (e.g., Carnap 1952).
In the late 19th and early 20th century Frege, followed by Russell and
(For details of Carnap’s approach see the section on logical probability in
Whitehead, showed how deductive logic may be represented in the kind of
the entry on interpretations of the probability calculus, in this
rigorous formal system we now call quantified predicate logic. For the
Encyclopedia.)
first time logicians had a fully formal deductive logic powerful enough to
represent all valid deductive arguments that arise in mathematics and the In the inductive logics of Keynes and Carnap, Bayes’ theorem, a
sciences. In this logic the validity of deductive arguments depends only on straightforward theorem of probability theory, plays a central role in
the logical structure of the sentences involved. This development in expressing how evidence comes to bear on hypotheses. Bayes’ theorem

8 Stanford Encyclopedia of Philosophy Spring 2021 Edition 9


Inductive Logic James Hawthorne

expresses how the probability of a hypothesis h on the evidence e, with their syntactic relationships to evidence statements). Are we to
P[h ∣ e] , depends on the probability that e should occur if h is true, evaluate alternative theories of gravitation, and alternative quantum
P[e ∣ h] , and on the probability of hypothesis h prior to taking the theories, this way? This seems an extremely dubious approach to the
evidence into account, P[h] (called the prior probability of h). (Later we’ll evaluation of real scientific hypotheses and theories. Thus, it seems that
examine Bayes’ theorem in detail.) So, such approaches might well be logical structure alone may not suffice for the inductive evaluation of
called Bayesian logicist inductive logics. Other prominent Bayesian scientific hypotheses. (This issue will be treated in more detail in Section
logicist attempts to develop a probabilistic inductive logic include the 3, after we first see how probabilistic logics employ Bayes’ theorem to
works of Jeffreys (1939), Jaynes (1968), and Rosenkrantz (1981). represent the evidential support for hypotheses as a function of prior
probabilities together with evidential likelihoods.)
It is now widely held that the core idea of this syntactic approach to
Bayesian logicism is fatally flawed—that syntactic logical structure cannot At about the time that the syntactic Bayesian logicist idea was developing,
be the sole determiner of the degree to which premises inductively support an alternative conception of probabilistic inductive reasoning was also
conclusions. A crucial facet of the problem faced by syntactic Bayesian emerging. This approach is now generally referred to as the Bayesian
logicism involves how the logic is supposed to apply in scientific contexts subjectivist or personalist approach to inductive reasoning (see, e.g.,
where the conclusion sentence is some scientific hypothesis or theory, and Ramsey 1926; De Finetti 1937; Savage 1954; Edwards, Lindman, &
the premises are evidence claims. The difficulty is that in any probabilistic Savage 1963; Jeffrey 1983, 1992; Howson & Urbach 1993; Joyce 1999).
logic that satisfies the usual axioms for probabilities, the inductive support This approach treats inductive probability as a measure of an agent’s
for a hypothesis must depend in part on its prior probability. This prior degree-of-belief that a hypothesis is true, given the truth of the evidence.
probability represents (arguably) how plausible the hypothesis is taken to This approach was originally developed as part of a larger normative
be on the basis of considerations other than the observational and theory of belief and action known as Bayesian decision theory. The
experimental evidence (e.g., perhaps due to various plausibility principal idea is that the strength of an agent’s desires for various possible
arguments). A syntactic Bayesian logicist must tell us how to assign values outcomes should combine with her belief-strengths regarding claims about
to these pre-evidential prior probabilities of hypotheses in a way that the world to produce optimally rational decisions. Bayesian subjectivists
relies only on the syntactic logical structure of the hypothesis, perhaps provide a logic of decision that captures this idea, and they attempt to
based on some measure of syntactic simplicity. There are severe problems justify this logic by showing that in principle it leads to optimal decisions
with getting this idea to work. Various kinds of examples seem to show about which of various risky alternatives should be pursued. On the
that such an approach must assign intuitively quite unreasonable prior Bayesian subjectivist or personalist account of inductive probability,
probabilities to hypotheses in specific cases (see the footnote cited near the inductive probability functions represent the subjective (or personal)
end of Section 3.2 for details). Furthermore, for this idea to apply to the belief-strengths of ideally rational agents, the kind of belief strengths that
evidential support of real scientific theories, scientists would have to figure into rational decision making. (See the section on subjective
formalize theories in a way that makes their relevant syntactic structures probability in the entry on interpretations of the probability calculus, in
apparent, and then evaluate theories solely on that syntactic basis (together this Encyclopedia.)

10 Stanford Encyclopedia of Philosophy Spring 2021 Edition 11


Inductive Logic James Hawthorne

Elements of a logicist conception of inductive logic live on today as part 2.2 Probabilistic Logic: Axioms and Characteristics
of the general approach called Bayesian inductive logic. However, among
philosophers and statisticians the term ‘Bayesian’ is now most closely All logics derive from the meanings of terms in sentences. What we now
associated with the subjectivist or personalist account of belief and recognize as formal deductive logic rests on the meanings (i.e., the truth-
decision. And the term ‘Bayesian inductive logic’ has come to carry the functional properties) of the standard logical terms. These logical terms,
connotation of a logic that involves purely subjective probabilities. This and the symbols we will employ to represent them, are as follows:
usage is misleading since, for inductive logics, the Bayesian/non-Bayesian
distinction should really turn on whether the logic gives Bayes’ theorem a ‘not’, ‘∼’;
prominent role, or the approach largely eschews the use of Bayes’ theorem ‘and’, ‘⋅ ’;
in inductive inferences, as do the classical approaches to statistical ‘inclusive or’, ‘∨’;
inference developed by R. A. Fisher (1922) and by Neyman & Pearson truth-functional ‘if-then’, ‘⊃’;
(1967)). Indeed, any inductive logic that employs the same probability ‘if and only if’, ‘≡’;
functions to represent both the probabilities of evidence claims due to the quantifiers
hypotheses and the probabilities of hypotheses due to those evidence ‘all’, ‘∀’, and
claims must be a Bayesian inductive logic in this broader sense; because ‘some’, ‘∃’;
Bayes’ theorem follows directly from the axioms that each probability and
function must satisfy, and Bayes’ theorem expresses a necessary the identity relation, ‘=’.
connection between the probabilities of evidence claims due to hypotheses
The meanings of all other terms, the non-logical terms such as names and
and the probabilities of hypotheses due to those evidence claims.
predicate and relational expressions, are permitted to “float free”. That is,
In this article the probabilistic inductive logic we will examine is a the logical validity of deductive arguments depends neither on the
Bayesian inductive logic in this broader sense. This logic will not meanings of the name and predicate and relation terms, nor on the truth-
presuppose the subjectivist Bayesian theory of belief and decision, and values of sentences containing them. It merely supposes that these non-
will avoid the objectionable features of the syntactic version of Bayesian logical terms are meaningful, and that sentences containing them have
logicism. We will see that there are good reasons to distinguish inductive truth-values. Deductive logic then tells us that the logical structures of
probabilities from degree-of-belief probabilities and from purely syntactic some sentences—i.e., the syntactic arrangements of their logical terms—
logical probabilities. So, the probabilistic logic articulated in this article preclude them from being jointly true of any possible state of affairs. This
will be presented in a way that depends on neither of these conceptions of is the notion of logical inconsistency. The notion of logical entailment is
what the probability functions are. However, this version of the logic will inter-definable with it. A collection of premise sentences logically entails a
be general enough that it may be fitted to a Bayesian subjectivist or conclusion sentence just when the negation of the conclusion is logically
Bayesian syntactic-logicist program, if one desires to do that. inconsistent with those premises.

12 Stanford Encyclopedia of Philosophy Spring 2021 Edition 13


Inductive Logic James Hawthorne

An inductive logic must, it seems, deviate from the paradigm provided by Although each support function satisfies these same axioms, the further
deductive logic in several significant ways. For one thing, logical issue of which among them provides an appropriate measure of inductive
entailment is an absolute, all-or-nothing relationship between sentences, support is not settled by the axioms alone. That may depend on additional
whereas inductive support comes in degrees-of-strength. For another, factors, such as the meanings of the non-logical terms (i.e., the names and
although the notion of inductive support is analogous to the deductive predicate expressions) of the language.
notion of logical entailment, and is arguably an extension of it, there
seems to be no inductive logic extension of the notion of logical A good way to specify the axioms of the logic of inductive support
inconsistency—at least none that is inter-definable with inductive support functions is as follows. These axioms are apparently weaker than the usual
in the way that logical inconsistency is inter-definable with logical axioms for conditional probabilities. For instance, the usual axioms
entailment. Indeed, it turns out that when the unconditional probability of assume that conditional probability values are restricted to real numbers
(B ⋅ ∼A) is very nearly 0 (i.e., when (B ⋅ ∼A) is “nearly inconsistent”), the between 0 and 1. The following axioms do not assume this, but only that
degree to which B inductively supports A, P[A ∣ B] , may range anywhere support functions assign some real numbers as values for support
between 0 and 1. strengths. However, it turns out that the following axioms suffice to derive
all the usual axioms for conditional probabilities (including the usual
Another notable difference is that when B logically entails A, adding a restriction to values between 0 and 1). We draw on these weaker axioms
premise C cannot undermine the logical entailment—i.e., (C ⋅ B) must only to forestall some concerns about whether the support function axioms
logically entail A as well. This property of logical entailment is called may assume too much, or may be overly restrictive.
monotonicity. But inductive support is nonmonotonic. In general,
depending on what A, B, and C mean, adding a premise C to B may Let L be a language for predicate logic with identity, and let ‘⊨ ’ be the
substantially raise the degree of support for A, or may substantially lower standard logical entailment relation—i.e., the expression ‘B ⊨ A’ says
it, or may leave it completely unchanged—i.e., P[A ∣ (C ⋅ B)] may have a “B logically entails A” and the expression ‘⊨ A ’ says “A is a
value much larger than P[A ∣ B] , or may have a much smaller value, or it tautology”. A support function is a function Pα from pairs of sentences
may have the same, or nearly the same value as P[A ∣ B] . of L to real numbers that satisfies the following axioms:

In a formal treatment of probabilistic inductive logic, inductive support is (1) Pα [E ∣ F] ≠ Pα [G ∣ H] for at least some sentences E, F, G , and
represented by conditional probability functions defined on sentences of a H.
formal language L. These conditional probability functions are constrained
For all sentence A, B, C, and D:
by certain rules or axioms that are sensitive to the meanings of the logical
terms (i.e., ‘not’, ‘and’, ‘or’, etc., the quantifiers ‘all’ and ‘some’, and the (2) If B ⊨ A, then Pα [A ∣ B] ≥ Pα [C ∣ D];
identity relation). The axioms apply without regard for what the other (3) Pα [A ∣ (B ⋅ C)] = Pα [A ∣ (C ⋅ B)];
terms of the language may mean. In essence the axioms specify a family (4) If C ⊨ ∼(B ⋅ A) , then either
of possible support functions, {Pβ , Pγ , … , Pδ , …} for a given language L.

14 Stanford Encyclopedia of Philosophy Spring 2021 Edition 15


Inductive Logic James Hawthorne

Pα [(A ∨ B) ∣ C] = Pα [A ∣ C] + Pα [B ∣ C] The above axioms are quite weak. For instance, they do not say that
logically equivalent sentences are supported by all other sentences to the
or else same degree; rather, that result is derivable from these axioms (see result 6

Pα [E ∣ C] = Pα [C ∣ C]
below). Nor do these axioms say that logically equivalent sentences
support all other sentences to the same degree; rather, that result is also
for every sentence E; derivable (see result 8 below). Indeed, from these axioms all of the usual
(5) Pα [(A ⋅ B) ∣ C] = Pα [A ∣ (B ⋅ C)] × Pα [B ∣ C]. theorems of probability theory may be derived. The following results are
particularly useful in probabilistic logic. Their derivations from these
This axiomatization takes conditional probability as basic, as seems axioms are provided in note 2.[2]
appropriate for evidential support functions. (These functions agree with
the more usual unconditional probability functions when the latter are a. If B ⊨ A, then Pα [A ∣ B] = 1 .
defined—just let Pα [A] = Pα [A ∣ (D ∨ ∼D)]. However, these axioms b. If C ⊨ ∼(B ⋅ A) , then either
permit conditional probabilities Pα [A ∣ C] to remain defined even when
Pα [(A ∨ B) ∣ C] = Pα [A ∣ C] + Pα [B ∣ C]
condition statement C has probability 0—i.e., even when
Pα [C ∣ (D ∨ ∼D)] = 0 .) or else Pα [E ∣ C] = 1 for every sentence E.
c. Pα [∼A ∣ B] = 1 − Pα [A ∣ B] or else Pα [C ∣ B] = 1 for every
Notice that conditional probability functions apply only to pairs of
sentence C.
sentences, a conclusion sentence and a premise sentence. So, in
d. 1 ≥ Pα [A ∣ B] ≥ 0.
probabilistic inductive logic we represent finite collections of premises by
e. If B ⊨ A, then Pα [A ∣ C] ≥ Pα [B ∣ C].
conjoining them into a single sentence. Rather than say,
f. If B ⊨ A and A ⊨ B, then Pα [A ∣ C] = Pα [B ∣ C].
A is supported to degree r by the set of premises {B1 , B2 , B3 ,…, Bn } , g. If C ⊨ B, then Pα [(A ⋅ B) ∣ C] = Pα [(B ⋅ A) ∣ C] = Pα [A ∣ C].
h. If C ⊨ B and B ⊨ C, then Pα [A ∣ B] = Pα [A ∣ C] .
we instead say that i. Pα [B ∣ C] > 0 , then

Pα [A ∣ C]
Pα [A ∣ (B ⋅ C)] = Pα [B ∣ (A ⋅ C)] ×
A is supported to degree r by the conjunctive premise
(((B1 ⋅ B2 ) ⋅ B3 ) ⋅ … ⋅ Bn ), Pα [B ∣ C]

and write this as (this is a simple form of Bayes’ theorem).


j. Pα [(A ∨ B) ∣ C] = Pα [A ∣ C] + Pα [B ∣ C] − Pα [(A ⋅ B) ∣ C].
P[A ∣ (((B1 ⋅ B2 ) ⋅ B3 ) ⋅ … ⋅ Bn )] = r. k. If {B1 , … , Bn } is any finite set of sentences such that for each pair Bi
and Bj , C ⊨ ∼(Bi ⋅ Bj ) (i.e., the members of the set are mutually
exclusive, given C), then either Pα [D ∣ C] = 1 for every sentence D,

16 Stanford Encyclopedia of Philosophy Spring 2021 Edition 17


Inductive Logic James Hawthorne

or Axiom 2 asserts that when B logically entail A, the support of A by B is as


strong as support can possibly be. This comports with the idea that an
n
Pα [((B1 ∨ B2 ) ∨ … ∨ Bn ) ∣ C] = ∑ Pα [Bi ∣ C]. inductive support function is a generalization of the deductive entailment
i=1 relation, where the premises of deductive entailments provide the strongest
possible support for their conclusions.
l. If {B1 , … , Bn , …} is any countably infinite set of sentences such that
for each pair Bi and Bj , C ⊨ ∼(Bi ⋅ Bj ) , then either Pα [D ∣ C] = 1 for Axiom 3 merely says that (B ⋅ C) supports sentences to precisely the same
every sentence D, or[3] degree that (C ⋅ B) supports them. This is an especially weak axiom. But
∞ taken together with the other axioms, it suffices to entail that logically
lim Pα [((B1 ∨ B2 ) ∨ … ∨ Bn ) ∣ C] = ∑ Pα [Bi ∣ C]. equivalent sentences support all sentences to precisely the same degree.
n
i=1
Axiom 4 says that inductive support adds up in a plausible way. When C
Let us now briefly consider each axiom to see how plausible it is as a logically entails the incompatibility of A and B, i.e., when no possible state
constraint on a quantitative measure of inductive support, and how it of affairs can make both A and B true together, the degrees of support that
extends the notion of deductive entailment. First notice that each degree- C provides to each of them individually must sum to the support it
of-support function Pα on L measures support strength with some real provides to their disjunction. The only exception is in those cases where C
number values, but the axioms don’t explicitly restrict these values to lie acts like a logical contradiction and supports all sentences to the maximum
between 0 and 1. It turns out that the all support values must lie between 0 possible degree (in deductive logic a logical contradiction logically entails
and 1, but this follows from the axioms, rather than being assumed by every sentence).
them. The scaling of inductive support via the real numbers is surely a
reasonable way to go. To understand what axiom 5 says, think of a support function Pα as
describing a measure on possible states of affairs. Read each degree-of-
Axiom 1 is a non-triviality requirement. It says that the support values support expression of form ‘Pα [D ∣ E] = r ’ to say that the proportion of
cannot be the same for all sentence pairs. This axiom merely rules out the states of affairs in which D is true among those states of affairs where E is
trivial support function that assigns the same amount of support to each true is r. Read this way, axiom 5 then says the following. Suppose B is true
sentence by every sentence. One might replace this axiom with the in proportion q of all the states of affairs where C is true, and suppose A is
following rule: true in fraction r of those states where B and C are true together. Then A

Pα [(A ∨ ∼A) ∣ (A ∨ ∼A)] ≠ Pα [(A ⋅ ∼A) ∣ (A ∨ ∼A)].


and B should be true together in what proportion of all the states where C
is true? In fraction r (the (A ⋅ B) part) of proportion q (the B portion) of all
But this alternative rule turns out to be derivable from axiom 1 together those states where C is true.
with the other axioms.

18 Stanford Encyclopedia of Philosophy Spring 2021 Edition 19


Inductive Logic James Hawthorne

The degree to which a sentence B supports a sentence A may well depend (6) If A is an axiom of set theory or any other piece of pure mathematics
on what these sentences mean. In particular it will usually depend on the employed by the sciences, or if A is analytically true (i.e., if the truth
meanings we associate with the non-logical terms (those terms other than of A depends only on the meanings of the words it contains, where
the logical terms not, and, or, etc., the quantifiers, and identity), that is, on the specific meanings for names and predicates are those associated
the meanings of the names, and the predicate and relation terms of the with the particular support function Pα ), then, for all sentences C,
language. For example, we should want Pα [A ∣ C] = Pα [C ∣ C] (i.e., Pα [A ∣ C] = 1) .

Pα [George is not married ∣ George is a bachelor] = 1, Here is how axiom 6 applies to the above example, yielding
Pα [∼Mg ∣ Bg] = 1 when the meaning assignments to non-logical terms
given the usual meanings of ‘bachelor’ and ‘married’, since “all bachelors associated with support function Pα makes ∀x(Bx ⊃ ∼Mx) analytically
are unmarried” is analytically true—i.e. no empirical evidence is required true. From axiom 6 (followed by results 7, 5, and 4) we have
to establish this connection. (In the formal language for predicate logic, if
we associate the meaning “is married” with predicate term ‘M’, the 1 = Pα [∀x(Bx ⊃ ∼Mx) ∣ Bg]
meaning “is a bachelor” with the predicate term ‘B’, and take the name = Pα [(Bg ⋅ ∀x(Bx ⊃ ∼Mx)) ∣ Bg]
term ‘g’ to refer to George, then we should want Pα [∼Mg ∣ Bg] = 1 , since ≤ Pα [∼Mg ∣ Bg]
∀x(Bx ⊃ ∼Mx) is analytically true on this meaning assignment to the non- ≤ 1;
logical terms.) So, let’s associate with each individual support function Pα
thus, Pα [∼Mg ∣ Bg] = 1 . The idea behind axiom 6 is that inductive logic
a specific assignment of meanings (primary intensions) to all the non-
is about evidential support for contingent claims. Nothing can count as
logical terms of the language. (However, evidential support functions
empirical evidence for or against non-contingent truths. In particular,
should not presuppose meaning assignments in the sense of so-called
analytic truths should be maximally supported by all premises C.
secondary intensions—e.g., those associated with rigid designators across
possible states of affairs. For, we should not want a confirmation function One important respect in which inductive logic should follow the
Pα to make deductive paradigm is that the logic should not presuppose the truth of

Pα [This glass is full of H 2O ∣ This glass is full of water] = 1,


contingent statements. If a statement C is contingent, then some other
statements should be able to count as evidence against C. Otherwise, a
since we presumably want the inductive logic to draw on explicit support function Pα will take C and all of its logical consequences to be
empirical evidence to support the claim that water is made of H2O. Thus, supported to degree 1 by all possible evidence claims. This is no way for
the meanings of terms we associate with a support function should only be an inductive logic to behave. The whole idea of inductive logic is to
their primary intensions, not their secondary intensions.) provide a measure of the extent to which premise statements indicate the
likely truth-values of contingent conclusion statements. This idea won’t
In the context of inductive logic it makes good sense to supplement the work properly if the truth-values of some contingent statements are
above axioms with two additional axioms. Here is the first of them: presupposed by assigning them support value 1 on every possible premise.

20 Stanford Encyclopedia of Philosophy Spring 2021 Edition 21


Inductive Logic James Hawthorne

Such probability assignments would make the inductive logic Axioms 6 and 7 taken together say that a support function Pα counts as
enthymematic by hiding significant premises in inductive support non-contingently true, and so not subject to empirical support, just those
relationships. It would be analogous to permitting deductive arguments to sentences that are assigned probability 1 by every premise.
count as valid in cases where the explicitly stated premises are insufficient
to logically entail the conclusion, but where the validity of the argument is Some Bayesian logicists have proposed that an inductive logic might be
permitted to depend on additional unstated premises. This is not how a made to depend solely on the logical form of sentences, as is the case for
rigorous approach to deductive logic should work, and it should not be a deductive logic. The idea is, effectively, to supplement axioms 1–7 with
common practice in a rigorous approach to inductive logic. additional axioms that depend only on the logical structures of sentences,
and to introduce enough such axioms to reduce the number of possible
Nevertheless, it is common practice for probabilistic logicians to sweep support functions to a single uniquely best support function. It is now
provisionally accepted contingent claims under the rug by assigning them widely agreed that this project cannot be carried out in a plausible way.
probability 1 (regardless of the fact that no explicit evidence for them is Perhaps support functions should obey some rules in addition to axioms
provided). This practice saves the trouble of repeatedly writing a given 1–7. But it is doubtful that any plausible collection of additional rules can
contingent sentence B as a premise, since Pγ [A ∣ B ⋅ C] will equal suffice to determine a single, uniquely qualified support function. Later, in
Pγ [A ∣ C] whenever Pγ [B ∣ C] = 1. Although this convention is useful, Section 3, we will briefly return to this issue, after we develop a more
such probability functions should be considered mere abbreviations for detailed account of how inductive probabilities capture the relationship
proper, logically explicit, non-enthymematic, inductive support relations. between hypotheses and evidence.
Thus, properly speaking, an inductive support function Pα should not
assign probability 1 to a sentence on every possible premise unless that 2.3 Two Conceptions of Inductive Probability
sentence is either (i) logically true, or (ii) an axiom of set theory or some
other piece of pure mathematics employed by the sciences, or (iii) unless Axioms 1–7 for conditional probability functions merely place formal
according to the interpretation of the language that Pα presupposes, the constraints on what may properly count as a degree of support function.
sentence is analytic (and so outside the realm of evidential support). Thus, Each function Pα that satisfies these axioms may be viewed as a possible
we adopt the following version of the so-called “axiom of regularity”. way of applying the notion of inductive support to a language L that
respects the meanings of the logical terms, much as each possible truth-
(7) If, for all C, Pα [A ∣ C] = Pα [C ∣ C] (i.e., Pα [A ∣ C] = 1 ), then A value assignment for a language represents a possible way of assigning
must be a logical truth or an axiom of set theory or some other piece truth-values to its sentences in a way that respects the meanings of the
of pure mathematics employed by the sciences, or A must be logical terms. The issue of which of the possible truth-value assignments
analytically true (according to the meanings of the terms of L to a language represents the actual truth or falsehood of its sentences
associated with support function Pα ). depends on more than this. It depends on the meanings of the non-logical
terms and on the state of the actual world. Similarly, the degree to which
some sentences actually support others in a fully meaningful language

22 Stanford Encyclopedia of Philosophy Spring 2021 Edition 23


Inductive Logic James Hawthorne

must rely on something more than the mere satisfaction of the axioms for tie such belief strengths to how much money (or how many units of utility)
support functions. It must, at least, rely on what the sentences of the the agent would be willing to bet on A turning out to be true. Roughly, the
language mean, and perhaps on much more besides. But, what more? idea is this. Suppose that an ideally rational agent α would be willing to
Perhaps a better understanding of what inductive probability is may accept a wager that would yield (no less than) $u if A turns out to be true
provide some help by filling out our conception of what inductive support and would lose him $1 if A turns out to be false. Then, under reasonable
is about. Let’s pause to discuss two prominent views—two interpretations assumptions about the agent’s desire money, it can be shown that the
of the notion of inductive probability. agent’s belief strength that A is true should be

1
One kind of non-syntactic logicist reading of inductive probability takes Pα [A] = .
each support function Pα to be a measure on possible states of affairs. The (u + 1)
idea is that, given a fully meaningful language (associated with support
And it can further be shown that any function Pα that expresses such
function Pα ) ‘Pα [A ∣ B] = r’ says that among those states of affairs in
betting-related belief-strengths on all statements in agent α ’s language
which B is true, A is true in proportion r of them. There will not generally
must satisfy axioms for unconditional probabilities analogous to axioms
be a single privileged way to define such a measure on possible states of
1–5.[4] Moreover, it can be shown that any function Pβ that satisfies these
affairs. Rather, each of a number of functions Pα , Pβ , Pγ ,…, etc., that
axioms is a possible rational belief function for some ideally rational agent
satisfy the constraints imposed by axioms 1–7 may represent a viable
β. These relationships between belief-strengths and the desirability of
measure of the inferential import of the propositions expressed by
outcomes (e.g., gaining money or goods on bets) are at the core of
sentences of the language. This idea needs more fleshing out, of course.
subjectivist Bayesian decision theory. Subjectivist Bayesians usually take
The next section will provide some indication of how that might go.
inductive probability to just be this notion of probabilistic belief-strength.
Subjectivist Bayesians offer an alternative reading of the support
Undoubtedly real agents do believe some claims more strongly than
functions. First, they usually take unconditional probability as basic, and
others. And, arguably, the belief strengths of real agents can be measured
take conditional probabilities as defined in terms of unconditional
on a probabilistic scale between 0 and 1, at least approximately. And
probabilities: the conditional probability ‘Pα [A ∣ B]’ is defined as a ratio of
clearly the inductive support of a hypothesis by evidence should influence
unconditional probabilities:
the strength of an agent’s belief in the truth of that hypothesis—that’s the
Pα [A ⋅ B] point of engaging in inductive reasoning, isn’t it? However, there is good
Pα [A ∣ B] = .
Pα [B] reason for caution about viewing inductive support functions as Bayesian
belief-strength functions, as we’ll see a bit later. So, perhaps an agent’s
Subjectivist Bayesians take each unconditional probability function Pα to support function is not simply identical to his belief function, and perhaps
represent the belief-strengths or confidence-strengths of an ideally rational the relationship between inductive support and belief-strength is somewhat
agent, α . On this understanding ‘Pα [A] = r’ says, “the strength of α’s more complicated.
belief (or confidence) that A is truth is r”. Subjectivist Bayesians usually

24 Stanford Encyclopedia of Philosophy Spring 2021 Edition 25


Inductive Logic James Hawthorne

In any case, some account of what support functions are supposed to theory) differ in empirical import, they count as distinct hypotheses. (This
represent is clearly needed. The belief function account and the logicist should not be confused with the converse positivistic assertion that
account (in terms of measures on possible states of affairs) are two theories with the same empirical content are really the same theory.
attempts to provide this account. But let us put this interpretative issue Inductive logic doesn’t necessarily endorse that view.)
aside for now. One may be able to get a better handle on what inductive
support functions really are after one sees how the inductive logic that The collection of competing hypotheses (or theories) to be evaluated by
draws on them is supposed to work. the logic may be finite in number, or may be countably infinite. No
realistic language contains more than a countable number of expressions;
3. The Application of Inductive Probabilities to the so it suffices for a logic to apply to countably infinite number of sentences.
From a purely logical perspective the collection of competing alternatives
Evaluation of Scientific Hypotheses
may consist of every rival hypothesis (or theory) about a given subject
One of the most important applications of an inductive logic is its matter that can be expressed within a given language — e.g., all possible
treatment of the evidential evaluation of scientific hypotheses. The logic theories of the origin and evolution of the universe expressible in English
should capture the structure of evidential support for all sorts of scientific and contemporary mathematics. In practice, alternative hypotheses (or
hypotheses, ranging from simple diagnostic claims (e.g., “the patient is theories) will often be constructed and evidentially evaluated over a long
infected by the HIV”) to complex scientific theories about the fundamental period of time. The logic of evidential support works in much the same
nature of the world, such as quantum mechanics or the theory of relativity. way regardless of whether all alternative hypotheses are considered
This section will show how evidential support functions (a.k.a. Bayesian together, or only a few alternative hypotheses are available at a time.
confirmation functions) represent the evidential evaluation of scientific
Evidence for scientific hypotheses consists of the results of specific
hypotheses and theories. This logic is essentially comparative. The
experiments or observations. For a given experiment or observation, let ‘c ’
evaluation of a hypothesis depends on how strongly evidence supports it
represent a description of the relevant conditions under which it is
over alternative hypotheses.
performed, and let ‘e’ represent a description of the result of the
Consider some collection of mutually incompatible, alternative hypotheses experiment or observation, the evidential outcome of conditions c .
(or theories) about a common subject matter, {h1 , h2 , …}. The collection
The logical connection between scientific hypotheses and the evidence
of alternatives may be very simple, e.g., {“the patient has HIV”, “the
often requires the mediation of background information and auxiliary
patient is free of HIV”}. Or, when the physician is trying to determine
hypotheses. Let ‘b ’ represent whatever background and auxiliary
which among a range of diseases is causing the patient’s symptoms, the
hypotheses are required to connect each hypothesis hi among the
competing hypotheses {h1 , h2 , …} to the evidence. Although the claims
collection of alternatives may consist of a long list of possible disease
hypotheses. For the cosmologist, the collection of alternatives may consist
expressed by the auxiliary hypotheses within b may themselves be subject
of several distinct gravitational theories, or several empirically distinct
to empirical evaluation, they should be the kinds of claims that are not at
variants of the “same” theory. Whenever two variants of a hypothesis (or

26 Stanford Encyclopedia of Philosophy Spring 2021 Edition 27


Inductive Logic James Hawthorne

issue in the evaluation of the alternative hypothesis in the collection Bayesian account of evidential support we will be describing below
{h1 , h2 , …}. Rather, each of the alternative hypotheses under extends this deductivist approach to include cases where the hypothesis hi
consideration draws on the same background and auxiliaries to logically (and its alternatives) may not be deductive related to the evidence, but
connect to the evidential events. (If competing hypotheses hi and hj draw may instead imply that the evidential outcome is likely or unlikely to some
on distinct auxiliary hypotheses ai and aj , respectively, in making logical specific degree r. That is, the Bayesian approach applies to cases where we
contact with evidential claims, then the following treatment should be may have neither hi ⋅ b ⋅ c ⊨ e nor hi ⋅ b ⋅ c ⊨ ∼e, but may instead only
applied to the respective conjunctive hypotheses, (hi ⋅ ai ) and (hj ⋅ aj ), have P[e ∣ hi ⋅ b ⋅ c] = r , where r is some “entailment strength” between
since these alternative conjunctive hypotheses will constitute the 0 and 1.
empirically distinct alternatives at issue.)
Before going on to describing the logic of evidential support in more
In cases where a hypothesis is deductively related to an outcome e of an detail, perhaps a few more words are in order about the background
observational or experimental condition c (via background and auxiliaries knowledge and auxiliary hypotheses, represented here by ‘b’. Duhem
b), we will have either hi ⋅ b ⋅ c ⊨ e or hi ⋅ b ⋅ c ⊨ ∼e. For example, hi (1906) and Quine (1953) are generally credited with alerting inductive
might be the Newtonian Theory of Gravitation. A test of the theory might logicians to the importance of auxiliary hypotheses in connecting scientific
involve a condition statement c that describes the results of some earlier hypotheses and theories to empirical evidence. (See the entry on Pierre
measurements of Jupiter’s position, and that describes the means by which Duhem.) They point out that scientific hypotheses often make little contact
the next position measurement will be made; the outcome description e with evidence claims on their own. Rather, in most cases scientific
states the result of this additional position measurement; and the hypotheses make testable predictions only relative to background
background information (and auxiliary hypotheses) b might state some information and auxiliary hypotheses that tie them to the evidence. (Some
already well confirmed theory about the workings and accuracy of the specific examples of such auxiliary hypotheses will be provided in the
devices used to make the position measurements. Then, from hi ⋅ b ⋅ c we next subsection.) Typically auxiliaries are highly confirmed hypotheses
may calculate the specific outcome e we expect to find; thus, the following from other scientific domains. They often describe the operating
logical entailment holds: hi ⋅ b ⋅ c ⊨ e. Then, provided that the characteristics of various devices (e.g., measuring instruments) used to
experimental and observational conditions stated by c are in fact true, if make observations or conduct experiments. Their credibility is usually not
the evidential outcome described by e actually occurs, the resulting at issue in the testing of hypothesis hi against its competitors, because hi
conjoint evidential claim (c ⋅ e) may be considered good evidence for hi , and its alternatives usually rely on the same auxiliary hypotheses to tie
given b . (This method of theory evaluation is called the hypothetical- them to the evidence. But even when an auxiliary hypothesis is already
deductive approach to evidential support.) On the other hand, when from well-confirmed, we cannot simply assume that it is unproblematic, or just
hi ⋅ b ⋅ c we calculate some outcome incompatible with the observed known to be true. Rather, the evidential support or refutation of a
evidential outcome e, then the following logical entailment holds: hypothesis hi is relative to whatever auxiliaries and background
hi ⋅ b ⋅ c ⊨ ∼e. In that case, from deductive logic alone we must also have information (in b) is being supposed in the confirmational context. In other
that b ⋅ c ⋅ e ⊨ ∼hi ; thus, hi is said to be falsified by b ⋅ c ⋅ e . The contexts the auxiliary hypotheses used to test hi may themselves be among

28 Stanford Encyclopedia of Philosophy Spring 2021 Edition 29


Inductive Logic James Hawthorne

a collection of alternative hypotheses that are subject to evidential support 3.1 Likelihoods
or refutation. Furthermore, to the extent that competing hypotheses
employ different auxiliary hypotheses in accounting for evidence, the In probabilistic inductive logic the likelihoods carry the empirical import
evidence only tests each such hypothesis in conjunction with its distinct of hypotheses. A likelihood is a support function probability of form
auxiliaries against alternative hypotheses packaged with their distinct P[e ∣ hi ⋅ b ⋅ c]. It expresses how likely it is that outcome e will occur
auxiliaries, as described earlier. Thus, what counts as a hypothesis to be according to hypothesis hi together with the background and auxiliaries b
tested, hi , and what counts as auxiliary hypotheses and background and the experimental (or observational) conditions c.[5] If a hypothesis
information, b , may depend on the epistemic context—on what class of together with auxiliaries and experimental/observation conditions
alternative hypotheses are being tested by a collection of experiments or deductively entails an evidence claim, the axioms of probability make the
observations, and on what claims are presupposed in that context. No corresponding likelihood objective in the sense that every support function
statement is intrinsically a test hypothesis, or intrinsically an auxiliary must agree on its values: P[e ∣ hi ⋅ b ⋅ c] = 1 if hi ⋅ b ⋅ c ⊨ e;
hypothesis or background condition. Rather, these categories are roles P[e ∣ hi ⋅ b ⋅ c] = 0 if hi ⋅ b ⋅ c ⊨ ∼e. However, in many cases a
statements may play in a particular epistemic context. hypothesis hi will not be deductively related to the evidence, but will only
imply it probabilistically. There are several ways this might happen: (1)
In a probabilistic inductive logic the degree to which the evidence (c ⋅ e) hypothesis hi may itself be an explicitly probabilistic or statistical
supports a hypothesis hi relative to background and auxiliaries b is hypothesis; (2) an auxiliary statistical hypothesis, as part of the
represented by the posterior probability of hi , Pα [hi ∣ b ⋅ c ⋅ e], according background b, may connect hypothesis hi to the evidence; (3) the
to an evidential support function Pα . It turns out that the posterior connection between the hypothesis and the evidence may be somewhat
probability of a hypothesis depends on just two kinds of factors: (1) its loose or imprecise, not mediated by explicit statistical claims, but
prior probability, Pα [hi ∣ b] , together with the prior probabilities of its nevertheless objective enough for the purposes of evidential evaluation.
competitors, Pα [hj ∣ b] , Pα [hk ∣ b] , etc.; and (2) the likelihood of Let’s briefly consider examples of the first two kinds. We’ll treat case (3)
evidential outcomes e according to hi in conjunction with with b and c, in Section 5, which addresses the the issue of vague and imprecise
P[e ∣ hi ⋅ b ⋅ c], together with the likelihoods of these same evidential likelihoods.
outcomes according to competing hypotheses, P[e ∣ hj ⋅ b ⋅ c] ,
P[e ∣ hk ⋅ b ⋅ c], etc. We will now examine each of these factors in some The hypotheses being tested may themselves be statistical in nature. One
detail. Following that we will see precisely how the values of posterior of the simplest examples of statistical hypotheses and their role in
probabilities depend on the values of likelihoods and prior probabilities. likelihoods are hypotheses about the chance characteristic of coin-tossing.
Let h[r] be a hypothesis that says a specific coin has a propensity (or
objective chance) r for coming up heads on normal tosses, let b say that
such tosses are probabilistically independent of one another. Let c state
that the coin is tossed n times in the normal way; and let e say that on
these tosses the coin comes up heads m times. In cases like this the value

30 Stanford Encyclopedia of Philosophy Spring 2021 Edition 31


Inductive Logic James Hawthorne

of the likelihood of the outcome e on hypothesis h[r] for condition c is patient is infected with HIV, h, and this patient is not infected with HIV,
given by the well-known binomial formula: ∼h. In this context the known test characteristics function as background
information, b. The experimental condition c merely states that this
n!
P[e ∣ h[r] ⋅ b ⋅ c] = × rm (1 − r)n−m . particular patient was subjected to this specific kind of blood test for HIV,
m! × (n − m)!
which was processed by the lab using proper procedures. Let us suppose
There are, of course, more complex cases of likelihoods involving that the outcome e states that the result is a positive test result for HIV.
statistical hypotheses. Consider, for example, the hypothesis that The relevant likelihoods then, are P[e ∣ h ⋅ b ⋅ c] = .99 and
plutonium 233 nuclei have a half-life of 20 minutes—i.e., that the P[e ∣ ∼h ⋅ b ⋅ c] = .05.
propensity (or objective chance) for a Pu-233 nucleus to decay within a 20
In this example the values of the likelihoods are entirely due to the
minute period is 1/2. The full statistical model for the lifetime of such a
statistical characteristics of the accuracy of the test, which is carried by the
system says that the propensity (or objective chance) for that system to
background/auxiliary information b . The hypothesis h being tested by the
remain intact (i.e., to not decay) within any time period x is governed by
the formula 1/2x/τ , where τ is the half-life of such a system. Let h be a
evidence is not itself statistical.

hypothesis that says that this statistical model applies to Pu-233 nuclei This kind of situation may, of course, arise for much more complex
with τ = 20 minutes; let c say that some specific Pu-233 nucleus is intact hypotheses. The alternative hypotheses of interest may be deterministic
within a decay detector (of some specific kind) at an initial time t0 ; let e physical theories, say Newtonian Gravitation Theory and some specific
say that no decay of this same Pu-233 nucleus is detected by the later time alternatives. Some of the experiments that test this theory relay on
t; and let b say that the detector is completely accurate (it always registers somewhat imprecise measurements that have known statistical error
a real decay, and it never registers false-positive detections). Then, the characteristics, which are expressed as part of the background or auxiliary
associated likelihood of e given h and c is this: hypotheses, b. For example, the auxiliary b may describe the error
P[e ∣ h ⋅ b ⋅ c] = 1/2 (t−t0 )/τ
, where the value of τ is 20 minutes. characteristics of a device that measures the torque imparted to a quartz
fiber, where the measured torque is used to assess the strength of the
An auxiliary statistical hypothesis, as part of the background b, may be
gravitational force between test masses. In that case b may say that for this
required to connect hypothesis hi to the evidence. For example, a blood
kind of device the measurement errors are normally distributed about
test for HIV has a known false-positive rate and a known true-positive
whatever value a given gravitational theory predicts, with some specified
rate. Suppose the false-positive rate is .05—i.e., the test tends to
standard deviation that is characteristic of the device. This results in
incorrectly show the blood sample to be positive for HIV in 5% of all
specific values ri for the likelihoods, P[e ∣ hi ⋅ b ⋅ c] = ri , for each of the
cases where HIV is not present. And suppose that the true-positive rate is
various gravitational theories, hi , being tested.
.99—i.e., the test tends to correctly show the blood sample to be positive
for HIV in 99% of all cases where HIV really is present. When a particular Likelihoods that arise from explicit statistical claims—either within the
patient’s blood is tested, the hypotheses under consideration are this hypotheses being tested, or from explicit statistical background claims that

32 Stanford Encyclopedia of Philosophy Spring 2021 Edition 33


Inductive Logic James Hawthorne

tie the hypotheses to the evidence—are often called direct inference relies on a high degree of objectivity or intersubjective agreement among
likelihoods. Such likelihoods should be completely objective. So, all scientists on the numerical values of likelihoods.
evidential support functions should agree on their values, just as all
support functions agree on likelihoods when evidence is logically entailed. To see the point more vividly, imagine what a science would be like if
Direct inference likelihoods are logical in an extended, non-deductive scientists disagreed widely about the values of likelihoods. Each
sense. Indeed, some logicians have attempted to spell out the logic of practitioner interprets a theory to say quite different things about how
direct inferences in terms of the logical form of the sentences involved.[6] likely it is that various possible evidence statements will turn out to be
But regardless of whether that project succeeds, it seems reasonable to true. Whereas scientist α takes theory h1 to probabilistically imply that
take likelihoods of this sort to have highly objective or intersubjectively event e is highly likely, his colleague β understands the empirical import
agreed values. of h1 to say that e is very unlikely. And, conversely, α takes competing
theory h2 to probabilistically imply that e is very unlikely, whereas β reads
Not all likelihoods of interest in confirmational contexts are warranted h2 to say that e is extremely likely. So, for α the evidential outcome e
deductively or by explicitly stated statistical claims. In such cases the supplies strong support for h1 over h2 , because
likelihoods may have vague, imprecise values, but values that are
determinate enough to still underwrite an objective evaluation of Pα [e ∣ h1 ⋅ b ⋅ c] ≫ Pα [e ∣ h2 ⋅ b ⋅ c].
hypotheses on the evidence. In Section 5 we’ll consider such cases, where
But his colleague β takes outcome e to show just the opposite, that h2 is
no underlying statistical theory is involved, but where likelihoods are
strongly supported over h1 , because
determinate enough to play their standard role in the evidential evaluation
of scientific hypotheses. However, the proper treatment of such cases will Pβ [e ∣ h2 ⋅ b ⋅ c] ≫ Pβ [e ∣ h1 ⋅ b ⋅ c].
be more easily understood after we have first seen how the logic works
when likelihoods are precisely known (such as cases where the likelihood If this kind of situation were to occur often, or for significant evidence
values are endorsed by explicit statistical hypotheses and/or explicit claims in a scientific domain, it would make a shambles of the empirical
statistical auxiliaries). In any case, the likelihoods that relate hypotheses to objectivity of that science. It would completely undermine the empirical
evidence claims in many scientific contexts will have such objective testability of such hypotheses and theories within that scientific domain.
values. So, although a variety of different support functions Pα , Pβ ,…, Pγ , Under these circumstances, although each scientist employs the same
etc., may be needed to represent the differing “inductive proclivities” of sentences to express a given theory hi , each understands the empirical
the various members of a scientific community, for now we will consider import of these sentences so differently that hi as understood by α is an
cases where all evidential support functions agree on the values of the empirically different theory than hi as understood by β. (Indeed, arguably,
likelihoods. For, the likelihoods represent the empirical content of a α must take at least one of the two sentences, h1 or h2 , to express a
scientific hypothesis, what the hypothesis (together with experimental different proposition than does β.) Thus, the empirical objectivity of the
conditions, c , and background and auxiliaries b) says or probabilistically sciences requires that experts should be in close agreement about the
implies about the evidence. Thus, the empirical objectivity of a science values of the likelihoods.[7]

34 Stanford Encyclopedia of Philosophy Spring 2021 Edition 35


Inductive Logic James Hawthorne

For now we will suppose that the likelihoods have objective or likelihood of the evidence stream will be equal to the product of the
intersubjectively agreed values, common to all agents in a scientific likelihoods of the individual outcomes:
community. We mark this agreement by dropping the subscript ‘α’, ‘β’,
etc., from expressions that represent likelihoods, since all support P[en ∣ hi ⋅ b ⋅ cn ] = P[e1 ∣ hi ⋅ b ⋅ c1 ] × ⋯ × P[ en ∣ hi ⋅ b ⋅ cn ].
functions under consideration are supposed to agree on the values for
When this equality holds, the individual bits of evidence are said to be
likelihoods. One might worry that this supposition is overly strong. There
probabilistically independent on the hypothesis (together with auxiliaries).
are legitimate scientific contexts where, although scientists should have
In the following account of the logic of evidential support, such
enough of a common understanding of the empirical import of hypotheses
probabilistic independence will not be assumed, except in those places
to assign quite similar values to likelihoods, precise agreement on their
where it is explicitly invoked.
numerical values may be unrealistic. This point is right in some important
kinds of cases. So later, in Section 5, we will see how to relax the
3.2 Posterior Probabilities and Prior Probabilities
supposition that precise likelihood values are available, and see how the
logic works in such cases. But for now the main ideas underlying The probabilistic logic of evidential support represents the net support of a
probabilistic inductive logic will be more easily explained if we focus on hypothesis by the posterior probability of the hypothesis,
those contexts were objective or intersubjectively agreed likelihoods are Pα [hi ∣ b ⋅ cn ⋅ en ] . The posterior probability represents the net support for
available. Later we will see that much the same logic continues to apply in the hypothesis that results from the evidence, cn ⋅ en , together with
contexts where the values of likelihoods may be somewhat vague, or whatever plausibility considerations are taken to be relevant to the
where members of the scientific community disagree to some extent about assessment of hi . Whereas the likelihoods are the means through which
their values. evidence contributes to the posterior probability of a hypothesis, all other
relevant plausibility consideration are represented by a separate factor,
An adequate treatment of the likelihoods calls for the introduction of one
called the prior probability of the hypothesis: Pα [hi ∣ b] . The prior
additional notational device. Scientific hypotheses are generally tested by
probability represents the weight of any important considerations not
a sequence of experiments or observations conducted over a period of
captured by the evidential likelihoods. Any relevant considerations that go
time. To explicitly represent the accumulation of evidence, let the series of
beyond the evidence itself may be explicitly stated within expression b (in
sentences c1 , c2 , …, cn , describe the conditions under which a sequence of
addition to whatever auxiliary hypotheses b may contain in support of the
experiments or observations are conducted. And let the corresponding
likelihoods). Thus, the prior probability of hi may depend explicitly on the
outcomes of these observations be represented by sentences e1 , e2 , …, en .
content of b. It turns out that posterior probabilities depend only on the
We will abbreviate the conjunction of the first n descriptions of
values of evidential likelihoods together with the values of prior
experimental or observational conditions by ‘cn ’, and abbreviate the
probabilities.
conjunction of descriptions of their outcomes by ‘en ’. Then, for a stream
of n observations or experiments and their outcomes, the likelihoods take
form P[en ∣ hi ⋅ b ⋅ cn ] = r, for appropriate values of r . In many cases the

36 Stanford Encyclopedia of Philosophy Spring 2021 Edition 37


Inductive Logic James Hawthorne

As an illustration of the role of prior probabilities, consider the HIV test More generally, in the evidential evaluation of scientific hypotheses and
example described in the previous section. What the physician and the theories, prior probabilities represent assessments of non-evidential
patient want to know is the value of the posterior probability, plausibility weightings among hypotheses. However, because the strengths
Pα [h ∣ b ⋅ c ⋅ e] , that the patient has HIV, h , given the evidence of the of such plausibility assessments may vary among members of a scientific
positive test, c ⋅ e, and given the error rates of the test, described within b. community, critics often brand such assessments as merely subjective, and
The value of this posterior probability depends on the likelihood (due to take their role in Bayesian inference to be highly problematic. Bayesian
the error rates) of this patient obtaining a true-positive result, inductivists counter that plausibility assessments play an important,
P[e ∣ h ⋅ b ⋅ c] = .99, and of obtaining a false-positive result, legitimate role in the sciences, especially when evidence cannot suffice to
P[e ∣ ∼h ⋅ b ⋅ c] = .05 . In addition, the value of the of the posterior distinguish among some alternative hypotheses. And, they argue, the
probability depends on how plausible it is that the patient has HIV prior to epithet “merely subjective” is unwarranted. Such plausibility assessments
taking the test results into account, Pα [h ∣ b]. In the context of medical are often backed by extensive arguments that may draw on forceful
diagnosis, this prior probability is usually assessed on the basis of the base conceptual considerations.
rate for HIV in the patient’s risk group (i.e., whether the patient is an IV
drug user, has unprotected sex with multiple partners, etc.). On a rigorous Scientists often bring plausibility arguments to bear in assessing
approach to the logic, such information and its risk-relevance should be competing views. Although such arguments are seldom decisive, they may
explicitly stated within the background information b. To see the bring the scientific community into widely shared agreement, especially
importance of this information, consider the following numerical results with regard to the implausibility of some logically possible alternatives.
(which may be calculated using the formula called Bayes’ Theorem, This seems to be the primary epistemic role of thought experiments.
presented in the next section). If the base rate for the patient’s risk group is Consider, for example, the kinds of plausibility arguments that have been
relatively high, say Pα [h ∣ b] = .10, then the positive test result yields a brought to bear on the various interpretations of quantum theory (e.g.,
posterior probability value for his having HIV of Pα [h ∣ b ⋅ c ⋅ e] = .69. those related to the measurement problem). These arguments go to the
However, if the patient is in a very low risk group, say Pα [h ∣ b] = .001, heart of conceptual issues that were central to the original development of
then a positive test result only raises the posterior probability of his having the theory. Many of these issues were first raised by those scientists who
an HIV infection to Pα [h ∣ b ⋅ c ⋅ e] = .02. This posterior probability is made the greatest contributions to the development of quantum theory, in
much higher than the prior probability of .001, but should not worry the their attempts to get a conceptual hold on the theory and its implications.
patient too much. This positive test result may well be due to the
Given any body of evidence, it is fairly easy to cook up a host of logically
comparatively high false-positive rate for the test, rather than to the
possible alternative hypotheses that make the evidence as probable as
presence of HIV. This sort of test, with a false-positive rate as large as .05,
desired. In particular, it is easy to cook up hypotheses that logically entail
is best used as a screening test; a positive result warrants conducting a
any given body evidence, providing likelihood values equal to 1 for all the
second, more rigorous, less error-prone test.
available evidence. Although most of these cooked up hypotheses will be
laughably implausible, evidential likelihoods cannot rule them out. But,

38 Stanford Encyclopedia of Philosophy Spring 2021 Edition 39


Inductive Logic James Hawthorne

the only factors other than likelihoods that figure into the values of plausibility values possessed by some hypotheses. In the next section we’ll
posterior probabilities for hypotheses are the values of their prior see precisely how this idea works, and we’ll return to it again in Section
probabilities; so only prior probability assessments provide a place for the 3.4.
Bayesian logic to bring important plausibility considerations to bear. Thus,
the Bayesian logic can only give implausible hypotheses their due via When sufficiently strong evidence becomes available, it turns out that the
prior probability assessments. contributions of prior plausibility assessments to the values of posterior
probabilities may be substantially “washed out”, overridden by the
It turns out that the mathematical structure of Bayesian inference makes evidence. That is, provided the prior probability of a true hypothesis isn’t
prior probabilities especially well-suited to represent plausibility assessed to be too close to zero, the influence of the values of the prior
assessments among competing hypotheses. For, in the fully fleshed out probabilities will very probably fade away as evidence accumulates. In
account of evidential support for hypotheses (spelled out below), it will Section 4 we’ll see precisely how this kind of Bayesian convergence to the
turn out that only ratios of prior probabilities for competing hypotheses, true hypothesis works. Thus, it turns out that prior plausibility assessments
Pα [hj ∣ b]/Pα [hi ∣ b], together with ratios of likelihoods, play their most important role when the distinguishing evidence
Pα [e ∣ hj ⋅ b ⋅ c]/Pα [e ∣ h2 ⋅ b ⋅ c] , play essential roles. The ratio of prior represented by the likelihoods remains weak.
probabilities is well-suited to represent how much more (or less) plausible
hypothesis hj is than competing hypothesis hi . Furthermore, the One more point before moving on to the logic of Bayes’ Theorem. Some
plausibility arguments on which such this comparative assessment is based Bayesian logicists have maintained that posterior probabilities of
may be explicitly stated within b. So, given that an inductive logic needs hypotheses should be determined by syntactic logical form alone. The idea
to incorporate well-considered plausibility assessments (e.g. in order to lay is that the likelihoods might reasonably be specified in terms of syntactic
low wildly implausible alternative hypotheses), the comparative logical form; so if syntactic form might be made to determine the values
assessment of Bayesian prior probabilities seems well-suited to do the job. of prior probabilities as well, then inductive logic would be fully “formal”
in the same way that deductive logic is “formal”. Keynes and Carnap tried
Thus, although prior probabilities may be subjective in the sense that to implement this idea through syntactic versions of the principle of
agents may disagree on the relative strengths of plausibility arguments, the indifference—the idea that syntactically similar hypotheses should be
priors used in scientific contexts need not represent mere subjective whims. assigned the same prior probability values. Carnap showed how to carry
Rather, the comparative strengths of the priors for hypotheses should be out this project in detail, but only for extremely simple formal languages.
supported by arguments about how much more plausible one hypothesis is Most logicians now take the project to have failed because of a fatal flaw
than another. The important role of plausibility assessments is captured by with the whole idea that reasonable prior probabilities can be made to
such received bits of scientific wisdom as the well-known scientific depend on logical form alone. Semantic content should matter.
aphorism, extraordinary claims require extraordinary evidence. That is, it Goodmanian grue-predicates provide one way to illustrate this point.[8]
takes especially strong evidence, in the form of extremely high values for Furthermore, as suggested earlier, for this idea to apply to the evidential
(ratios of) likelihoods, to overcome the extremely low pre-evidential support of real scientific theories, scientists would have to assess the prior

40 Stanford Encyclopedia of Philosophy Spring 2021 Edition 41


Inductive Logic James Hawthorne

probabilities of each alternative theory based only on its syntactic The simplest version of Bayes’ Theorem as it applies to evidence for a
structure. That seems an unreasonable way to proceed. Are we to evaluate hypothesis goes like this:
the prior probabilities of alternative theories of gravitation, or for
alternative quantum theories, by exploring only their syntactic structures, Bayes’ Theorem: Simple Form
with absolutely no regard for their content—with no regard for what they Pα [e ∣ hi ] × Pα [hi ]
Pα [hi ∣ e] =
say about the world? This seems an extremely dubious approach to the Pα [e]
evaluation of real scientific theories. Logical structure alone cannot, and
should not suffice for determining reasonable prior probability values for This equation expresses the posterior probability of hypothesis hi due to
real scientific theories. Moreover, real scientific hypotheses and theories evidence e , Pα [hi ∣ e], in terms of the likelihood of the evidence on that
are inevitably subject to plausibility considerations based on what they say hypothesis, Pα [e ∣ hi ], the prior probability of the hypothesis, Pα [hi ], and
about the world. Prior probabilities are well-suited to represent the the simple probability of the evidence, Pα [e]. The factor Pα [e] is often
comparative weight of plausibility considerations for alternative called the expectedness of the evidence. Written this way, the theorem
hypotheses. But no reasonable assessment of comparative plausibility can suppresses the experimental (or observational) conditions, c , and all
derive solely from the logical form of hypotheses. background information and auxiliary hypotheses, b. As discussed earlier,
both of these terms play an important role in logically connecting the
We will return to a discussion of prior probabilities a bit later. Let’s now hypothesis at issue, hi , to the evidence e . In scientific contexts the
see how Bayesian logic combines likelihoods with prior probabilities to objectivity of the likelihoods, Pα [e ∣ hi ⋅ b ⋅ c], almost always depends on
yield posterior probabilities for hypotheses. such terms. So, although the suppression of experimental (or
observational) conditions and auxiliary hypotheses is a common practice
3.3 Bayes’ Theorem in accounts of Bayesian inference, the treatment below, and throughout the
remainder of this article will make the role of these terms explicit.
Any probabilistic inductive logic that draws on the usual rules of
probability theory to represent how evidence supports hypotheses must be The subscript α on the evidential support function Pα is there to remind us
a Bayesian inductive logic in the broad sense. For, Bayes’ Theorem that more than one such function exists. A host of distinct probability
follows directly from the usual axioms of probability theory. Its functions satisfy axioms 1–5, so each of them satisfies Bayes’ Theorem.
importance derives from the relationship it expresses between hypotheses Some of these probability functions may provide a better fit with our
and evidence. It shows how evidence, via the likelihoods, combines with intuitive conception of how the evidential support for hypotheses should
prior probabilities to produce posterior probabilities for hypotheses. We work. Nevertheless, there are bound to be reasonable differences among
now examine several forms of Bayes’ Theorem, each derivable from Bayesian agents regarding to the initial plausibility of a hypothesis hi .
axioms 1–5. This diversity in initial plausibility assessments is represented by diverse
values for prior probabilities for the hypothesis: Pα [hi ], Pβ [hi ], Pγ [hi ], etc.
This usually results in diverse values for posterior probabilities for

42 Stanford Encyclopedia of Philosophy Spring 2021 Edition 43


Inductive Logic James Hawthorne

hypotheses: Pα [hi ∣ e], Pβ [hi ∣ e], Pγ [hi ∣ e], etc. So it is important to keep auxiliaries) objectively says about the likelihood of possible evidential
the diversity among evidential support functions in mind. outcomes of the experimental conditions. So, all reasonable support
functions should agree on the values for likelihoods. (Section 5 will treat
Here is how the Simple Form of Bayes’ Theorem looks when terms for the cases where the likelihoods may lack this kind of objectivity.)
experimental (or observational) conditions, c , and the background
information and auxiliary hypotheses b are made explicit: This version of Bayes’ Theorem includes a term that represents the ratio of
the likelihood of the experimental conditions on the hypothesis and
Bayes’ Theorem: Simple Form with explicit Experimental background information (and auxiliaries) to the “likelihood” of the
Conditions, Background Information and Auxiliary Hypotheses experimental conditions on the background (and auxiliaries) alone:
P[e ∣ hi ⋅ b ⋅ c] × Pα [hi ∣ b] Pα [c ∣ hi ⋅ b]/Pα [c ∣ b]. Arguably the value of this term should be 1, or
(8) Pα [hi ∣ b ⋅ c ⋅ e] =
Pα [e ∣ b ⋅ c] very nearly 1, since the truth of the hypothesis at issue should not
Pα [c ∣ hi ⋅ b] significantly affect how likely it is that the experimental conditions are
×
Pα [c ∣ b] satisfied. If various alternative hypotheses assign significantly different
likelihoods to the experimental conditions themselves, then such
P[e ∣ hi ⋅ b ⋅ c] × Pα [hi ∣ b] conditions should more properly be included as part of the evidential
=
Pα [e ∣ b ⋅ c] outcome e.

when Pα [c ∣ hj ⋅ b] = Pα [c ∣ b]. Both the prior probability of the hypothesis and the expectedness tend to
be somewhat subjective factors in that various agents from the same
This version of the theorem determines the posterior probability of the scientific community may legitimately disagree on what values these
hypothesis, Pα [hi ∣ b ⋅ c ⋅ e], from the value of the likelihood of the factors should take. Bayesian logicians usually accept the apparent
evidence according to that hypothesis (taken together with background subjectivity of the prior probabilities of hypotheses, but find the
and auxiliaries and the experimental conditions), P[e ∣ hi ⋅ b ⋅ c], the value subjectivity of the expectedness to be more troubling. This is due at least
of the prior probability of the hypothesis (on background and auxiliaries), in part to the fact that in a Bayesian logic of evidential support the value of
Pα [hi ∣ b] , and the value of the expectedness of the evidence (on the expectedness cannot be determined independently of likelihoods and
background and auxiliaries and the experimental conditions), Pα [e ∣ b ⋅ c]. prior probabilities of hypotheses. That is, when, for each member of a
Notice that in the factor for the likelihood, P[e ∣ hi ⋅ b ⋅ c], the subscript α collection of alternative hypotheses, the likelihood P[e ∣ hj ⋅ b ⋅ c] has an
has been dropped. This marks the fact that in scientific contexts the objective (or intersubjectively agreed) value, the expectedness is
likelihood of an evidential outcome e on the hypothesis together with constrained by the following equation (where the sum ranges over a
explicit background and auxiliary hypotheses and the description of the mutually exclusive and exhaustive collection of alternative hypotheses
experimental conditions, hi ⋅ b ⋅ c, is usually objectively determinate. This {h1 , h2 , … , hm , …} , which may be finite or countably infinite):
factor represents what the hypothesis (in conjunction with background and

= ∑ P[e ∣ j ⋅ b ⋅ c] × α[ j ∣ b ⋅ c].
44 Stanford Encyclopedia of Philosophy Spring 2021 Edition 45
Inductive Logic James Hawthorne

Pα [e ∣ b ⋅ c] = ∑ P[e ∣ hj ⋅ b ⋅ c] × Pα [hj ∣ b ⋅ c]. — i.e., the experimental or observation conditions are no more likely
j according to one hypothesis than according to the other.[9]

This equation shows that the values for the prior probabilities together This Ratio Form of Bayes’ Theorem expresses how much more plausible,
with the values of the likelihoods uniquely determine the value for the on the evidence, one hypothesis is than another. Notice that the likelihood
expectedness of the evidence. Furthermore, it implies that the value of the ratios carry the full import of the evidence. The evidence influences the
expectedness must lie between the largest and smallest of the various evaluation of hypotheses in no other way. The only other factor that
likelihood values implied by the alternative hypotheses. However, the influences the value of the ratio of posterior probabilities is the ratio of the
precise value of the expectedness can only be calculated this way when prior probabilities. When the likelihoods are fully objective, any
every alternative to hypothesis hj is specified. In cases where some subjectivity that affects the ratio of posteriors can only arise via
alternative hypotheses remain unspecified (or undiscovered), the value of subjectivity in the ratio of the priors.
the expectedness is constrained in principle by the totality of possible
alternative hypotheses, but there is no way to figure out precisely what its This version of Bayes’s Theorem shows that in order to evaluate the
value should be. posterior probability ratios for pairs of hypotheses, the prior probabilities
of hypotheses need not be evaluated absolutely; only their ratios are
Troubles with determining a numerical value for the expectedness of the needed. That is, with regard to the priors, the Bayesian evaluation of
evidence may be circumvented by appealing to another form of Bayes’ hypotheses only relies on how much more plausible one hypothesis is than
Theorem, a ratio form that compares hypotheses one pair at a time: another (due to considerations expressed within b). This kind of Bayesian
evaluation of hypotheses is essentially comparative in that only ratios of
Bayes’ Theorem: Ratio Form
likelihoods and ratios of prior probabilities are ever really needed for the
Pα [hj ∣ b ⋅ c ⋅ e] P[e ∣ hj ⋅ b ⋅ c] Pα [hj ∣ b] assessment of scientific hypotheses. Furthermore, we will soon see that the
(9) = ×
Pα [hi ∣ b ⋅ c ⋅ e] P[e ∣ hi ⋅ b ⋅ c] Pα [hi ∣ b] absolute values of the posterior probabilities of hypotheses entirely derive
Pα [c ∣ hj ⋅ b] from the posterior probability ratios provided by the Ratio Form of Bayes’
×
Pα [c ∣ hi ⋅ b] Theorem.

P[e ∣ hj ⋅ b ⋅ c] Pα [hj ∣ b] When the evidence consists of a collection of n distinct experiments or


= ×
P[e ∣ hi ⋅ b ⋅ c] Pα [hi ∣ b] observations, we may explicitly represent this fact by replacing the term ‘
c ’ by the conjunction of experimental or observational conditions,
when Pα [c ∣ hj ⋅ b] = Pα [c ∣ hi ⋅ b]. (c1 ⋅ c2 ⋅ … ⋅ cn ) , and replacing the term ‘e ’ by the conjunction of their
respective outcomes, (e1 ⋅ e2 ⋅ … ⋅ en ) . For notational convenience, let’s
The clause Pα [c ∣ hj ⋅ b] = Pα [c ∣ hi ⋅ b] says that the experimental (or
use the term ‘cn ’ to abbreviate the conjunction of n the experimental
observation) condition described by c is as likely on (hi ⋅ b) as on (hj ⋅ b)
conditions, and we use the term ‘en ’ to abbreviate the corresponding

46 Stanford Encyclopedia of Philosophy Spring 2021 Edition 47


Inductive Logic James Hawthorne

P[ n ∣ i ⋅ b ⋅ n ] α [ i ∣ b]
conjunction of n their respective outcomes. Relative to any given Let’s consider a simple example of how the Ratio Form of Bayes’
hypothesis h, the evidential outcomes of distinct experiments or Theorem applies to a collection of independent evidential events. Suppose
observations will usually be probabilistically independent of one another, we possess a warped coin and want to determine its propensity for heads
and also independent of the experimental conditions for one another. In when tossed in the usual way. Consider two hypotheses, h[p] and h[q],
that case we have: which say that the propensities for the coin to come up heads on the usual
kinds of tosses are p and q, respectively. Let cn report that the coin is
P[en ∣ h ⋅ b ⋅ cn ] = P[e1 ∣ h ⋅ b ⋅ c1 ] × ⋯ × P[en ∣ h ⋅ b ⋅ cn ]. tossed n times in the normal way, and let en report that precisely m
occurrences of heads has resulted. Supposing that the outcomes of such
When the Ratio Form of Bayes’ Theorem is extended to explicitly
tosses are probabilistically independent (asserted by b), the respective
represent the evidence as consisting of a collection of n of distinct
likelihoods take the binomial form
experiments (or observations) and their respective outcomes, it takes the
following form. n!
P[en ∣ h[r] ⋅ b ⋅ cn ] = × rm (1 − r)n−m ,
m! × (n − m)!
Bayes’ Theorem: Ratio Form for a Collection of n Distinct
Evidence Claims with r standing in for p and for q , respectively. Then, Equation 9** yields
the following formula, where the likelihood ratio is the ratio of the
Pα [hj ∣ b ⋅ cn ⋅ en ] P[en ∣ hj ⋅ b ⋅ cn ] Pα [hj ∣ b]
(9*) = × respective binomial terms:
Pα [hi ∣ b ⋅ c ⋅ e ]
n n P[e ∣ hi ⋅ b ⋅ c ]
n n Pα [hi ∣ b]
Pα [h[p] ∣ b ⋅ cn ⋅ en ] pm (1 − p)n−m Pα [h[p] ∣ b]
Pα [cn ∣ hj ⋅ b] = ×
× Pα [h[q] ∣ b ⋅ cn ⋅ en ] qm (1 − q)n−m Pα [h[q] ∣ b]
Pα [cn ∣ hi ⋅ b]
When, for instance, the coin is tossed n = 100 times and comes up heads
P[en ∣ hj ⋅ b ⋅ cn ] Pα [hj ∣ b] m = 72 times, the evidence for hypothesis h[1/2] as compared to h[3/4] is
= ×
P[e ∣ hi ⋅ b ⋅ c ]
n n Pα [hi ∣ b] given by the likelihood ratio

when Pα [cn ∣ hj ⋅ b] = Pα [cn ∣ hi ⋅ b]. P[en ∣ h[1/2] ⋅ b ⋅ cn ] [(1/2)72 (1/2)28 ]


= = .000056269.
P[e ∣ h[3/4] ⋅ b ⋅ c ]
n n [(3/4)72 (1/4)28 ]
Furthermore, when evidence claims are probabilistically independent
of one another, we have In that case, even if the prior plausibility considerations (expressed within
b) make it 100 times more plausible that the coin is fair than that it is
Pα [hj ∣ b ⋅ cn ⋅ en ] P[e1 ∣ hj ⋅ b ⋅ c1 ]
(9**) = ×⋯
Pα [hi ∣ b ⋅ c ⋅ e ] P[e1 ∣ hi ⋅ b ⋅ c1 ]
n n warped towards heads with propensity 3/4 — i.e., even if
Pα [h[1/2] ∣ b]/Pα [h[3/4] ∣ b] = 100 — the evidence provided by these
P[en ∣ hj ⋅ b ⋅ cn ] Pα [hj ∣ b] tosses makes the posterior plausibility that the coin is fair only about
× × .
P[en ∣ hi ⋅ b ⋅ cn ] Pα [hi ∣ b]

48 Stanford Encyclopedia of Philosophy Spring 2021 Edition 49


Inductive Logic James Hawthorne

6/1000ths as plausible as the hypothesis that it is warped towards heads P[en ∣ hj ⋅ b ⋅ cn ]


with propensity 3/4: P[en ∣ hi ⋅ b ⋅ cn ]

Pα [h[1/2] ∣ b ⋅ cn ⋅ en ] toward 0 (as n increases), then Equation 9∗ says that each false hj will
= .0056269.
Pα [h[3/4] ∣ b ⋅ cn ⋅ en ] become effectively refuted — each of their posterior probabilities will
approaches 0 (as n increases). As a result, the posterior probability of hi
Thus, such evidence strongly refutes the “fairness hypothesis” relative to must approach 1. The next two equations show precisely how this works.
the “3/4-heads hypothesis”, provided the assessment of prior prior
plausibilities doesn’t make the latter hypothesis too extremely implausible If we sum the ratio versions of Bayes’ Theorem in Equation 9∗ over all
to begin with. Notice, however, that strong refutation is not absolute alternatives to hypothesis hi (including the catch-all alternative hK , if
refutation. Additional evidence could reverse this trend towards the appropriate), we get the Odds Form of Bayes’ Theorem. By definition, the
refutation of the fairness hypothesis. odds against a statement A given B is related to the probability of A given
B as follows:
This example employs repetitions of the same kind of experiment—
Pα [∼A ∣ B] 1 − Pα [A ∣ B]
repeated tosses of a coin. But the point holds more generally. If, as the Ωα [∼A ∣ B] = = .
evidence increases, the likelihood ratios Pα [A ∣ B] Pα [A ∣ B]

P[en ∣ hj ⋅ b ⋅ cn ] This notion of odds gives rise to the following version of Bayes’ Theorem:
P[en ∣ hi ⋅ b ⋅ cn ] Bayes’ Theorem: Odds Form
approach 0, then the Ratio Forms of Bayes’ Theorem, Equations 9∗) and Pα [hj ∣ b ⋅ cn ⋅ en ]
9 ∗ ∗), show that the posterior probability of hj must approach 0 as well, (10) Ωα [∼hi ∣ b ⋅ cn ⋅ en ] = ∑
P [h ∣ b ⋅ cn ⋅ en ]
j≠i α i
since
Pα [hK ∣ b ⋅ cn ⋅ en ]
+
Pα [hj ∣ b ⋅ ⋅
cn en ] Pα [hi ∣ b ⋅ cn ⋅ en ]
Pα [hj ∣ b ⋅ cn ⋅ en ] ≤ .
Pα [hi ∣ b ⋅ cn ⋅ en ]
P[en ∣ hj ⋅ b ⋅ cn ] Pα [hj ∣ b]
=∑ ×
Such evidence comes to strongly refute hj , with little regard for its prior
j≠i
P[e ∣ hi ⋅ b ⋅ c ]
n n Pα [hi ∣ b]
plausibility value. Indeed, Bayesian induction turns out to be a version of
eliminative induction, and Equation 9∗ and 9 ∗ ∗ begin to illustrate this. Pα [en ∣ hK ⋅ b ⋅ cn ] P [h ∣ b]
+ × α K
For, suppose that hi is the true hypothesis, and consider what happens to P[e ∣ hi ⋅ b ⋅ c ]
n n Pα [hi ∣ b]
each of its false competitors, hj . If enough evidence becomes available to
drive each of the likelihood ratios where the factor following the ‘+’ sign is only required in cases where
a catch-all alternative hypothesis, hK , is needed.
P[ n ∣ hj ⋅ b ⋅ n]

50 Stanford Encyclopedia of Philosophy Spring 2021 Edition 51


Inductive Logic James Hawthorne

Recall that when we have a finite collection of concrete alternative Bayes’ Theorem: General Probabilistic Form
hypotheses available, {h1 , h2 , … , hm }, but where this set of alternatives is
1
not exhaustive (where additional, unarticulated, undiscovered alternative (11) Pα [hi ∣ b ⋅ cn ⋅ en ] = .
1 + Ωα [∼ hi ∣ b ⋅ cn ⋅ en ]
hypotheses may exist), the catch-all alternative hypothesis hK is just the
denial of each of the concrete alternatives, (∼h1 ⋅ ∼h2 ⋅ … ⋅ ∼hm ) . The odds against a hypothesis depends only on the values of ratios of
Generally, the likelihood of evidence claims relative to a catch-all posterior probabilities, which entirely derive from the Ratio Form of
hypothesis will not enjoy the same kind of objectivity possessed by the Bayes’ Theorem. Thus, we see that the individual value of the posterior
likelihoods for concrete alternative hypotheses. So, we leave the subscript probability of a hypothesis depends only on the ratios of posterior
α attached to the likelihood for the catch-all hypothesis to indicate this probabilities, which come from the Ratio Form of Bayes’ Theorem. Thus,
lack of objectivity. the Ratio Form of Bayes’ Theorem captures all the essential features of the
Bayesian evaluation of hypothesis. It shows how the impact of evidence
Although the catch-all hypothesis may lack objective likelihoods, the
(in the form of likelihood ratios) combines with comparative plausibility
influence of the catch-all term in Bayes’ Theorem diminishes as additional
assessments of hypotheses (in the form of ratios of prior probabilities) to
concrete hypotheses are articulated. That is, as new hypotheses are
provide a net assessment of the extent to which hypotheses are refuted or
discovered they are “peeled off” of the catch-all. So, when a new
supported via contests with their rivals.
hypothesis hm+1 is formulated and made explicit, the old catch-all
hypothesis hK is replaced by a new catch-all, hK∗ , of form There is a result, a kind of Bayesian Convergence Theorem, that shows
(∼h1 ⋅ ⋅∼h2 ⋅ … ⋅ ∼hm ⋅ ∼hm+1 ); and the prior probability for the new that if hi (together with b ⋅ cn ) is true, then the likelihood ratios
catch-all hypothesis is gotten by diminishing the prior of the old catch-all:
Pα [hK∗ ∣ b] = Pα [hK ∣ b] − Pα [hm+1 ∣ b]. Thus, the influence of the P[en ∣ hj ⋅ b ⋅ cn ]
catch-all term should diminish towards 0 as new alternative hypotheses are P[en ∣ hi ⋅ b ⋅ cn ]
made explicit.[10]
comparing evidentially distinguishable alternative hypothesis hj to hi will
If increasing evidence drives towards 0 the likelihood ratios comparing very probably approach 0 as evidence accumulates (i.e., as n increases).
each competitor hj with hypothesis hi , then the odds against hi , Let’s call this result the Likelihood Ratio Convergence Theorem. When
Ωα [∼hi ∣ b ⋅ cn ⋅ en ], will approach 0 (provided that priors of catch-all this theorem applies, Equation 9∗ shows that the posterior probability of a
terms, if needed, approach 0 as well, as new alternative hypotheses are false competitor hj will very probably approach 0 as evidence
made explicit and peeled off). And, as Ωα [∼hi ∣ b ⋅ cn ⋅ en ] approaches 0, accumulates, regardless of the value of its prior probability Pα [hj ∣ b]. As
the posterior probability of hi goes to 1. This derives from the fact that the this happens to each of hi ’s false competitors, Equations 10 and 11 say that
odds against hi is related to and its posterior probability by the following the posterior probability of the true hypothesis, hi , will approach 1 as
formula: evidence increases.[11] Thus, Bayesian induction is at bottom a version of
induction by elimination, where the elimination of alternatives comes by

52 Stanford Encyclopedia of Philosophy Spring 2021 Edition 53


Inductive Logic James Hawthorne

way of likelihood ratios approaching 0 as evidence accumulates. Thus, on privately held opinions. (Formally, the logic may represent comparative
when the Likelihood Ratio Convergence Theorem applies, the Criterion of plausibility arguments by explicit statements expressed within b .) It would
Adequacy for an Inductive Logic described at the beginning of this article be highly unscientific for a member of the scientific community to
will be satisfied: As evidence accumulates, the degree to which the disregard or dismiss a hypothesis that other members take to be a
collection of true evidence statements comes to support a hypothesis, as reasonable proposal with only the comment, “don’t ask me to give my
measured by the logic, should very probably come to indicate that false reasons, it’s just my opinion”. Even so, agents may be unable to specify
hypotheses are probably false and that true hypotheses are probably true. precisely how much more strongly the available plausibility arguments
We will examine this Likelihood Ratio Convergence Theorem in Section 4. support a hypothesis over an alternative; so prior probability ratios for
[12] hypotheses may be vague. Furthermore, agents in a scientific community
may disagree about how strongly the available plausibility arguments
A view called Likelihoodism relies on likelihood ratios in much the same support a hypothesis over a rival hypothesis; so prior probability ratios
way as the Bayesian logic articulated above. However, Likelihoodism may be somewhat diverse as well.
attempts to avoid the use of prior probabilities. For an account of this
alternative view, see the supplement Likelihood Ratios, Likelihoodism, Both the vagueness of comparative plausibilities assessments for
and the Law of Likelihood. For more discussion of Bayes’ Theorem and individual agents and the diversity of such assessments among the
its application, see the entries on Bayes’ Theorem and on Bayesian community of agents can be represented formally by sets of support
Epistemology in this Encyclopedia. functions, {Pα , Pβ , …} , that agree on the values for the likelihoods but
encompass a range of values for the (ratios of) prior probabilities of
3.4 On Prior Probabilities and Representations of Vague hypotheses. Vagueness and diversity are somewhat different issues, but
and Diverse Plausibility Assessments they may be represented in much the same way. Let’s briefly consider each
in turn.
Given that a scientific community should largely agree on the values of the
likelihoods, any significant disagreement among them with regard to the Assessments of the prior plausibilities of hypotheses will often be vague—
values of posterior probabilities of hypotheses should derive from not subject to the kind of precise quantitative treatment that a Bayesian
disagreements over their assessments of values for the prior probabilities version of probabilistic inductive logic may seem to require for prior
of those hypotheses. We saw in Section 3.3 that the Bayesian logic of probabilities. So, it may seem that the kind of assessment of prior
evidential support need only rely on assessments of ratios of prior probabilities required to get the Bayesian algorithm going cannot be
probabilities—on how much more plausible one hypothesis is than accomplished in practice. To see how Bayesian inductivists address this
another. Thus, the logic of evidential support only requires that scientists worry, first recall the Ratio Form of Bayes’ Theorem, Equation 9∗ .
can assess the comparative plausibilities of various hypotheses. Pα [hj ∣ b ⋅ cn ⋅ en ] P[en ∣ hj ⋅ b ⋅ cn ] Pα [hj ∣ b]
= ×
Pα [hi ∣ b ⋅ cn ⋅ en ] P[en ∣ hi ⋅ b ⋅ cn ] Pα [hi ∣ b]
Presumably, in scientific contexts the comparative plausibility values for
hypotheses should depend on explicit plausibility arguments, not merely

54 Stanford Encyclopedia of Philosophy Spring 2021 Edition 55


Inductive Logic James Hawthorne

P[en ∣ hj ⋅ b ⋅ cn ]
= LRn
Recall that this Ratio Form of the theorem captures the essential features
of the logic of evidential support, even though it only provides a value for P[en ∣ hi ⋅ b ⋅ cn ]
the ratio of the posterior probabilities. Notice that the ratio form of the
results in a posterior support ratio in the interval
theorem easily accommodates situations where we don’t have precise
numerical values for prior probabilities. It only depends on our ability to Pα [hj ∣ b ⋅ cn ⋅ en ]
(LRn × q) ≤ ≤ (LRn × r).
assess how much more or less plausible alternative hypothesis hj is than Pα [hi ∣ b ⋅ cn ⋅ en ]
hypothesis hi —only the value of the ratio Pα [hj ∣ b]/Pα [hi ∣ b] need be
assessed; the values of the individual prior probabilities are not needed. (Technically each probabilistic support function assigns a specific
Such comparative plausibilities are much easier to assess than specific numerical value to each pair of sentences; so when we write an
numerical values for the prior probabilities of individual hypotheses. inequality like
When combined with the ratio of likelihoods, this ratio of priors suffices Pα [hj ∣ b]
q≤ ≤r
to yield an assessment of the ratio of posterior plausibilities, Pα [hi ∣ b]
Pα [hj ∣ b ⋅ cn ⋅ en ]
. we are really referring to a set of probability functions Pα , a vagueness
Pα [hi ∣ b ⋅ cn ⋅ en ] set, for which the inequality holds. Thus, technically, the Bayesian
Although such posterior ratios don’t supply values for the posterior logic employs sets of probabilistic support functions to represent the
probabilities of individual hypotheses, they place a crucial constraint on vagueness in comparative plausibility values for hypotheses.)
the posterior support of hypothesis hj , since
Observe that if the likelihood ratio values LRn approach 0 as the amount
Pα [hj ∣ b ⋅ cn ⋅ en ] of evidence en increases, the interval of values for the posterior probability
Pα [hj ∣ b ⋅ cn ⋅ en ] < ratio must become tighter as the upper bound (LRn × r) approaches 0.
Pα [hi ∣ b ⋅ cn ⋅ en ]
P[en ∣ hj ⋅ b ⋅ cn ] Pα [hj ∣ b] Furthermore, the absolute degree of support for hj , Pα [hj ∣ b ⋅ cn ⋅ en ] ,
= ×
P[e ∣ hi ⋅ b ⋅ c ]
n n Pα [hi ∣ b] must also approach 0.

This Ratio Form of Bayes’ Theorem tolerates a good deal of vagueness or This observation is really useful. For, it can be shown that when hi ⋅ b ⋅ cn
imprecision in assessments of the ratios of prior probabilities. In practice is true and hj is empirically distinct from hi , the continual pursuit of
one need only assess bounds for these prior plausibility ratios to achieve evidence is very likely to result in evidential outcomes en that (as n
meaningful results. Given a prior ratio in a specific interval, increases) yield values of likelihood ratios
P[e ∣ hj ⋅ b ⋅ c ]/P[e ∣ hi ⋅ b ⋅ c ] that approach 0 as the amount of
n n n n
Pα [hj ∣ b]
q≤ ≤r evidence increases. This result, called the Likelihood Ratio Convergence
Pα [hi ∣ b] Theorem, will be investigated in more detail in Section 4. When that kind
a likelihood ratio of convergence towards 0 for likelihood ratios occurs, the upper bound on
P[ n ∣ hj ⋅ b ⋅ n]
= n

56 Stanford Encyclopedia of Philosophy Spring 2021 Edition 57


Inductive Logic James Hawthorne

the posterior probability ratio also approaches 0, driving the posterior that cover the ranges of values for comparative plausibility assessments
probability of hj to approach 0 as well, effectively refuting hypothesis hj . for pairs of competing hypotheses
Thus, false competitors of a true hypothesis will effectively be eliminated
Pα [hj ∣ b]
by increasing evidence. As this happens, Equations 9* through 11 show q≤ ≤r
that the posterior probability Pα [hi ∣ b ⋅ cn ⋅ en ] of the true hypothesis hi
Pα [hi ∣ b]
approaches 1. as assessed by the scientific community. But, once again, if accumulating
evidence drives the likelihood ratios comparing various alternative
Thus, Bayesian logic of inductive support for hypotheses is a form of
hypotheses to the true hypothesis towards 0, the range of support functions
eliminative induction, where the evidence effectively refutes false
in a diversity set will come to near agreement, near 0, on the values for
alternatives to the true hypothesis. Because of its eliminative nature, the
posterior probabilities of false competitors of the true hypothesis. So, not
Bayesian logic of evidential support doesn’t require precise values for
only does such evidence firm up each agent’s vague initial plausibility
prior probabilities. It only needs to draw on bounds on the values of
assessment, it also brings the whole community into agreement on the
comparative plausibility ratios, and these bounds only play a significant
near refutation of empirically distinct competitors of a true hypothesis. As
role while evidence remains fairly weak. If the true hypothesis is assessed
this happens, the posterior probability of the true hypothesis may approach
to be comparatively plausible (due to plausibility arguments contained in
1. The Likelihood Ratio Convergence Theorem implies that this kind of
b), then plausibility assessments give it a leg-up over alternatives. If the
convergence to the truth should very probably happen, provided that the
true hypothesis is assessed to be comparatively implausible, the
true hypothesis is empirically distinct enough from its rivals.
plausibility assessments merely slow down the rate at which it comes to
dominate its rivals, reflecting the idea that extraordinary hypotheses One more point about prior probabilities and Bayesian convergence
require extraordinary evidence (or an extraordinary accumulation of should be mentioned before proceeding to Section 4. Some subjectivist
evidence) to overcome their initial implausibilities. Thus, as evidence versions of Bayesian induction seem to suggest that an agent’s prior
accumulates, the agent’s vague initial plausibility assessments transform plausibility assessments for hypotheses should stay fixed once-and-for-all,
into quite sharp posterior probabilities that indicate their strong refutation and that all plausibility updating should be brought about via the
or support by the evidence. likelihoods in accord with Bayes’ Theorem. Critics argue that this is
unreasonable. The members of a scientific community may quite
When the various agents in a community may widely disagree over the
legitimately revise their (comparative) prior plausibility assessments for
non-evidential plausibilities of hypotheses, the Bayesian logic of
hypotheses from time to time as they rethink plausibility arguments and
evidential support may represent this kind of diversity across the
bring new considerations to bear. This seems a natural part of the
community of agents as a collection of the agents’ vagueness sets of
conceptual development of a science. It turns out that such reassessments
support functions. Let’s call such a collection of support functions a
of the comparative plausibilities of hypotheses poses no difficulty for the
diversity set. That is, a diversity set is just a set of support functions Pα
probabilistic inductive logic discussed here. Such reassessments may be

58 Stanford Encyclopedia of Philosophy Spring 2021 Edition 59


Inductive Logic James Hawthorne

represented by the addition or modification of explicit statements that theorem places an explicit lower bound on the “rate of probable
modify the background information b. Such reassessments may result in convergence” of these likelihood ratios towards 0. That is, it puts a lower
(non-Bayesian) transitions to new vagueness sets for individual agents and bound on how likely it is, if hi is true, that a stream of outcomes will occur
new diversity sets for the community. The logic of Bayesian induction (as that yields likelihood ratio values against hj as compared to hi that lie
described here) has nothing to say about what values the prior plausibility within any specified small distance above 0.
assessments for hypotheses should have; and it places no restrictions on
how they might change over time. Provided that the series of The theorem itself does not require the full apparatus of Bayesian
reassessments of (comparative) prior plausibilities doesn’t happen to probability functions. It draws only on likelihoods. Neither the statement
diminish the (comparative) prior plausibility value of the true hypothesis of the theorem nor its proof employ prior probabilities of any kind. So
towards zero (or, at least, doesn’t do so too quickly), the Likelihood Ratio even likelihoodists, who eschew the use of Bayesian prior probabilities,
Convergence Theorem implies that the evidence will very probably bring may embrace this result. Given the forms of Bayes’ Theorem, 9*-11 from
the posterior probabilities of empirically distinct rivals of the true the previous section, the Likelihood Ratio Convergence Theorem further
hypothesis to approach 0 via decreasing likelihood ratios; and as this implies the likely convergence to 0 of the posterior probabilities of false
happens, the posterior probability of the true hypothesis will head towards competitors of a true hypothesis. That is, when the ratios
1. P[en ∣ hj ⋅ b ⋅ cn ]/P[en ∣ hi ⋅ b ⋅ cn ] approach 0 for increasing n, the Ratio
Form of Bayes’ Theorem, Equation 9*, says that the posterior probability
(Those interested in a Bayesian account of Enumerative Induction and the of hj must also approach 0 as evidence accumulates, regardless of the
estimation of values for relative frequencies of attributes in populations value of its prior probability. So, support functions in collections
should see the supplement, Enumerative Inductions: Bayesian Estimation representing vague prior plausibilities for an individual agent (i.e., a
and Convergence.) vagueness set) and representing the diverse range of priors for a
community of agents (i.e., a diversity set) will come to agree on the near 0
4. The Likelihood Ratio Convergence Theorem posterior probability of empirically distinct false rivals of a true
hypothesis. And as the posterior probabilities of false competitors fall, the
In this section we will investigate the Likelihood Ratio Convergence posterior probability of the true hypothesis heads towards 1. Thus, the
Theorem. This theorem shows that under certain reasonable conditions, theorem establishes that the inductive logic of probabilistic support
when hypothesis hi (in conjunction with auxiliaries in b) is true and an functions satisfies the Criterion of Adequacy (CoA) suggested at the
alternative hypothesis hj is empirically distinct from hi on some possible beginning of this article.
outcomes of experiments or observations described by conditions ck , then
it is very likely that a long enough sequence of such experiments and The Likelihood Ratio Convergence Theorem merely provides some
observations cn will produce a sequence of outcomes en that yields sufficient conditions for probable convergence. But likelihood ratios may
likelihood ratios P[en ∣ hj ⋅ b ⋅ cn ]/P[en ∣ hi ⋅ b ⋅ cn ] that approach 0, well converge towards 0 (in the way described by the theorem) even when
favoring hi over hj , as evidence accumulates (i.e., as n increases). This the antecedent conditions of the theorem are not satisfied. This theorem

60 Stanford Encyclopedia of Philosophy Spring 2021 Edition 61


Inductive Logic James Hawthorne

overcomes many of the objections raised by critics of Bayesian probability of the true hypothesis towards 0 too rapidly, the theorem
convergence results. First, this theorem does not employ second-order implies that the posterior probabilities of each empirically distinct false
probabilities; it says noting about the probability of a probability. It only competitor will very probably approach 0 as evidence increases.[13]
concerns the probability of a particular disjunctive sentence that expresses
a disjunction of various possible sequences of experimental or 4.1 The Space of Possible Outcomes of Experiments and
observational outcomes. The theorem does not require evidence to consist Observations
of sequences of events that, according to the hypothesis, are identically
distributed (like repeated tosses of a die). The result is most easily To specify the details of the Likelihood Ratio Convergence Theorem we’ll
expressed in cases where the individual outcomes of a sequence of need a few additional notational conventions and definitions. Here they
experiments or observations are probabilistically independent, given each are.
hypothesis. So that is the version that will be presented in this section.
For a given sequence of n experiments or observations cn , consider the set
However, a version of the theorem also holds when the individual
of those possible sequences of outcomes that would result in likelihood
outcomes of the evidence stream are not probabilistically independent,
ratios for hj over hi that are less than some chosen small number ε > 0 .
given the hypotheses. (This more general version of the theorem will be
This set is represented by the expression,
presented in a supplement on the Probabilistic Refutation Theorem, below,
P[en ∣ hj ⋅ b ⋅ cn ]
{ }
where the proof of both versions is provided.) In addition, this result does
en : <ε .
not rely on supposing that the probability functions involved are countably P[en ∣ hi ⋅ b ⋅ cn ]
additive. Furthermore, the explicit lower bounds on the rate of
convergence provided by this result means that there is no need to wait for Placing the disjunction symbol ‘∨ ’ in front of this expression yields an
the infinitely long run before convergence occurs (as some critics seem to expression,
think).
P[en ∣ hj ⋅ b ⋅ cn ]
{ }
∨ en : <ε ,
It is sometimes claimed that Bayesian convergence results only work P[en ∣ hi ⋅ b ⋅ cn ]
when an agent locks in values for the prior probabilities of hypotheses
that we’ll use to represent the disjunction of all outcome sequences en in
once-and-for-all, and then updates posterior probabilities from there only
this set. So,
by conditioning on evidence via Bayes Theorem. The Likelihood Ratio
P[en ∣ hj ⋅ b ⋅ cn ]
{ }
Convergence Theorem, however, applies even if agents revise their prior
∨ en : <ε
probability assessments over time. Such non-Bayesian shifts from one P[en ∣ hi ⋅ b ⋅ cn ]
support function (or vagueness set) to another may arise from new
plausibility arguments or from reassessments of the strengths of old ones. is just a particular sentence that says, in effect, “one of the sequences of
The Likelihood Ratio Convergence Theorem itself only involves the values outcomes of the first n experiments or observations will occur that makes
of likelihoods. So, provided such reassessments don’t push the prior the likelihood ratio for hj over hi less than ε ”.

62 Stanford Encyclopedia of Philosophy Spring 2021 Edition 63


Inductive Logic James Hawthorne

The Likelihood Ratio Convergence Theorem says that under certain We now let expressions of form ‘ek ’ act as variables that range over the
conditions (covered in detail below), the likelihood of a disjunctive possible outcomes of condition ck —i.e., ek ranges over the members of
sentence of this sort, given that ‘hi ⋅ b ⋅ cn ’ is true, Ok . As before, ‘cn ’ denotes the conjunction of the first n test conditions,
(c1 ⋅ c2 ⋅ … ⋅ cn ) , and ‘en ’ represents possible sequences of corresponding
P[en ∣ hj ⋅ b ⋅ cn ]
[ { } ]
P ∨ en : < ε ∣ hi ⋅ b ⋅ cn , outcomes, (e1 ⋅ e2 ⋅ … ⋅ en ) . Let’s use the expression ‘En’ to represent the
P[en ∣ hi ⋅ b ⋅ cn ] set of all possible outcome sequences that may result from the sequence of
must be at least 1 − (ψ/n), for some explicitly calculable term ψ . Thus, the conditions cn. So, for each hypothesis hj (including hi ),
true hypothesis hi probabilistically implies that as the amount of evidence, ∑en ∈E n P[en ∣ hj ⋅ b ⋅ cn ] = 1.
n, increases, it becomes highly likely (as close to 1 as you please) that one
Everything introduced in this subsection is mere notational convention. No
of the outcome sequences en will occur that yields a likelihood ratio
substantive suppositions (other than the axioms of probability theory) have
P[en ∣ hj ⋅ b ⋅ cn ]/P[en ∣ hi ⋅ b ⋅ cn ] less than ε ; and this holds for any
yet been introduced. The version of the Likelihood Ratio Convergence
specific value of ε you may choose. As this happens, the posterior
Theorem I’ll present below does, however, draw on one substantive
probability of hi ’s false competitor, hj , must approach 0, as required by the
supposition, although a rather weak one. The next subsection will discuss
Ratio Form of Bayes’ Theorem, Equation 9*.
that supposition in detail.
The term ψ in the lower bound of this probability depends on a measure of
the empirical distinctness of the two hypotheses hj and hi for the proposed 4.2 Probabilistic Independence
sequence of experiments and observations cn . To specify this measure we
In most scientific contexts the outcomes in a stream of experiments or
need to contemplate the collection of possible outcomes of each
observations are probabilistically independent of one another relative to
experiment or observation. So, consider some sequence of experimental or
observational conditions described by sentences c1 , c2 , … , cn .
each hypothesis under consideration, or can at least be divided up into
probabilistically independent parts. For our purposes probabilistic
Corresponding to each condition ck there will be some range of possible
alternative outcomes. Let Ok = {ok1 , ok2 , … , okw } be a set of statements
independence of evidential outcomes on a hypothesis divides neatly into
two types.
describing the alternative possible outcomes for condition ck . (The number
of alternative outcomes will usually differ for distinct experiments among Definition: Independent Evidence Conditions:
those in the sequence c1 , … , cn ; so, the value of w may depend on ck .) For
each hypothesis hj , the alternative outcomes of ck in Ok are mutually 1. A sequence of outcomes ek is condition-independent of a
exclusive and exhaustive, so we have: condition for an additional experiment or observation ck+1 , given
h ⋅ b together with its own conditions ck , if and only if
w
P[oku ⋅ okv ∣ hj ⋅ b ⋅ ck ] = 0 and ∑ P[ oku ∣ hj ⋅ b ⋅ ck ] = 1.
P[ek ∣ h ⋅ b ⋅ ck ⋅ ck+1 ] = P[ek ∣ h ⋅ b ⋅ ck ].
u=1

64 Stanford Encyclopedia of Philosophy Spring 2021 Edition 65


Inductive Logic James Hawthorne

2. An individual outcome ek is result-independent of a sequence of In scientific contexts the evidence can almost always be divided into parts
other observations and their outcomes (ck−1 ⋅ ek−1 ), given h ⋅ b that satisfy both clauses of the Independent Evidence Condition with
and its own condition ck , if and only if respect to each alternative hypothesis. To see why, let us consider each
independence condition more carefully.
P[ek ∣ h ⋅ b ⋅ ck ⋅ ( ck−1 ⋅ ek−1 )] = P[ek ∣ h ⋅ b ⋅ ck ].
Condition-independence says that the mere addition of a new observation
When these two conditions hold, the likelihood for an evidence sequence condition ck+1 , without specifying one of its outcomes, does not alter the
may be decomposed into the product of the likelihoods for individual likelihood of the outcomes ek of other experiments ck . To appreciate the
experiments or observations. To see how the two independence conditions significance of this condition, imagine what it would be like if it were
affect the decomposition, first consider the following formula, which holds violated. Suppose hypothesis hj is some statistical theory, say, for
even when neither independence condition is satisfied: example, a quantum theory of superconductivity. The conditions expressed
n in ck describe a number of experimental setups, perhaps conducted in
(12) P[en ∣ hj ⋅ b ⋅ cn ] = ∏ P[ek ∣ hj ⋅ b ⋅ cn ⋅ ek−1 ]. numerous labs throughout the world, that test a variety of aspects of the
k=1 theory (e.g., experiments that test electrical conductivity in different
When condition-independence holds, the likelihood of the whole evidence materials at a range of temperatures). An outcome sequence ek describes
stream parses into a product of likelihoods that probabilistically depend on the results of these experiments. The violation of condition-independence
only past observation conditions and their outcomes. They do not depend would mean that merely adding to hj ⋅ b ⋅ ck a statement ck+1 describing
on the conditions for other experiments whose outcomes are not yet how an additional experiment has been set up, but with no mention of its
specified. Here is the formula: outcome, changes how likely the evidence sequence ek is taken to be.
What (hj ⋅ b) says via likelihoods about the outcomes ek of experiments ck
n
differs as a result of merely supplying a description of another
(13) P[en ∣ hj ⋅ b ⋅ cn ] = ∏ P[ek ∣ hj ⋅ b ⋅ ck ⋅ (ck−1 ⋅ ek−1 )].
experimental arrangement, ck+1 . Condition-independence, when it holds,
k=1
rules out such strange effects.
Finally, whenever both independence conditions are satisfied we have the
following relationship between the likelihood of the evidence stream and Result-independence says that the description of previous test conditions
the likelihoods of individual experiments or observations: together with their outcomes is irrelevant to the likelihoods of outcomes
for additional experiments. If this condition were widely violated, then in
n
(14) P[en ∣ hj ⋅ b ⋅ cn ] = ∏ P[ek ∣ hj ⋅ b ⋅ ck ].
order to specify the most informed likelihoods for a given hypothesis one
k=1
would need to include information about volumes of past observations and
their outcomes. What a hypothesis says about future cases would depend
(For proofs of Equations 12–14 see the supplement Immediate on how past cases have gone. Such dependence had better not happen on a
Consequences of Independent Evidence Conditions.) large scale. Otherwise, the hypothesis would be fairly useless, since its

66 Stanford Encyclopedia of Philosophy Spring 2021 Edition 67


Inductive Logic James Hawthorne

empirical import in each specific case would depend on taking into 1. Each sequence of possible outcomes ek of a sequence of
account volumes of past observational and experimental results. However, conditions ck is condition-independent of additional conditions
even if such dependencies occur, provided they are not too pervasive, ck+1 —i.e.,
result-independence can be accommodated rather easily by packaging
each collection of result-dependent data together, treating it like a single P[ek ∣ h ⋅ b ⋅ ck ⋅ ck+1 ] = P[ek ∣ h ⋅ b ⋅ ck ].
extended experiment or observation. The result-independence condition
2. Each possible outcome ek of condition ck is result-independent
will then be satisfied by letting each term ‘ck ’ in the statement of the
of sequences of other observations and possible outcomes
(ck−1 ⋅ ek−1 )—i.e.,
independence condition represent a conjunction of test conditions for a
collection of result-dependent tests, and by letting each term ‘ek ’ (and each
term ‘oku ’) stand for a conjunction of the corresponding result-dependent P[ek ∣ h ⋅ b ⋅ ck ⋅ ( ck−1 ⋅ ek−1 )] = P[ek ∣ h ⋅ b ⋅ ck ].
outcomes. Thus, by packaging result-dependent data together in this way,
the result-independence condition is satisfied by those (conjunctive) We now have all that is needed to begin to state the Likelihood Ratio
statements that describe the separate, result-independent chunks.[14] Convergence Theorem.

The version of the Likelihood Ratio Convergence Theorem we will 4.3 Likelihood Ratio Convergence when Falsifying
examine depends only on the Independent Evidence Conditions (together Outcomes are Possible
with the axioms of probability theory). It draws on no other assumptions.
Indeed, an even more general version of the theorem can be established, a The Likelihood Ratio Convergence Theorem comes in two parts. The first
version that draws on neither of the Independent Evidence Conditions. part applies only to those experiments or observations ck within the total
However, the Independent Evidence Conditions will be satisfied in almost evidence stream cn for which some of the possible outcomes have 0
all scientific contexts, so little will be lost by assuming them. (And the likelihood of occurring according to hypothesis hj but have non-0
presentation will run more smoothly if we side-step the added likelihood of occurring according to hi . Such outcomes are highly
complications needed to explain the more general result.) desirable. If they occur, the likelihood ratio comparing hj to hi will
become 0, and hj will be falsified. So-called crucial experiments are a
From this point on, let us assume that the following versions of the
special case of this, where for at least one possible outcome oku ,
Independent Evidence Conditions hold.
P[oku ∣ hi ⋅ b ⋅ ck ] = 1 and P[oku ∣ hj ⋅ b ⋅ ck ] = 0. In the more general
Assumption: Independent Evidence Assumptions. For each case hi together with b says that one of the outcomes of ck is at least
hypothesis h and background b under consideration, we assume that minimally probable, whereas hj says that this outcome is impossible—i.e.,
the experiments and observations can be packaged into condition P[oku ∣ hi ⋅ b ⋅ ck ] > 0 and P[oku ∣ hj ⋅ b ⋅ ck ] = 0. It will be convenient
statements, c1 , … , ck , ck+1 , …, and possible outcomes in a way that to define a term for this situation.
satisfies the following conditions:
k

68 Stanford Encyclopedia of Philosophy Spring 2021 Edition 69


Inductive Logic James Hawthorne

Definition: Full Outcome Compatibility. Let’s call hj fully outcome- two hypotheses. Furthermore, suppose there is a lower bound δ > 0
compatible with hi on experiment or observation ck just when, for such that for each ck on which hj fails to be fully outcome-compatible
each of its possible outcomes ek , if P[ek ∣ hi ⋅ b ⋅ ck ] > 0, then with hi ,
P[ek ∣ hj ⋅ b ⋅ ck ] > 0. Equivalently, hj is fails to be fully outcome-
compatible with hi on experiment or observation ck just when, for at P[∨{oku : P[oku ∣ hj ⋅ b ⋅ ck ] = 0} ∣ hi ⋅ b ⋅ ck ] ≥ δ
least one of its possible outcomes ek , P[ek ∣ hi ⋅ b ⋅ ck ] > 0 but
—i.e., hi together with b ⋅ ck says, with likelihood at least as large as
P[ek ∣ hj ⋅ b ⋅ ck ] = 0.
δ, that one of the outcomes will occur that hj says cannot occur. Then,

P[en ∣ hj ⋅ b ⋅ cn ]
[ { } ]
The first part of the Likelihood Ratio Convergence Theorem applies to that
P ∨ en : = 0 ∣ hi ⋅ b ⋅ cn
P[en ∣ hi ⋅ b ⋅ cn ]
part of the total stream of evidence (i.e., that subsequence of the total
evidence stream) on which hypothesis hj fails to be fully outcome-
compatible with hypothesis hi ; the second part of the theorem applies to = P [∨ {en : P[en ∣ hj ⋅ b ⋅ cn ] = 0} ∣ hi ⋅ b ⋅ cn ]
the remaining part of the total stream of evidence, that subsequence of the ≥ 1 − (1 − δ)m ,
total evidence stream on which hj is fully outcome-compatible with hi . It
turns out that these two kinds of cases must be treated differently. (This is which approaches 1 for large m. (For proof see Proof of the
due to the way in which the expected information content for empirically Falsification Theorem.)
distinguishing between the two hypotheses will be measured for
In other words, we only suppose that for each of m observations, ck , hi
experiments and observations that are fully outcome compatible; this
says observation ck has at least a small likelihood δ of producing one of
measure of information content blows up (becomes infinite) for
the outcomes oku that hj says is impossible. If the number m of such
experiments and observations that fail to be fully outcome compatible).
experiments or observations is large enough (or if the lower bound δ on
Thus, the following part of the convergence theorem applies to just that
the likelihoods of getting such outcomes is large enough), and if hi
part of the total stream of evidence that consists of experiments and
(together with b ⋅ cn ) is true, then it is highly likely that one of the
observations that fail to be fully outcome compatible for the pair of
outcomes held to be impossible by hj will actually occur. If one of these
hypotheses involved. Here, then, is the first part of the convergence
outcomes does occur, then the likelihood ratio for hj as compared to over
theorem.
hi will become 0. According to Bayes’ Theorem, when this happen, hj is
Likelihood Ratio Convergence Theorem 1—The Falsification absolutely refuted by the evidence—its posterior probability becomes 0.
Theorem:
The Falsification Theorem is quite commonsensical. First, notice that if
Suppose that the total stream of evidence cn contains precisely m
there is a crucial experiment in the evidence stream, the theorem is
experiments or observations on which hj fails to be fully outcome-
completely obvious. That is, suppose for the specific experiment ck (in
compatible with hi . And suppose that the Independent Evidence
evidence stream cn ) there are two incompatible possible outcomes okv and
Conditions hold for evidence stream cn with respect to each of these

70 Stanford Encyclopedia of Philosophy Spring 2021 Edition 71


Inductive Logic James Hawthorne

oku such that P[okv ∣ hj ⋅ b ⋅ ck ] = 1 and P[oku ∣ hi ⋅ b ⋅ ck ] = 1. Then, scientific hypotheses hi and hj one may directly compute the likelihood,
clearly, P[∨{oku : P[oku ∣ hj ⋅ b ⋅ ck ] = 0} ∣ hi ⋅ b ⋅ ck ] = 1, since oku is given (hi ⋅ b ⋅ cn ) , that a proposed sequence of experiments or
one of the oku such that P[oku ∣ hj ⋅ b ⋅ ck ] = 0. So, where a crucial observations cn will result in one of the sequences of outcomes that would
experiment is available, the theorem applies with m = 1 and δ = 1. yield low likelihood ratios. So, given a specific pair of hypotheses and a
proposed sequence of experiments, we don’t need a general Convergence
The theorem is equally commonsensical for cases where no crucial Theorem to tell us the likelihood of obtaining refuting evidence. The
experiment is available. To see what it says in such cases, consider an specific hypotheses hi and hj tell us this themselves. They tell us the
example. Let hi be some theory that implies a specific rate of proton likelihood of obtaining each specific outcome stream, including those that
decay, but a rate so low that there is only a very small probability that any either refute the competitor or produce a very small likelihood ratio for it.
particular proton will decay in a given year. Consider an alternative theory Furthermore, after we’ve actually performed an experiment and recorded
hj that implies that protons never decay. If hi is true, then for a persistent its outcome, all that matters is the actual ratio of likelihoods for that
enough sequence of observations (i.e., if proper detectors can keep outcome. Convergence theorems become moot.
trillions of protons under observation for long enough), eventually a
proton decay will almost surely be detected. When this happens, the The point of the Likelihood Ratio Convergence Theorem (both the
likelihood ratio becomes 0. Thus, the posterior probability of hj becomes Falsification Theorem and the part of the theorem still to come) is to
0. assure us in advance of considering any specific pair of hypotheses that if
the possible evidence streams that test hypotheses have certain
It is instructive to plug some specific values into the formula given by the characteristics which reflect the empirical distinctness of the two
Falsification Theorem, to see what the convergence rate might look like. hypotheses, then it is highly likely that one of the sequences of outcomes
For example, the theorem tells us that if we compare any pair of will occur that yields a very small likelihood ratio. These theorems
hypotheses hi and hj on an evidence stream cn that contains at least provide finite lower bounds on how quickly such convergence is likely to
m = 19 observations or experiments, where each has a likelihood δ ≥ .10 be. Thus, they show that the CoA is satisfied in advance of our using the
of yielding a falsifying outcome, then the likelihood (on hi ⋅ b ⋅ cn ) of logic to test specific pairs of hypotheses against one another.
obtaining an outcome sequence en that yields likelihood-ratio

P[en ∣ hj ⋅ b ⋅ cn ] 4.4 Likelihood Ratio Convergence When No Falsifying


= 0,
P[en ∣ hi ⋅ b ⋅ cn ] Outcomes are Possible

will be at least as large as (1 − (1 − .1)19 ) = .865. (The reader is invited The Falsification Theorem applies whenever the evidence stream includes
to try other values of δ and m.) possible outcomes that may falsify the alternative hypothesis. However, it
completely ignores the influence of any experiments or observations in the
A comment about the need for and usefulness of such convergence evidence stream on which hypothesis hj is fully outcome-compatible with
theorems is in order, now that we’ve seen one. Given some specific pair of hypothesis hi . We now turn to a theorem that applies to those evidence

72 Stanford Encyclopedia of Philosophy Spring 2021 Edition 73


Inductive Logic James Hawthorne

streams (or to parts of evidence streams) consisting only of experiments employ a symmetric measure. The logarithm of the likelihood ratio
and observations on which hypothesis hj is fully outcome-compatible with provides such a measure.
hypothesis hi . Evidence streams of this kind contain no possibly falsifying
outcomes. In such cases the only outcomes of an experiment or Definition: QI—the Quality of the Information.
observation ck for which hypothesis hj may specify 0 likelihoods are those For each experiment or observation ck , define the quality of the
for which hypothesis hi specifies 0 likelihoods as well. information provided by possible outcome oku for distinguishing hj
from hi , given b, as follows (where henceforth we take “logs” to be
Hypotheses whose connection with the evidence is entirely statistical in base-2):
nature will usually be fully outcome-compatible on the entire evidence
P[oku ∣ hi ⋅ b ⋅ ck ]
[ P[oku ∣ hj ⋅ b ⋅ ck ] ]
stream. So, evidence streams of this kind are undoubtedly much more QI[oku ∣ hi /hj ∣ b ⋅ ck ] = log .
common in practice than those containing possibly falsifying outcomes.
Furthermore, whenever an entire stream of evidence contains some Similarly, for the sequence of experiments or observations cn , define
mixture of experiments and observations on which the hypotheses are not the quality of the information provided by possible outcome en for
fully outcome compatible along with others on which they are fully distinguishing hj from hi , given b, as follows:
outcome compatible, we may treat the experiments and observations for
P[en ∣ hi ⋅ b ⋅ cn ]
[ P[en ∣ hj ⋅ b ⋅ cn ] ]
QI[en ∣ hi /hj ∣ b ⋅ cn ] = log .
which full outcome compatibility holds as a separate subsequence of the
entire evidence stream, to see the likely impact of that part of the evidence
in producing values for likelihood ratios.
That is, QI is the base-2 logarithm of the likelihood ratio for hi over
To cover evidence streams (or subsequences of evidence streams) that for hj .
consisting entirely of experiments or observations on which hj is fully
outcome-compatible with hypothesis hi we will first need to identify a So, we’ll measure the Quality of the Information an outcome would yield
useful way to measure the degree to which hypotheses are empirically in distinguishing between two hypotheses as the base-2 logarithm of the
distinct from one another on such evidence. Consider some particular likelihood ratio. This is clearly a symmetric measure of the outcome’s
sequence of outcomes en that results from observations cn . The likelihood evidential strength at distinguishing between the two hypotheses. On this
ratio P[en ∣ hj ⋅ b ⋅ cn ]/P[en ∣ hi ⋅ b ⋅ cn ] itself measures the extent to measure hypotheses hi and hj assign the same likelihood value to a given
which the outcome sequence distinguishes between hi and hj . But as a outcome oku just when QI[oku ∣ hi /hj ∣ b ⋅ ck ] = 0. Thus, QI measures
measure of the power of evidence to distinguish among hypotheses, raw information on a logarithmic scale that is symmetric about the natural no-
likelihood ratios provide a rather lopsided scale, a scale that ranges from 0 information midpoint, 0. This measure is set up so that positive
to infinity with the midpoint, where en doesn’t distinguish at all between information favors hi over hj , and negative information favors hj over hi .
hi and hj , at 1. So, rather than using raw likelihood ratios to measure the
ability of en to distinguish between hypotheses, it proves more useful to

74 Stanford Encyclopedia of Philosophy Spring 2021 Edition 75


Inductive Logic James Hawthorne

Given the Independent Evidence Assumptions with respect to each For hj fully outcome-compatible with hi on experiment or observation
hypothesis, it’s easy to show that the QI for a sequence of outcomes is just ck , define
the sum of the QIs of the individual outcomes in the sequence:
EQI[ck ∣ hi /hj ∣ b] = ∑ QI[oku ∣ hi /hj ∣ b ⋅ ck ] × P[oku ∣ hi ⋅ b ⋅ ck ].
n u
(15) QI[en ∣ hi / hj ∣ b ⋅ cn ] = ∑ QI[ek ∣ hi /hj ∣ b ⋅ ck ].
k=1 Also, for hj fully outcome-compatible with hi on each experiment and
observation in the sequence cn , define
Probability theorists measure the expected value of a quantity by first
multiplying each of its possible values by their probabilities of occurring, EQI[cn ∣ hi /hj ∣ b] = ∑ QI[en ∣ hi /hj ∣ b ⋅ cn ] × P[en ∣ hi ⋅ b ⋅ cn ].
and then summing these products. Thus, the expected value of QI is given n
e ∈E
n

by the following formula:


The EQI of an experiment or observation is the Expected Quality of its
Definition: EQI—the Expected Quality of the Information. Information for distinguishing hi from hj when hi is true. It is a measure of
We adopt the convention that if P[oku ∣ hi ⋅ b ⋅ ck ] = 0, then the term the expected evidential strength of the possible outcomes of an experiment
QI[oku ∣ hi /hj ∣ b ⋅ ck ] × P[oku ∣ hi ⋅ b ⋅ ck ] = 0. This convention will or observation at distinguishing between the hypotheses when hi (together
make good sense in the context of the following definition because, with b ⋅ c) is true. Whereas QI measures the ability of each particular
whenever the outcome oku has 0 probability of occurring according to outcome or sequence of outcomes to empirically distinguish hypotheses,
hi (together with b ⋅ ck ), it makes good sense to give it 0 impact on the EQI measures the tendency of experiments or observations to produce
ability of the evidence to distinguish between hj and hi when hi distinguishing outcomes. It can be shown that EQI tracks empirical
(together with b ⋅ ck ) is true. Also notice that the full outcome- distinctness in a very precise way. We return to this in a moment.
compatibility of hj with hi on ck means that whenever
P[ek ∣ hj ⋅ b ⋅ ck ] = 0, we must have P[ek ∣ hi ⋅ b ⋅ ck ] = 0 as well; so It is easily seen that the EQI for a sequence of observations cn is just the
whenever the denominator would be 0 in the term sum of the EQIs of the individual observations ck in the sequence:
n
P[oku ∣ hi ⋅ b ⋅ ck ]
[ P[oku ∣ hj ⋅ b ⋅ ck ] ]
QI[oku ∣ hi /hj ∣ b ⋅ ck ] = log , (16) EQI[cn ∣ hi /hj ∣ b] = ∑ EQI[ck ∣ hi /hj ∣ b].
k=1

the convention just described makes the term (For proof see the supplement Proof that the EQI for cn is the sum of the
EQI for the individual ck .)
QI[ oku ∣ hi /hj ∣ b ⋅ ck ] × P[oku ∣ hi ⋅ b ⋅ ck ] = 0.
This suggests that it may be useful to average the values of the
Thus the following notion is well-defined: EQI[ck ∣ hi /hj ∣ b] over the number of observations n to obtain a measure

76 Stanford Encyclopedia of Philosophy Spring 2021 Edition 77


Inductive Logic James Hawthorne

of the average expected quality of the information among the experiments (For proof, see the supplement The Effect on EQI of Partitioning the
and observations that make up the evidence stream cn . Outcome Space More Finely—Including Proof of the Nonnegativity
of EQI.)
Definition: The Average Expected Quality of Information
For hj fully outcome-compatible with hi on each experiment and In fact, the more finely one partitions the outcome space
observation in the evidence stream cn , define the average expected Ok = {ok1 , … , okv , … , okw } into distinct outcomes that differ on
⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯
quality of information, EQI, from cn for distinguishing hj from hi , likelihood ratio values, the larger EQI becomes.[15] This shows that EQI
given hi ⋅ b , as follows: tracks empirical distinctness in a precise way. The importance of the Non-
negativity of EQI result for the Likelihood Ratio Convergence Theorem
⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯ EQI[cn ∣ hi /hj ∣ b]
EQI[cn ∣ hi /hj ∣ b] = will become clear in a moment.
n
n
= (1/n) × ∑ EQI[ ck ∣ hi /hj ∣ b].
We are now in a position to state the second part of the Likelihood Ratio
Convergence Theorem. It applies to all evidence streams not containing
k=1
possibly falsifying outcomes for hj when hi holds—i.e., it applies to all
It turns out that the value of EQI[ck ∣ hi /hj ∣ b] cannot be less than 0; and evidence streams for which hj is fully outcome-compatible with hi on each
it must be greater than 0 just in case hi is empirically distinct from hj on at ck in the stream.
least one outcome oku —i.e., just in case it is empirically distinct in the
sense that P[oku ∣ hi ⋅ b ⋅ ck ] ≠ P[oku ∣ hj ⋅ b ⋅ ck ] , for at least one Likelihood Ratio Convergence Theorem 2—The Probabilistic
⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯
outcome oku . The same goes for the average, EQI[cn ∣ hi /hj ∣ b]. Refutation Theorem.

Theorem: Nonnegativity of EQI. Suppose the evidence stream cn contains only experiments or
observations on which hj is fully outcome-compatible with hi —i.e.,
EQI[ck ∣ hi /hj ∣ b] ≥ 0; and EQI[ck ∣ hi /hj ∣ b] > 0 if and only if for suppose that for each condition ck in sequence cn , for each of its
at least one of its possible outcomes oku , possible outcomes possible outcomes oku , either
P[oku ∣ hi ⋅ b ⋅ ck ] = 0 or P[oku ∣ hj ⋅ b ⋅ ck ] > 0. In addition (as a
P[oku ∣ hi ⋅ b ⋅ ck ] ≠ P[oku ∣ hj ⋅ b ⋅ ck ]. slight strengthening of the previous supposition), for some γ > 0 a
⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯ ⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯ number smaller than 1/e2 (≈ .135 ; where e’ is the base of the natural
As a result, EQI[cn ∣ hi /hj ∣ b] ≥ 0; and EQI[cn ∣ hi /hj ∣ b] > 0 if and
logarithm), suppose that for each possible outcome oku of each
only if at least one experiment or observation ck has at least one
observation condition ck in cn , either P[oku ∣ hi ⋅ b ⋅ ck ] = 0 or
possible outcome oku such that
P[oku ∣ hj ⋅ b ⋅ ck ]
P[oku ∣ hi ⋅ b ⋅ ck ] ≠ P[oku ∣ hj ⋅ b ⋅ ck ]. ≥ γ.
P[oku ∣ hi ⋅ b ⋅ ck ]

78 Stanford Encyclopedia of Philosophy Spring 2021 Edition 79


Inductive Logic James Hawthorne

And suppose that the Independent Evidence Conditions hold for of ε you may choose). The likelihood of getting such an evidential
evidence stream cn with respect to each of these hypotheses. Now, outcome en is quite close to 1—i.e., no more than the amount
choose any positive ε < 1, as small as you like, but large enough (for
1 (log γ)2
the number of observations n being contemplated) that the value of ×
⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯ (log ε) 2
(EQI[c ∣ hi /hj ∣ b] + n )
n n
⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯ (log ε)
EQI[cn ∣ hi /hj ∣ b] > − .
n
below 1. (Notice that this amount below 1 goes to 0 as n increases.)
Then:
It turns out that in almost every case (for almost any pair of hypotheses)
P[en ∣ hj ⋅ b ⋅ cn ]
[ { } ]
P ∨ en : < ε ∣ hi ⋅ b ⋅ cn the actual likelihood of obtaining such evidence (i.e., evidence that has a
P[en ∣ hi ⋅ b ⋅ cn ] likelihood ratio value less than ε) will be much closer to 1 than this factor
indicates.[16] Thus, the theorem provides an overly cautious lower bound
1 (log γ)2
>1− × ⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯ on the likelihood of obtaining small likelihood ratios. It shows that the
⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯
n (EQI[cn ∣ hi /hj ∣ b] + (log ε)/n)2 larger the value of EQI for an evidence stream, the more likely that stream
is to produce a sequence of outcomes that yield a very small likelihood
For ε = 1/2m and γ = 1/2q , this formula becomes, ⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯
ratio value. But even if EQI remains quite small, a long enough evidence
P[en ∣ hj ⋅ b ⋅ cn ]
[ { } ]
stream, n, of such low-grade evidence will, nevertheless, almost surely
P ∨ en : < 1/2m ∣ hi ⋅ b ⋅ cn
P[en ∣ hi ⋅ b ⋅ cn ] produce an outcome sequence having a very small likelihood ratio value.
[17]
1 q2
> 1 − × ⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯
n (EQI[cn ∣ hi /hj ∣ b] − (m/n))2 Notice that the antecedent condition of the theorem, that “either

(For proof see the supplement Proof of the Probabilistic Refutation P[oku ∣ hi ⋅ b ⋅ ck ] = 0
Theorem.)
or
This theorem provides sufficient conditions for the likely refutation of
P[oku ∣ hj ⋅ b ⋅ ck ]
false alternatives via exceeding small likelihood ratios. The conditions ≥ γ,
under which this happens characterize the degree to which the hypotheses P[oku ∣ hi ⋅ b ⋅ ck ]
involved are empirically distinct from one another. The theorem says that for some γ > 0 but less than 1/e2 (≈ .135 )”, does not favor hypothesis hi
when these conditions are met, according to hypothesis hi (taken together over hj in any way. The condition only rules out the possibility that some
with b ⋅ cn ), the likelihood is near 1 that that one of the outcome sequence outcomes might furnish extremely strong evidence against hj relative to
en will occur for which the likelihood ratio is smaller than ε (for any value hi —by making P[oku ∣ hi ⋅ b ⋅ ck ] = 0 or by making
P[ ku ∣ hj ⋅ b ⋅ k]

80 Stanford Encyclopedia of Philosophy Spring 2021 Edition 81


Inductive Logic James Hawthorne

P[oku ∣ hj ⋅ b ⋅ ck ] The point of the two Convergence Theorems explored in this section is to
P[oku ∣ hi ⋅ b ⋅ ck ] assure us, in advance of the consideration of any specific pair of
hypotheses, that if the possible evidence streams that test them have
less than some quite small γ. This condition is only needed because our
certain characteristics which reflect their evidential distinguishability, it is
measure of evidential distinguishability, QI, blows up when the ratio
highly likely that outcomes yielding small likelihood ratios will result.
P[oku ∣ hj ⋅ b ⋅ ck ] These theorems provide finite lower bounds on how quickly convergence
P[oku ∣ hi ⋅ b ⋅ ck ] is likely to occur. Thus, there is no need to wait through some infinitely
long run for convergence to occur. Indeed, for any evidence sequence on
is extremely small. Furthermore, this condition is really no restriction at which the probability distributions are at all well behaved, the actual
all on possible experiments or observations. If ck has some possible likelihood of obtaining outcomes that yield small likelihood ratio values
outcome sentence oku that would make will inevitably be much higher than the lower bounds given by Theorems
P[oku ∣ hj ⋅ b ⋅ ck ] 1 and 2.

P[oku ∣ hi ⋅ b ⋅ ck ] In sum, according to Theorems 1 and 2, each hypothesis hi says, via
(for a given small γ of interest), one may disjunctively lump oku together likelihoods, that given enough observations, it is very likely to dominate
with some other outcome sentence okv for ck . Then, the antecedent its empirically distinct rivals in a contest of likelihood ratios. The true
condition of the theorem will be satisfied, but with the sentence ‘ hypothesis speaks truthfully about this, and its competitors lie. Even a
(oku ∨ okv ) ’ treated as a single outcome. It can be proved that the only sequence of observations with an extremely low average expected quality
⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯
effect of such “disjunctive lumping” is to make EQI smaller than it would of information is very likely to do the job if that evidential sequence is
⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯
otherwise be (whereas larger values of EQI are more desirable). If the too
long enough. Thus (by Equation 9*), as evidence accumulates, the degree
of support for false hypotheses will very probably approach 0, indicating
strongly refuting disjunct oku actually occurs when the experiment or
that they are probably false; and as this happens, (by Equations 10 and 11)
observation ck is conducted, all the better, since this results in a likelihood
the degree of support for the true hypothesis will approach 1, indicating its
ratio
probable truth. Thus, the Criterion of Adequacy (CoA) is satisfied.
P[oku ∣ hj ⋅ b ⋅ ck ]
P[oku ∣ hi ⋅ b ⋅ ck ] 5. When the Likelihoods are Vague or Diverse
smaller than γ on that particular evidential outcome. We merely failed to Up to this point we have been supposing that likelihoods possess objective
take this more strongly refuting possibility into account when computing or agreed numerical values. Although this supposition is often satisfied in
our lower bound on the likelihood that refutation via likelihood ratios scientific contexts, there are important settings where it is unrealistic,
would occur. where hypotheses only support vague likelihood values, and where there is
enough ambiguity in what hypotheses say about evidential claims that the

82 Stanford Encyclopedia of Philosophy Spring 2021 Edition 83


Inductive Logic James Hawthorne

scientific community cannot agree on precise values for the likelihoods of it also makes
evidential claims.[18] Let us now see how the supposition of precise,
Pβ [en ∣ hj ⋅ b ⋅ cn ]
agreed likelihood values may be relaxed in a reasonable way. < 1;
Pβ [en ∣ hi ⋅ b ⋅ cn ]
Recall why agreement, or near agreement, on precise values for
whenever possible outcome sequence en makes
likelihoods is so important to the scientific enterprise. To the extent that
members of a scientific community disagree on the likelihoods, they Pα [en ∣ hj ⋅ b ⋅ cn ]
> 1,
disagree about the empirical content of their hypotheses, about what each Pα [en ∣ hi ⋅ b ⋅ cn ]
hypothesis says about how the world is likely to be. This can lead to
disagreement about which hypotheses are refuted or supported by a given it also makes
body of evidence. Similarly, to the extent that the values of likelihoods are Pβ [en ∣ hj ⋅ b ⋅ cn ]
> 1;
Pβ [en ∣ hi ⋅ b ⋅ cn ]
only vaguely implied by hypotheses as understood by an individual agent,
that agent may be unable to determine which of several hypotheses is
refuted or supported by a given body of evidence. each of these likelihood ratios is either close to 1 for both of these
support functions, or is quite far from 1 for both of them.[19]
We have seen, however, that the individual values of likelihoods are not
really crucial to the way evidence impacts hypotheses. Rather, as When this condition holds, the evidence will support hi over hj according
Equations 9–11 show, it is ratios of likelihoods that do the heavy lifting. to Pα just in case it does so for Pβ as well, although the strength of support
So, even if two support functions Pα and Pβ disagree on the values of may differ. Furthermore, although the rate at which the likelihood ratios
individual likelihoods, they may, nevertheless, largely agree on the increase or decrease on a stream of evidence may differ for the two
refutation or support that accrues to various rival hypotheses, provided support functions, the impact of the cumulative evidence should ultimately
that the following condition is satisfied: affect their refutation or support in much the same way.

Directional Agreement Condition: When likelihoods are vague or diverse, we may take an approach similar
The likelihood ratios due to each of a pair of support functions Pα and to that we employed for vague and diverse prior plausibility assessments.
Pβ are said to agree in direction (with respect to the possible outcomes We may extend the vagueness sets for individual agents to include a
of experiments or observations relevant to a pair of hypotheses) just in collection of inductive support functions that cover the range of values for
case likelihood ratios of evidence claims (as well as cover the ranges of
comparative support strengths for hypotheses due to plausibility
whenever possible outcome sequence en makes arguments within b, as represented by ratios of prior probabilities).
Pα [en ∣ hj ⋅ b ⋅ cn ] Similarly, we may extend the diversity sets for communities of agents to
< 1,
Pα [en ∣ hi ⋅ b ⋅ cn ] include support functions that cover the ranges of likelihood ratio values

84 Stanford Encyclopedia of Philosophy Spring 2021 Edition 85


Inductive Logic James Hawthorne

that arise within the vagueness sets of members of the scientific community comes to agree on the refutation of these competitors, and the
community. true hypothesis rises to the top of the heap.[20]

This broadening of vagueness and diversity sets to accommodate vague What if the true hypothesis has evidentially equivalent rivals? Their
and diverse likelihood values makes no trouble for the convergence to posterior probabilities must rise as well. In that case we are only assured
truth results for hypotheses. For, provided that the Directional Agreement that the disjunction of the true hypothesis with its evidentially equivalent
Condition is satisfied by all support functions in an extended vagueness or rivals will be driven to 1 as evidence lays low its evidentially distinct
diversity set under consideration, the Likelihood Ratio Convergence rivals. The true hypothesis will itself approach 1 only if either it has no
Theorem applies to each individual support function in that set. For, the evidentially equivalent rivals, or whatever equivalent rivals it does have
the proof of that convergence theorem doesn’t depend on the supposition can be laid low by plausibility arguments of a kind that don’t depend on
that likelihoods are objective or have intersubjectively agreed values. the evidential likelihoods, but only show up via the comparative
Rather, it applies to each individual support function Pα . The only possible plausibility assessments represented by ratios of prior probabilities.
problem with applying this result across a range of support functions is
that when their values for likelihoods differ, function Pα may disagree List of Supplements
with Pβ on which of the hypotheses is favored by a given sequence of
evidence. That can happen because different support functions may Enumerative Inductions: Bayesian Estimation and Convergence
represent the evidential import of hypotheses differently, by specifying Some Prominent Approaches to the Representation of Uncertain
different likelihood values for the very same evidence claims. So, an Inference
evidence stream that favors hi according to Pα may instead favor hj Likelihood Ratios, Likelihoodism, and the Law of Likelihood
according to Pβ . However, when the Directional Agreement Condition Immediate Consequences of the Independent Evidence Conditions
holds for a given collection of support functions, this problem cannot Proof of the Falsification Theorem
arise. Directional Agreement means that the evidential import of Proof that the EQI for cn is the sum of EQI for the individual ck
hypotheses is similar enough for Pα and Pβ that a sequence of outcomes The Effect on EQI of Partitioning the Outcome Space More Finely—
may favor a hypothesis according to Pα only if it does so for Pβ as well. Including Proof of the Nonnegativity of EQI
Proof of the Probabilistic Refutation Theorem
Thus, when the Directional Agreement Condition holds for all support
functions in a vagueness or diversity set that is extended to include vague Bibliography
or diverse likelihoods, and provided that enough evidentially
distinguishing experiments or observations can be performed, all support Boole, George, 1854, The Laws of Thought, London: MacMillan.
functions in the extended vagueness or diversity set will very probably Republished in 1958 by Dover: New York.
come to agree that the likelihood ratios for empirically distinct false Bovens, Luc and Stephan Hartmann, 2003, Bayesian Epistemology,
competitors of a true hypothesis are extremely small. As that happens, the Oxford: Oxford University Press. doi:10.1093/0199269750.001.0001

86 Stanford Encyclopedia of Philosophy Spring 2021 Edition 87


Inductive Logic James Hawthorne

Carnap, Rudolf, 1950, Logical Foundations of Probability, Chicago: Chevalier et Riviere; translated by P.P. Wiener, The Aim and
University of Chicago Press. Structure of Physical Theory, Princeton, NJ: Princeton University
–––, 1952, The Continuum of Inductive Methods, Chicago: University of Press, 1954.
Chicago Press. Earman, John, 1992, Bayes or Bust? A Critical Examination of Bayesian
–––, 1963, “Replies and Systematic Expositions”, in The Philosophy of Confirmation Theory, Cambridge, MA: MIT Press.
Rudolf Carnap, Paul Arthur Schilpp (ed.),La Salle, IL: Open Court. Edwards, A.W.F., 1972, Likelihood: an account of the statistical concept
Chihara, Charles S., 1987, “Some Problems for Bayesian Confirmation of likelihood and its application to scientific inference, Cambridge:
Theory”, British Journal for the Philosophy of Science, 38(4): 551– Cambridge University Press.
560. doi:10.1093/bjps/38.4.551 Edwards, Ward, Harold Lindman, and Leonard J. Savage, 1963, “Bayesian
Christensen, David, 1999, “Measuring Confirmation”, Journal of Statistical Inference for Psychological Research”, Psychological
Philosophy, 96(9): 437–61. doi:10.2307/2564707 Review, 70(3): 193–242. doi:10.1037/h0044139
–––, 2004, Putting Logic in its Place: Formal Constraints on Rational Eells, Ellery, 1985, “Problems of Old Evidence”, Pacific Philosophical
Belief, Oxford: Oxford University Press. Quarterly, 66(3–4): 283–302. doi:10.1111/j.1468-
doi:10.1093/0199263256.001.0001 0114.1985.tb00254.x
De Finetti, Bruno, 1937, “La Prévision: Ses Lois Logiques, Ses Sources –––, 2006, “Confirmation Theory”, Sarkar and Pfeifer 2006..
Subjectives”, Annales de l’Institut Henri Poincaré, 7: 1–68; Eells, Ellery and Branden Fitelson, 2000, “Measuring Confirmation and
translated by Henry E. Kyburg, Jr. as “Foresight. Its Logical Laws, Its Evidence”, Journal of Philosophy, 97(12): 663–672.
Subjective Sources”, in Studies in Subjective Probability, Henry E. doi:10.2307/2678462
Kyburg, Jr. and H.E. Smokler (eds.), Robert E. Krieger Publishing Field, Hartry H., 1977, “Logic, Meaning, and Conceptual Role”, Journal
Company, 1980. of Philosophy, 74(7): 379–409. doi:10.2307/2025580
Dowe, David L., Steve Gardner, and Graham Oppy, 2007, “Bayes, Not Fisher, R.A., 1922, “On the Mathematical Foundations of Theoretical
Bust! Why Simplicity is No Problem for Bayesians”, British Journal Statistics”, Philosophical Transactions of the Royal Society, series A ,
for the Philosophy of Science, 58(4): 709–754. 222(594–604): 309–368. doi:10.1098/rsta.1922.0009
doi:10.1093/bjps/axm033 Fitelson, Branden, 1999, “The Plurality of Bayesian Measures of
Dubois, Didier J. and Henri Prade, 1980, Fuzzy Sets and Systems, Confirmation and the Problem of Measure Sensitivity”, Philosophy of
(Mathematics in Science and Engineering, 144), New York: Science, 66: S362–S378. doi:10.1086/392738
Academic Press. –––, 2001, “A Bayesian Account of Independent Evidence with
–––, 1990, “An Introduction to Possibilistic and Fuzzy Logics”, in Glenn Applications”, Philosophy of Science, 68(S3): S123–S140.
Shafer and Judea Pearl (eds.), Readings in Uncertain Reasoning, San doi:10.1086/392903
Mateo, CA: Morgan Kaufmann, 742–761.. –––, 2002, “Putting the Irrelevance Back Into the Problem of Irrelevant
Duhem, P., 1906, La theorie physique. Son objet et sa structure, Paris: Conjunction”, Philosophy of Science, 69(4): 611–622.

88 Stanford Encyclopedia of Philosophy Spring 2021 Edition 89


Inductive Logic James Hawthorne

doi:10.1086/344624 doi:10.1017/CBO9780511817557
–––, 2006, “Inductive Logic”, Sarkar and Pfeifer 2006.. –––, 2001, An Introduction to Probability and Inductive Logic,
–––, 2006, “Logical Foundations of Evidential Support”, Philosophy of Cambridge: Cambridge University Press.
Science, 73(5): 500–512. doi:10.1086/518320 doi:10.1017/CBO9780511801297
–––, 2007, “Likelihoodism, Bayesianism, and Relational Confirmation”, Hájek, Alan, 2003a, “What Conditional Probability Could Not Be”,
Synthese, 156(3): 473–489. doi:10.1007/s11229-006-9134-9 Synthese, 137(3):, 273–323.
Fitelson, Branden and James Hawthorne, 2010, “How Bayesian doi:10.1023/B:SYNT.0000004904.91112.16
Confirmation Theory Handles the Paradox of the Ravens”, in Eells –––, 2003b, “Interpretations of the Probability Calculus”, in the Stanford
and Fetzer (eds.), The Place of Probability in Science, Open Court. Encyclopedia of Philosophy, (Summer 2003 Edition), Edward N.
[Fitelson & Hawthorne 2010 preprint available from the author Zalta (ed.), URL =
(PDF)] <https://plato.stanford.edu/archives/sum2003/entries/probability-
Forster, Malcolm and Elliott Sober, 2004, “Why Likelihood”, in Mark L. interpret/>
Taper and Subhash R. Lele (eds.), The Nature of Scientific Evidence, –––, 2005, “Scotching Dutch Books?” Philosophical Perspectives, 19
Chicago: University of Chicago Press. (Epistemology): 139–151. doi:10.1111/j.1520-8583.2005.00057.x
Friedman, Nir and Joseph Y. Halpern, 1995, “Plausibility Measures: A –––, 2007, “The Reference Class Problem is Your Problem Too”,
User’s Guide”, in UAI 95: Proceedings of the Eleventh Conference Synthese, 156(3): 563–585. doi:10.1007/s11229-006-9138-5
on Uncertainty in Artificial Intelligence, 175–184. Halpern, Joseph Y., 2003, Reasoning About Uncertainty, Cambridge, MA:
Gaifman, Haim and Marc Snir, 1982, “Probabilities Over Rich Languages, MIT Press.
Testing and Randomness”, Journal of Symbolic Logic, 47(3): 495– Harper, William L., 1976, “Rational Belief Change, Popper Functions and
548. doi:10.2307/2273587 Counterfactuals”, in Harper and Hooker 1976: 73–115.
Gillies, Donald, 2000, Philosophical Theories of Probability, London: doi:10.1007/978-94-010-1853-1_5
Routledge. Harper, William L. and Clifford Alan Hooker (eds.), 1976, Foundations of
Glymour, Clark N., 1980, Theory and Evidence, Princeton, NJ: Princeton Probability Theory, Statistical Inference, and Statistical Theories of
University Press. Science, volume I Foundations and Philosophy of Epistemic
Goodman, Nelson, 1983, Fact, Fiction, and Forecast, 4th edition, Applications of Probability Theory, (The Western Ontario Series in
Cambridge, MA: Harvard University Press. Philosophy of Science, 6a), Dordrecht: Reidel. doi:10.1007/978-94-
Hacking, Ian, 1965, Logic of Statistical Inference, Cambridge: Cambridge 010-1853-1
University Press. Hawthorne, James, 1993, “Bayesian Induction is Eliminative Induction”,
–––, 1975, The Emergence of Probability: a Philosophical Study of Early Philosophical Topics, 21(1): 99–138.
Ideas about Probability, Induction and Statistical Inference, doi:10.5840/philtopics19932117
Cambridge: Cambridge University Press. –––, 1994,“On the Nature of Bayesian Convergence”, PSA: Proceedings

90 Stanford Encyclopedia of Philosophy Spring 2021 Edition 91


Inductive Logic James Hawthorne

of the Biennial Meeting of the Philosophy of Science Association Bayesian Approach, La Salle, IL: Open Court. [3rd edition, 2005.]
1994, 1: 241–249. doi:10.1086/psaprocbienmeetp.1994.1.193029 Huber, Franz, 2005a, “Subjective Probabilities as Basis for Scientific
–––, 2005, “Degree-of-Belief and Degree-of-Support: Why Bayesians Reasoning?” British Journal for the Philosophy of Science, 56(1):
Need Both Notions”, Mind, 114(454): 277–320. 101–116. doi:10.1093/phisci/axi105
doi:10.1093/mind/fzi277 –––, 2005b, “What Is the Point of Confirmation?” Philosophy of Science,
–––, 2009, “The Lockean Thesis and the Logic of Belief”, in Franz Huber 72(5): 1146–1159. doi:10.1086/508961
and Christoph Schmidt-Petri (eds.), Degrees of Belief, (Synthese Jaynes, Edwin T., 1968, “Prior Probabilities”, IEEE Transactions on
Library, 342), Dordrecht: Springer, pp. 49–74. doi:10.1007/978-1- Systems Science and Cybernetics, SSC–4(3): 227–241.
4020-9198-8_3 doi:10.1109/TSSC.1968.300117
Hawthorne, James and Luc Bovens, 1999, “The Preface, the Lottery, and Jeffrey, Richard C., 1983, The Logic of Decision, 2nd edition, Chicago:
the Logic of Belief”, Mind, 108(430): 241–264. University of Chicago Press.
doi:10.1093/mind/108.430.241 –––, 1987, “Alias Smith and Jones: The Testimony of the Senses”,
Hawthorne, James and Branden Fitelson, 2004, “Discussion: Re-solving Erkenntnis, 26(3): 391–399. doi:10.1007/BF00167725
Irrelevant Conjunction With Probabilistic Independence”, Philosophy –––, 1992, Probability and the Art of Judgment, New York: Cambridge
of Science, 71(4): 505–514. doi:10.1086/423626 University Press. doi:10.1017/CBO9781139172394
Hellman, Geoffrey, 1997, “Bayes and Beyond”, Philosophy of Science, –––, 2004, Subjective Probability: The Real Thing, Cambridge:
64(2): 191–221. doi:10.1086/392548 Cambridge University Press. doi:10.1017/CBO9780511816161
Hempel, Carl G., 1945, “Studies in the Logic of Confirmation”, Mind, Jeffreys, Harold, 1939, Theory of Probability, Oxford: Oxford University
54(213): 1–26, 54(214):97–121. doi:10.1093/mind/LIV.213.1 Press.
doi:10.1093/mind/LIV.214.97 Joyce, James M., 1998, “A Nonpragmatic Vindication of Probabilism”,
Horwich, Paul, 1982, Probability and Evidence, Cambridge: Cambridge Philosophy of Science, 65(4): 575–603. doi:10.1086/392661
University Press. doi:10.1017/CBO9781316494219 –––, 1999, The Foundations of Causal Decision Theory, New York:
Howson, Colin, 1997, “A Logic of Induction”, Philosophy of Science, Cambridge University Press. doi:10.1017/CBO9780511498497
64(2): 268–290. doi:10.1086/392551 –––, 2003, “Bayes’ Theorem”, in the Stanford Encyclopedia of
–––, 2000, Hume’s Problem: Induction and the Justification of Belief, Philosophy, (Summer 2003 Edition), Edward N. Zalta (ed.), URL =
Oxford: Oxford University Press. doi:10.1093/0198250371.001.0001 <https://plato.stanford.edu/archives/win2003/entries/bayes-theorem/>
–––, 2002, “Bayesianism in Statistics“, in Swinburne 2002: 39–71. –––, 2004, “Bayesianism”, in Alfred R. Mele and Piers Rawling (eds.),
doi:10.5871/bacad/9780197263419.003.0003 The Oxford Handbook of Rationality, Oxford: Oxford University
–––, 2007, “Logic With Numbers”, Synthese, 156(3): 491–512. Press, pp. 132–153. doi:10.1093/0195145399.003.0008
doi:10.1007/s11229-006-9135-8 –––, 2005, “How Probabilities Reflect Evidence”, Philosophical
Howson, Colin and Peter Urbach, 1993, Scientific Reasoning: The Perspectives, 19: 153–179. doi:10.1111/j.1520-8583.2005.00058.x

92 Stanford Encyclopedia of Philosophy Spring 2021 Edition 93


Inductive Logic James Hawthorne

Kaplan, Mark, 1996, Decision Theory as Philosophy, Cambridge: Lenhard Johannes, 2006, “Models and Statistical Inference: The
Cambridge University Press. Controversy Between Fisher and Neyman-Pearson”, British Journal
Kelly, Kevin T., Oliver Schulte, and Cory Juhl, 1997, “Learning Theory for the Philosophy of Science, 57(1): 69–91. doi:10.1093/bjps/axi152
and the Philosophy of Science”, Philosophy of Science, 64(2): 245– Levi, Isaac, 1967, Gambling with Truth: An Essay on Induction and the
267. doi:10.1086/392550 Aims of Science, New York: Knopf.
Keynes, John Maynard, 1921, A Treatise on Probability, London: –––, 1977, “Direct Inference”, Journal of Philosophy, 74(1): 5–29.
Macmillan and Co. doi:10.2307/2025732
Kolmogorov, A.N., 1956, Foundations of the Theory of Probability –––, 1978, “Confirmational Conditionalization”, Journal of Philosophy,
(Grundbegriffe der Wahrscheinlichkeitsrechnung, 2nd edition, New 75(12): 730–737. doi:10.2307/2025516
York: Chelsea Publishing Company. –––, 1980, The Enterprise of Knowledge: An Essay on Knowledge, Credal
Koopman, B.O., 1940, “The Bases of Probability”, Bulletin of the Probability, and Chance, Cambridge, MA: MIT Press.
American Mathematical Society, 46(10): 763–774. Reprinted in H. Lewis, David, 1980, “A Subjectivist’s Guide to Objective Chance”, in
Kyburg and H. Smokler (eds.), 1980, Studies in Subjective Richard C. Jeffrey, (ed.), Studies in Inductive Logic and Probability,
Probability, 2nd edition, Huntington, NY: Krieger Publ. Co. vol. 2, Berkeley: University of California Press, 263–293.
[Koopman 1940 available online] Maher, Patrick, 1993, Betting on Theories, Cambridge: Cambridge
Kyburg, Henry E., Jr., 1974, The Logical Foundations of Statistical University Press.
Inference, Dordrecht: Reidel. doi:10.1007/978-94-010-2175-3 –––, 1996, “Subjective and Objective Confirmation”, Philosophy of
–––, 1977, “Randomness and the Right Reference Class”, Journal of Science, 63(2): 149–174. doi:10.1086/289906
Philosophy, 74(9): 501–520. doi:10.2307/2025794 –––, 1997, “Depragmatized Dutch Book Arguments”, Philosophy of
–––, 1978, “An Interpolation Theorem for Inductive Relations”, Journal of Science, 64(2): 291–305. doi:10.1086/392552
Philosophy, 75:93–98. –––, 1999, “Inductive Logic and the Ravens Paradox”, Philosophy of
–––, 2006, “Belief, Evidence, and Conditioning”, Philosophy of Science, Science, 66(1): 50–70. doi:10.1086/392676
73(1): 42–65. doi:10.1086/510174 –––, 2004, “Probability Captures the Logic of Scientific Confirmation”, in
Lange, Marc, 1999, “Calibration and the Epistemological Role of Christopher Hitchcock (ed.), Contemporary Debates in Philosophy of
Bayesian Conditionalization”, Journal of Philosophy, 96(6): 294– Science, Oxford: Blackwell, 69–93.
324. doi:10.2307/2564680 –––, 2005, “Confirmation Theory”, The Encyclopedia of Philosophy, 2nd
–––, 2002, “Okasha on Inductive Scepticism”, The Philosophical edition, Donald M. Borchert (ed.), Detroit: Macmillan.
Quarterly, 52(207): 226–232. doi:10.1111/1467-9213.00264 –––, 2006a, “The Concept of Inductive Probability”, Erkenntnis, 65(2):
Laudan, Larry, 1997, “How About Bust? Factoring Explanatory Power 185–206. doi:10.1007/s10670-005-5087-5
Back into Theory Evaluation”, Philosophy of Science, 64(2): 206– –––, 2006b, “A Conception of Inductive Logic”, Philosophy of Science,
216. doi:10.1086/392553 73(5): 513–523. doi:10.1086/518321

94 Stanford Encyclopedia of Philosophy Spring 2021 Edition 95


Inductive Logic James Hawthorne

–––, 2010, “Bayesian Probability”, Synthese, 172(1): 119–127. Quine, W.V., 1953, “Two Dogmas of Empiricism”, in From a Logical
doi:10.1007/s11229-009-9471-6 Point of View, New York: Harper Torchbooks. Routledge
Mayo, Deborah G., 1996, Error and the Growth of Experimental Encyclopedia of Philosophy, Version 1.0, London: Routledge
Knowledge, Chicago: University of Chicago Press. Ramsey, F.P., 1926, “Truth and Probability”, in Foundations of
–––, 1997, “Duhem’s Problem, the Bayesian Way, and Error Statistics, or Mathematics and other Essays, R.B. Braithwaite (ed.), Routledge &
‘What’s Belief Got to do with It?’”, Philosophy of Science, 64(2): P. Kegan,1931, 156–198. Reprinted in Studies in Subjective
222–244. doi:10.1086/392549 Probability, H. Kyburg and H. Smokler (eds.), 2nd ed., R.E. Krieger
Mayo Deborah and Aris Spanos, 2006, “Severe Testing as a Basic Publishing Company, 1980, 23–52. Reprinted in Philosophical
Concept in a Neyman-Pearson Philosophy of Induction“, British Papers, D.H. Mellor (ed.), Cambridge: University Press, Cambridge,
Journal for the Philosophy of Science, 57(2): 323–357. 1990,
doi:10.1093/bjps/axl003 Reichenbach, Hans, 1938, Experience and Prediction: An Analysis of the
McGee, Vann, 1994, “Learning the Impossible”, in E. Eells and B. Skyrms Foundations and the Structure of Knowledge, Chicago: University of
(eds.), Probability and Conditionals: Belief Revision and Rational Chicago Press.
Decision, New York: Cambridge University Press, 179–200. Rényi, Alfred, 1970, Foundations of Probability, San Francisco, CA:
McGrew, Timothy J., 2003, “Confirmation, Heuristics, and Explanatory Holden-Day.
Reasoning”, British Journal for the Philosophy of Science, 54: 553– Rosenkrantz, R.D., 1981, Foundations and Applications of Inductive
567. Probability, Atascadero, CA: Ridgeview Publishing.
McGrew, Lydia and Timothy McGrew, 2008, “Foundationalism, Roush, Sherrilyn , 2004, “Discussion Note: Positive Relevance
Probability, and Mutual Support”, Erkenntnis, 68(1): 55–77. Defended”, Philosophy of Science, 71(1): 110–116.
doi:10.1007/s10670-007-9062-1 doi:10.1086/381416
Neyman, Jerzy and E.S. Pearson, 1967, Joint Statistical Papers, –––, 2006, “Induction, Problem of”, Sarkar and Pfeifer 2006..
Cambridge: Cambridge University Press. –––, 2006, Tracking Truth: Knowledge, Evidence, and Science, Oxford:
Norton, John D., 2003, “A Material Theory of Induction”, Philosophy of Oxford University Press.
Science, 70(4): 647–670. doi:10.1086/378858 Royall, Richard M., 1997, Statistical Evidence: A Likelihood Paradigm,
–––, 2007, “Probability Disassembled”, British Journal for the Philosophy New York: Chapman & Hall/CRC.
of Science, 58(2): 141–171. doi:10.1093/bjps/axm009 Salmon, Wesley C., 1966, The Foundations of Scientific Inference,
Okasha, Samir, 2001, “What Did Hume Really Show About Induction?”, Pittsburgh, PA: University of Pittsburgh Press.
The Philosophical Quarterly, 51(204): 307–327. doi:10.1111/1467- –––, 1975, “Confirmation and Relevance”, in H. Feigl and G. Maxwell
9213.00231 (eds.), Induction, Probability, and Confirmation, (Minnesota Studies
Popper, Karl, 1968, The Logic of Scientific Discovery, 3rd edition, London: in the Philosophy of Science, 6), Minneapolis: University of
Hutchinson. Minnesota Press, 3–36.

96 Stanford Encyclopedia of Philosophy Spring 2021 Edition 97


Inductive Logic James Hawthorne

Sarkar, Sahotra and Jessica Pfeifer (eds.), 2006, The Philosophy of doi:10.1023/B:SYNT.0000044991.73791.f7
Science: An Encyclopedia, 2 volumes, New York: Routledge. Suppes, Patrick, 2007, “Where do Bayesian Priors Come From?”
Savage, Leonard J., 1954, The Foundations of Statistics, John Wiley (2nd Synthese, 156(3): 441–471. doi:10.1007/s11229-006-9133-x
ed., New York: Dover 1972). Swinburne, Richard, 2002, Bayes’ Theorem, Oxford: Oxford University
Savage, Leonard J., et al., 1962, The Foundations of Statistical Inference, Press. doi:10.5871/bacad/9780197263419.001.0001
London: Methuen. Talbot, W., 2001, “Bayesian Epistemology”, in the Stanford Encyclopedia
Schlesinger, George N., 1991, The Sweep of Probability, Notre Dame, IN: of Philosophy, (Fall 2001 Edition), Edward N. Zalta (ed.), URL =
Notre Dame University Press. <https://plato.stanford.edu/archives/fall2001/entries/epistemology-
Seidenfeld, Teddy, 1978, “Direct Inference and Inverse Inference”, bayesian/>
Journal of Philosophy, 75(12): 709–730. doi:10.2307/2025515 Teller, Paul, 1976, “Conditionalization, Observation, and Change of
–––, 1992, “R.A. Fisher’s Fiducial Argument and Bayes’ Theorem”, Preference”, in Harper and Hooker 1976: 205–259. doi:10.1007/978-
Statistical Science, 7(3): 358–368. doi:10.1214/ss/1177011232 94-010-1853-1_9
Shafer, Glenn, 1976, A Mathematical Theory of Evidence, Princeton, NJ: Van Fraassen, Bas C., 1983, “Calibration: A Frequency Justification for
Princeton University Press. Personal Probability ”, in R.S. Cohen and L. Laudan (eds.), Physics,
–––, 1990, “Perspectives on the Theory and Practice of Belief Functions”, Philosophy, and Psychoanalysis: Essays in Honor of Adolf
International Journal of Approximate Reasoning, 4(5–6): 323–362. Grunbaum, Dordrecht: Reidel. doi:10.1007/978-94-009-7055-7_15
doi:10.1016/0888-613X(90)90012-Q Venn, John, 1876, The Logic of Chance, 2nd ed., Macmillan and co;
Skyrms, Brian, 1984, Pragmatics and Empiricism, New Haven, CT: Yale reprinted, New York, 1962.
University Press. Vineberg, Susan, 2006, “Dutch Book Argument”, Sarkar and Pfeifer
–––, 1990, The Dynamics of Rational Deliberation, Cambridge, MA: 2006..
Harvard University Press. Vranas, Peter B.M., 2004, “Hempel’s Raven Paradox: A Lacuna in the
–––, 2000, Choice and Chance: An Introduction to Inductive Logic, 4th Standard Bayesian Solution”, British Journal for the Philosophy of
edition, Belmont, CA: Wadsworth, Inc. Science, 55(3): 545–560. doi:10.1093/bjps/55.3.545
Sober, Elliott, 2002, “Bayesianism—Its Scope and Limits”, in Swinburne Weatherson, Brian, 1999, “Begging the Question and Bayesianism”,
2002: 21–38. doi:10.5871/bacad/9780197263419.003.0002 Studies in History and Philosophy of Science [Part A], 30(4): 687–
Spohn, Wolfgang, 1988, “Ordinal Conditional Functions: A Dynamic 697. doi:10.1016/S0039-3681(99)00020-5
Theory of Epistemic States”, in William L. Harper and Brian Skyrms Williamson, Jon, 2007, “Inductive Influence”, British Journal for
(eds.), Causation in Decision, Belief Change, and Statistics, vol. 2, Philosophy of Science, 58(4): 689–708. doi:10.1093/bjps/axm032
Dordrecht: Reidel, 105–134. doi:10.1007/978-94-009-2865-7_6 Zadeh, Lotfi A., 1965, “Fuzzy Sets”, Information and Control, 8(3): 338–
Strevens, Michael, 2004, “Bayesian Confirmation Theory: Inductive 353. doi:10.1016/S0019-9958(65)90241-X
Logic, or Mere Inductive Framework?” Synthese, 141(3): 365–379. –––, 1978, “Fuzzy Sets as a Basis for a Theory of Possibility”, Fuzzy Sets

98 Stanford Encyclopedia of Philosophy Spring 2021 Edition 99


Inductive Logic James Hawthorne

and Systems, vol. 1, 3–28. extensive list of links to readings. The Notes, Handouts, & Links
page has Fitelson’s weekly course notes and some links to useful
Academic Tools internet resources on confirmation theory.
Fitelson’s course on Probability and Induction. Main page of Branden
How to cite this entry. Fitelson’s course on Probability and Induction. The Syllabus provides
Preview the PDF version of this entry at the Friends of the SEP an extensive list of links to readings on the subject. The Notes &
Society. Handouts page has Fitelson’s powerpoint slides for each of his
Look up this entry topic at the Internet Philosophy Ontology
lectures and some links to handouts for the course. The Links page
Project (InPhO).
contains links to some useful internet resources.
Enhanced bibliography for this entry at PhilPapers, with links
to its database.
Related Entries
Other Internet Resources
Bayes’ Theorem | epistemology: Bayesian | probability, interpretations of
Confirmation and Induction. Really nice overview by Franz Huber in
the Internet Encyclopedia of Philosophy. Acknowledgments
Inductive Logic, (in PDF), by Branden Fitelson, Philosophy of
Thanks to Alan Hájek, Jim Joyce, and Edward Zalta for many valuable
Science: An Encyclopedia, (J. Pfeifer and S. Sarkar, eds.), Routledge.
comments and suggestions. The editors and author also thank Greg
An extensive encyclopedia article on inductive logic.
Stokley and Philippe van Basshuysen for carefully reading an earlier
Teaching Theory of Knowledge: Probability and Induction. A very
version of the entry and identifying a number of typographical errors.
extensive outline of issues in Probability and Induction, each topic
accompanied by a list of relevant books and articles (without links),
compiled by Brad Armendt and Martin Curd.
Enumerative Inductions: Bayesian Estimation and
Bayesian Networks Without Tears, (in PDF), by Eugene Charniak Convergence
(Computer Science and Cognitive Science, Brown University). An
In this section we’ll see that for the special case of enumerative inductions
introductory article on Bayesian inference.
probabilistic inductive logic satisfies the Criterion of Adequacy (CoA)
Miscellany of Works on Probabilistic Thinking. A collection of on-
stated at the beginning of this article. That is, under some plausible
line articles on Subjective Probability and probabilistic reasoning by
conditions, given a reasonable amount of evidence, the degree to which
Richard Jeffrey and by several other philosophers writing on related
that evidence comes to support a hypothesis through enumerative
issues.
induction is very likely to approach 1 for true hypotheses. We will now see
Fitelson’s course on Confirmation Theory. Main page of Branden
precisely how this works.
Fitelson’s course on Confirmation Theory. The Syllabus provides an

100 Stanford Encyclopedia of Philosophy Spring 2021 Edition 101


Inductive Logic James Hawthorne

Recall that in enumerative inductions the idea is to infer the proportion, or m/n? And to what extent does that bound depend on the prior probabilities
relative frequency, of an attribute in a population from how frequently the of the various possible alternative frequency hypotheses. More generally,
attribute occurs in a sample of the population. Examples 1 and 2 at the for a given vagueness or diversity set, what bounds can we place on the
beginning of the article describe two such inferences. Enumerative values of p.
induction is only a rather special case of inductive inference. However,
such inferences are very common, and so worthy of careful attention. Put more formally, we are asking for what values of p and q does the
They arise, for example, in the context of polling and in many other cases following inequality hold:
where a population frequency is estimated from a sample. We will ⎡ (m/n) − q < F[A, B] < (m/n) + q ⎤
⎢ ⎥
∣ b ⋅ F[A, S] = m/n
Pα ⎢⎢ ⎥ > p?
establish conditions under which such inferences give rise to highly
objective posterior probabilities, posterior probabilities that are extremely ⋅ Rnd[S, ⎥
⎢ B, A] ⎥
stable over a wide range of reasonable prior plausibility assessments. That ⎣ ⋅ Size[S] = n ⎦
is, we will consider all of the inductive support functions in an agent’s
vagueness set V or in a community’s diversity set D. We will see that It turns out that we need only a very weak supposition about the values of
under some very weak suppositions about the make up of V or of D, a prior probabilities of support functions Pα in vagueness or diversity sets to
reasonable amount of data will bring all of the support functions in these legitimize such inferences, an supposition that almost always holds in the
sets to agree that the posterior degree of support for a particular frequency context of enumerative inductions.
hypothesis is very close to 1. And, we will see, it is very likely these
Boundedness Assumption for Estimation:
support functions will converge to agreement on a true hypothesis.
There is a region R of possible values near the sample frequency m/n
1. Convergence to Agreement (e.g., R is the region between (m/n) − q and (m/n) + q , for some
2. Convergence to the Truth margin of error q of interest) such that no frequency hypothesis
3 Tighter Bounds on the Margin of Error outside of region R is overwhelmingly more initially plausible than
those frequency hypotheses inside of region R.

1. Convergence to Agreement What does it mean for no frequency hypothesis outside of region R to be
overwhelmingly more initially plausible than those frequency hypotheses
Suppose we want to know the frequency with which attribute A occurs inside of region R (where R is some specific region in which the sample
among members of population B. We randomly select a sample S from B frequency, F[A, S] = m/n, lies)? The main idea is that there is some
consisting of n members, and find that it contains m members that exhibit (perhaps very large) bound K on how much more plausible frequency
attribute A.[21] On the basis of this evidence, what is the posterior hypotheses outside of region R may be than those frequency hypotheses
probability p of the hypothesis that the true proportion (or frequency) of inside of region R. We state this condition carefully by considering two
As among Bs is within a given region R around the sample proportion kinds of cases, depending on whether or not the population B is known to

102 Stanford Encyclopedia of Philosophy Spring 2021 Edition 103


Inductive Logic James Hawthorne

be bounded in size by some specific (perhaps overly large) integer u. (The for each integer k from 0 through u.
first case will be simpler because it doesn’t suppose that the support
functions involved may be characterized by probability density functions, Case 2. Alternatively, suppose that there is no positive integer u at
while the second case does suppose this.) least as large as the size of population B that satisfies the conditions of
case 1. But suppose that the prior probabilities of the various
Case 1. Suppose the size of the population B is finite. We need not competing hypotheses can be represented (at least very nearly) by a
know how big B is. We merely suppose that for some positive integer probability density function pα [F[A, B] = r ∣ b] —i.e., for any specific
u that is at least as large as the size of B, but might well be many times values v and u, the value of
larger, the following condition holds for all support functions Pα in the u

∫v
vagueness or diversity set under consideration. Pα [v < F[A, B] < u ∣ b] = pα [F[A, B] = r ∣ b]dr,

There is some specific positive factor K (possibly very large, perhaps


or at least very nearly so. Then we just need the following condition to
as large as 1000, or larger) such that for any pair of hypotheses of
be satisfied by all support functions Pα in the vagueness or diversity
form F[A, B] = v/u inside region R and of form F[A, B] = w/u
set under consideration.
outside of region R (where u, v, and w are non-negative integers), the
hypothesis outside of region R is no more than K times more plausible There is some specific positive factor K (possibly very large, perhaps
than the hypothesis within region R (given plausibility consideration as large as 1000, or larger) such that for any pair of hypotheses of
within b)—i.e., for all ratios v/u inside region R and all ratios v/u form F[A, B] = r inside region R and of form F[A, B] = s outside of
outside region R, region R (where r and s are non-negative real numbers no larger than
Pα [F[A, B] = w/u ∣ b] 1), the value of the probability density function for the hypothesis
≤ K.
Pα [F[A, B] = v/u ∣ b] outside of region R is no more than K times larger than the value of the
probability density function for the hypothesis within region R (given
For Case 1 we also assume (as seems reasonable) that in the absence of plausibility consideration within b), where the density function within
information about the observed sample frequency, the claim region R is never less than some (perhaps very tiny) positive lower
‘Random[S, B, A] ⋅ Size[S] = n’, that the sample is randomly selected and bound—i.e., for all values r inside region R and all values s outside
of size n, should be irrelevant to the initial plausibilities of possible region R,
population frequencies—i.e., we suppose that
pα [F[A, B] = s ∣ b]
≤ K,
⎡ F[A, B] = k/u ⎤ pα [F[A, B] = r ∣ b]
Pα ⎢⎢ ∣ Rnd[S, B, A] ⋅ Size[S] = n ⎥ = P [F[A, B] = k/u ∣ b]
⎥ where for all r within region R, pα [F[A, B] = r ∣ b] ≥ g for some
α
⎣ ⋅b ⎦
small g > 0 .

104 Stanford Encyclopedia of Philosophy Spring 2021 Edition 105


Inductive Logic James Hawthorne

For Case 2 we also assume (as seems reasonable) that in the absence of For any given region R containing sample frequencies m/n, this lower
information about the observed sample frequency, the claim bound approaches 1 rapidly as n increases.
‘Random[S, B, A] ⋅ Size[S] = n’, that the sample is randomly selected and
of size n, should be irrelevant to the initial plausibilities of possible The expression ‘β[R, m + 1, n − m + 1]’ represents the beta-distribution
population frequencies—i.e., in particular, we suppose that for each function with parameters m + 1 and n − m + 1 evaluated over region R.
probability density function pα under consideration, By definition

⎡ F[A, B] = q ⎤ ∫R rm (1 − r)n−m dr
β[R, m + 1, n − m + 1] = .
pα ⎢⎢ ∣ Rnd[S, B, A] ⋅ Size[S] = n ⎥ = p [F[A, B] = q ∣ b]
⎥ α
1
∫0 rm (1 − r)n−m dr
⎣ ⋅b ⎦
When region R contains an interval around m/n , the value of this function
for real numbers q from 0 through 1. is a fraction that approaches 1 for large n. In a moment we will see some
concrete illustrations of the implications of this theorem for specific values
When either of these two Cases hold, let us say that for the support
of m and n and specific regions R.
functions Pα in the vagueness or diversity sets under consideration, the
prior probabilities are K bounded with respect to region R. Then we have The values of the beta-distribution function may be easily computed using
the following theorem about enumerative inductions, which shows that the a version of the function supplied with most mathematics and spreadsheet
posterior probability that the true frequency must lie within a small region programs. The version of the function supplied by such programs usually
R around the sample frequency m/n quickly approaches 1 as the sample takes the form BETADIST(x, y, z) , which computes the value of the beta
size n becomes large. distribution from 0 up to to x, and where y and z are the “parameters of the
distribution”. For our purposes, where the sample S of n selections from B
Theorem: Frequency Estimation Theorem:[22]
yields m that exhibit As, these parameters need to be m + 1 and
Suppose, for all support functions Pα in the vagueness or diversity set
n − m + 1. So if the region R of interest (wherein the sample frequency
under consideration, the prior probabilities are K bounded with respect
m/n lies) is between the values v and u (where v is the lower bound on
to region R, where region R contains the fraction m/n (for positive
integer n and non-negative integer m ≤ n) . Then, for all support
region R and u is the upper bound on region R), then the program should
be asked to compute the value of
functions Pα in the vagueness or diversity set,
∫v rm (1 − r)n−m dr
u
⎡ F[A, B] ∈ R ⎤ β[R, m + 1, n − m + 1] =
⎢ ⎥ 1
∫0 rm (1 − r)n−m dr
∣ b ⋅ F[A, S] = m/n 1
Pα ⎢⎢ ⎥≥
⎥ .
⋅ Rnd[S, B, A] ⎥ 1 + K × [( β[R,m+1,n−m+1]
⎢ ) − 1]
1
⎣ ⋅ Size[S] = n ⎦ by having it compute

BETADIST[u, m + 1, n − m + 1] − BETADIST[v, m + 1, n − m + 1].

106 Stanford Encyclopedia of Philosophy Spring 2021 Edition 107


Inductive Logic James Hawthorne

So, to have your mathematics or spreadsheet program compute a lower The first section of this article provided two examples of enumerative
bound on the value of inductive inferences. Consider Example 1. Let ‘B’ represent the

⎡ v ≤ F[A, B] ≤ u ⎤
population of all ravens. Let ‘A’ represent the class of black ravens. Now
⎢ ⎥ consider those hypotheses of form ‘F[A, B] = r ’ for r in the interval
∣ b ⋅ F[A, S] = m/n
Pα ⎢⎢ ⎥

between .99 and 1. This collection of hypotheses includes the claim that
⎢ ⋅ Rnd[S, B, A] ⎥
⎣ ⎦
“all ravens are black” together with those alternative hypotheses that claim
⋅ Size[S] = n the frequency of being black among ravens is within .01 of 1. The
for a given upper bound K (on how much more initially plausible it is that alternatives to these hypotheses are just those that assert ‘F[A, B] = s ’ for
the true population frequency lies outside the region between v and u than values of s below .99.
it is that the true population frequency lies inside that region), you may be
Suppose none of the support functions represented in the vagueness or
able to simply paste the following expression into your program and then
diversity set under consideration rates the prior plausibility of any of the
hypotheses ‘F[A, B] = s’ with s less than .99 to be more than twice as
plug in desired values for K, u, v, m, n in this expression:

1 plausible as the hypotheses ‘F[A, B] = r ’ for which r is between .99 and


1 + K ∗ ( BETADIST(u,m+1,n−m+1)−BETADIST(v,m+1,n−m+1)
1
− 1) 1. That is, suppose, for each Pα in the vagueness or diversity set under
consideration, the prior plausibility Pα [F[A, B] = s ∣ b] for hypotheses
In many real cases it will not be initially more plausible that the true with s below .99 is never more than K = 2 times greater than the prior
frequency value lies outside of the region of interest between v and u than plausibility Pα [F[A, B] = r ∣ b] for hypotheses with r between .99 and 1.
that it lies inside that region. In such cases set the value of K to 1. Then, on the evidence of 400 ravens selected randomly with respect to
However, you will find that for any moderately large sample size n, this color, the theorem yields the following bound for all Pα in the vagueness
function yields very similar values for all plausible values of K you might or diversity set:

⎡ F[A, B] > .99 ⎤


try out, even when the values of K are quite large. (We’ll see examples of
this fact in the computed tables below.) ⎢ ⎥
∣ b ⋅ F[A, S] = 1
Pα ⎢⎢ ⎥ ≥ .9651.

This theorem implies that for large samples the values of prior ⎢ ⋅ Rnd[S, B, A] ⎥
probabilities don’t matter much. Given such evidence, a vary wide range ⎣ ⋅ Size[S] = 400 ⎦
of inductive support functions Pα will come to agree on high posterior
The following table describes similar results for other upper bounds K on
probabilities that the proportion of attribute A in population B is very close
values of prior probability ratios and other sample sizes n:
to the sample frequency. Thus, all support functions in such vagueness or
diversity sets come to near agreement. Let us look at several numerical
examples to make clear how strong this result really is.

108 Stanford Encyclopedia of Philosophy Spring 2021 Edition 109


Inductive Logic James Hawthorne

TABLE 1: Values of lower bound p on the posterior probability random sample of 1600 black ravens will, nevertheless, pull the posterior
m/n = 1 Sample-Size = n plausibility level that “the true frequency is above .99” to a value above
F[A, B] > .99 (number of As in Sample of Bs = m = n) .9898, for every support function in the set. And if the vagueness or
Prior Ratio K 400 800 1600 3200

diversity set contains support functions that assign even more extreme
1 0.9822 0.9997 1.0000 1.0000
priors, say, priors that are nearly ten million times higher for some
2 0.9651 0.9994 1.0000 1.0000 hypotheses asserting frequencies below .99 than for hypotheses within .99
5 0.9170 0.9984 1.0000 1.0000 of 1 (the table’s last row), this poses no great problem for convergence-to-
10 0.8468 0.9968 1.0000 1.0000 agreement. A random sample of 3200 black ravens will yield posterior
100 0.3560 0.9691 1.0000 1.0000
probabilities (i.e., posterior plausibilities) indistinguishable from 1 for the
1,000 0.0524 0.7581 0.9999 1.0000
10,000 0.0055 0.2386 0.9990 1.0000 claim that “more than 99% of all ravens are black”.
100,000 0.0006 0.0304 0.9898 1.0000
1,000,000 0.0001 0.0031 0.9068 1.0000 Strong support can be gotten for a narrower range of hypotheses about the
10,000,000 0.0000 0.0003 0.4931 1.0000 percentage of black birds among the ravens, provided that the sample size
⎡ F[A, B] > .99 ⎤
is increased enough. We’ll return to this in a later subsection.
⎢ ⎥
∣ b ⋅ F[A, S] = 1
Pα ⎢⎢ ⎥ ≥ p,

⋅ Rnd[S, B, A]
Now consider the second example of an enumerative induction provided at
⎢ ⎥
⎣ ⋅ Size[S] = n ⎦ the beginning of this article, involving the poll about the presidential
preferences of voters. The posterior probabilities for this example follow a
for a range of Sample-Sizes n (from 400 to 3200), when the prior
probability of any specific frequency hypothesis outside the region pattern similar to that of the first example. Let ‘B’ represent the class of all
between .99 and 1 is no more than K times more than the lowest prior registered voters on February 20, 2004, and let ‘A’ represent those who
probability for any specific frequency hypothesis inside of the region
prefer Kerry to Bush. In sample S (randomly drawn from B with respect to
between .99 and 1.
A) consisting of 400 voters, 248 report preference for Kerry over Bush—
(All probabilities with entries ‘1.0000’ in this table and the next actually i.e., F[A, B] = 248/400 = .62 . Suppose, as seems reasonable, that none of
have values slightly less than one, but nearly equal 1.0000 to four the support functions in the vagueness or diversity set under consideration
significant decimal places.) rates the hypotheses ‘F[A, B] = r’ for values of r outside the interval
.62 ± .05 as more initially plausible than they rate alternative frequency
To see what the table tells us, consider the third to last row. It represents hypotheses having values of r inside this interval. That is, suppose, for
what happens when a vagueness or diversity set contains at least some each Pα under consideration, the prior probabilities Pα [F[A, B] = s ∣ b]
support functions that assign prior probabilities (i.e., prior plausibilities) when s is not within .62 ± .05 is never more than K = 1 times as great as
nearly one hundred thousand times higher to some hypotheses asserting the prior probabilities Pα [F[A, B] = r ∣ b] for hypotheses having r within
frequencies not between .99 and 1 than it assigns to hypotheses asserting .62 ± .05. Then, the theorem yields the following lower bound on the
frequencies between .99 and 1. The table shows that even in such cases, a

110 Stanford Encyclopedia of Philosophy Spring 2021 Edition 111


Inductive Logic James Hawthorne

posterior plausibility ratings, for all Pα in the vagueness or diversity set TABLE 2: Values of lower bound p on the posterior probability
under consideration: m/n = .62 F[A, B] = Sample-Size = n
m/n ± q (number of As in Sample of Bs = m:
⎡ .57 < F[A, B] < .67 ⎤ where m/n = .62 )
⎢ ⎥ Prior Ratio K q = .05
∣ b ⋅ F[A, S] = .62
Pα ⎢⎢ ⎥ ≥ .9614.
400 800 1600 3200 6400 12800


or .025 (248) (496) (992) (1984) (3968) (7936)
⎢ ⋅ Rnd[S, B, A] ⎥ .05 →
⎣ ⎦
1 0.9614 0.9965 1.0000 1.0000 1.0000 1.0000
⋅ Size[S] = 400 .025 → 0.6982 0.8554 0.9608 0.9964 1.0000 1.0000
2 .05 → 0.9256 0.9930 0.9999 1.0000 1.0000 1.0000
The following table gives similar results for other sample sizes, and for .025 → 0.5364 0.7474 0.9246 0.9929 0.9999 1.0000
5 .05 → 0.8327 0.9827 0.9998 1.0000 1.0000 1.0000
.025 →
upper bounds on ratios of prior probabilities that may be much larger than 0.3163 0.5420 0.8306 0.9825 0.9998 1.0000
1. In addition, this table shows what happens when we tighten up the 10 .05 → 0.7133 0.9661 0.9996 1.0000 1.0000 1.0000
interval around the frequency hypotheses being supported to .62 ± .025— .025 → 0.1879 0.3717 0.7103 0.9656 0.9996 1.0000
100 .05 → 0.1992 0.7402 0.9963 1.0000 1.0000 1.0000
i.e., it shows the bounds p on support for the hypothesis .025 → 0.0226 0.0559 0.1969 0.7371 0.9962 1.0000
.595 < F[A, B] < .645 as well: 1,000 .05 → 0.0243 0.2217 0.9639 1.0000 1.0000 1.0000
.025 → 0.0023 0.0059 0.0239 0.2190 0.9637 1.0000
10,000 .05 → 0.0025 0.0277 0.7277 0.9999 1.0000 1.0000
.025 → 0.0002 0.0006 0.0024 0.0273 0.7261 0.9999
100,000 .05 → 0.0002 0.0028 0.2109 0.9994 1.0000 1.0000
.025 → 0.0000 0.0001 0.0002 0.0028 0.2096 0.9994
1,000,000 .05 → 0.0000 0.0003 0.0260 0.9940 1.0000 1.0000
.025 → 0.0000 0.0000 0.0000 0.0003 0.0258 0.9943
10,000,000 .05 → 0.0000 0.0000 0.0027 0.9433 1.0000 1.0000
.025 → 0.0000 0.0000 0.0000 0.0000 0.0026 0.9457

⎡ .62 − q < F[A, B] < .62 + q ⎤


Pα ⎢⎢ ∣ F[A, S] = .62 ⋅ Rnd[S, B, A] ⎥ ≥ p,

⎣ ⋅ Size[S] = n ⎦

for two values of q (.05 and .025) and a range of Sample-Sizes n (from
400 to 12800), when the prior probability of any specific frequency
hypothesis outside of .62 ± q is no more than K times more than the
lowest prior probability for any specific frequency hypothesis inside of
.62 ± q.

Notice that even if the vagueness or diversity set includes prior


plausibilities nearly ten million times higher for hypotheses asserting
frequency values outside of .62 ± .025 than for hypotheses asserting
frequencies within .62 ± .025 , a random sample of 12800 registered

112 Stanford Encyclopedia of Philosophy Spring 2021 Edition 113


Inductive Logic James Hawthorne

voters will, nevertheless, bring about a posterior plausibility value greater Theorem: Weak Law of Large Numbers for Enumerative
than .9457 for the claim that “the true frequency of preference for Kerry Inductions.
over Bush among all registered voters is within .62 ± .025 ”, for all Let r be any frequency between 0 and 1.
support functions Pα in the set.
For r = 0 ,
2. Convergence to the Truth ⎡ F[A, S] = 0 ⎤
P ⎢⎢ ∣ F[A, B] = 0 ⋅ Rnd[S, B, A] ⎥ = 1.

⎣ ⋅ Size[S] = n ⎦
The Frequency Estimation Theorem is a Bayesian Convergence-to-
Agreement result. It does not, on its own, show that the Criterion of
Adequacy (CoA) is satisfied. The theorem shows, for enumerative For r = 1 ,

⎡ F[A, S] = 1 ⎤
inductions, that as evidence accumulates, diverse support functions will

P ⎢⎢ ∣ F[A, B] = 1 ⋅ Rnd[S, B, A] ⎥ = 1.
come to near agreement on high posterior support strengths for those

⎣ ⎦
hypotheses expressing population frequencies near the sample frequency.
But, it does not show that the true hypothesis is among them—it does not ⋅ Size[S] = n
show that the sample frequency is near the true population frequency. So, For 0 < r < 1, let q be any real number such that r is in the
it does not show that these converging support functions converge on region,
strong support for the true hypothesis, as a CoA result is supposed to do.
0 < (r − q) < r < (r + q) < 1.
However, there is such a CoA result close at hand. It is a Weak Law of
Large Numbers result that establishes that each frequency hypothesis of Given a specific q (which identifies a specific small region of interest
form ‘F[A, B] = r’ implies, via direct inference likelihoods, that randomly around r), for each given positive integer n that’s large enough to
selected sample data is highly likely to result in sample frequencies very permit it, we define associated non-negative integers v and u such that
close to the value r that it claims to be the true frequency. Of course each v < u , where by definition:
frequency hypothesis says that the sample frequency will be near its own
frequency value; but only the true hypothesis says this truthfully. Add this v is the non-negative integer for which v/n is the smallest fraction
result to the previous theorem and we get that, for large sample sizes, it is greater than (r − q) , and
very likely that a sample frequency will occur that yields a very high
u is the non-negative integer for which u/n is the largest fraction
less than (r + q) .
degree of support for the true hypothesis. Thus the CoA is satisfied.

Here is the needed result.


Then,

⎡ ⎤
⎢ ⎥
⎢ Spring 2021 Edition ⎥
⎣ ⎦
114 Stanford Encyclopedia of Philosophy 115
Inductive Logic James Hawthorne

⎡ r − q < F[A, S] < r + q ⎤


P⎢⎢ ∣ F[A, B] = r ⋅ Rnd[S, B, A] ⎥
where competing hypotheses are empirically distinct enough to disagree,
⎥ at least a little, on the likelihoods of possible evidential outcomes.
⎣ ⋅ Size[S] = n ⎦
u
n! 3. Tighter Bounds on the Margin of Error
=∑ × rm (1 − r)n−m
m=v
m! × (n − m)!
⎡ ⎤
If we want strong support for hypotheses claiming more than 99.9% of all
−q
≈ 1 − 2 × Φ ⎢⎢ 1


ravens are black, the following extension of Table 1 applies.
⎣ ((r × (1 − r))/n) 2 ⎦ TABLE 3: Values of lower bound p on the posterior probability
≥ 1 − 2 × Φ [−2 × q × n ] ,
1
2
m/n = 1 Sample-Size = n
F[A, B] > .999 (number of As in Sample of Bs = m = n)
which goes to 1 quickly as n increases. Prior Ratio: K 400 800 1600 3200 6400 12800 25600

Here Φ[x] is the area under the Standard Normal Distribution up to point
1 0.3305 0.5513 0.7985 0.9593 0.9983 1.0000 1.0000
2 0.1980 0.3805 0.6645 0.9219 0.9967 1.0000 1.0000
x. The first equality is a version of the binomial theorem. The 5 0.0899 0.1973 0.4421 0.8252 0.9918 1.0000 1.0000
approximation of the binomial formula by the normal distribution is 10 0.0470 0.1094 0.2838 0.7023 0.9837 1.0000 1.0000
guaranteed by the Central Limit Theorem. This approximation is very 100 0.0049 0.0121 0.0381 0.1909 0.8578 0.9997 1.0000
close for n near 20, and gets extremely close as n gets larger. 1,000 0.0005 0.0012 0.0039 0.0231 0.3763 0.9973 1.0000
10,000 0.0000 0.0001 0.0004 0.0024 0.0569 0.9733 1.0000
100,000 0.0000 0.0000 0.0000 0.0002 0.0060 0.7849 1.0000
Notice that the degree of support probability in this theorem is a direct
1,000,000 0.0000 0.0000 0.0000 0.0000 0.0006 0.2674 1.0000
inference likelihood—all support functions should agree on these values. 10,000,000 0.0000 0.0000 0.0000 0.0000 0.0001 0.0352 0.9999
[23]
⎡ F[A, B] > .999 ⎤
⎢ ⎥
∣ b ⋅ F[A, S] = 1
This Weak Law result together with the Simple Estimation Theorem yields Pα ⎢⎢ ⎥ ≥ p,

the promised CoA result: for large sample sizes, it is very likely that a ⎢ ⋅ Rnd[S, B, A] ⎥
⎣ ⋅ Size[S] = n ⎦
sample frequency will occur that has a value very near the true frequency;
and whenever such a sample frequency does occur, it yields a very high for a range of Sample-Sizes n (from 400 to 25600), when the prior
probability of any specific frequency hypothesis outside the region
degree of support for the true frequency hypothesis. between .999 and 1 is no more than K times more than the lowest prior
probability for any specific frequency hypothesis inside of the region
This result only applies to enumerative inductions. In the next section of between .999 and 1.
the main article we establish a CoA result that applies much more
The lower right corner of the table shows that even when the vagueness or
generally. It applies to the inductive support of hypotheses in any context
diversity sets include support functions with prior plausibilities up to ten

116 Stanford Encyclopedia of Philosophy Spring 2021 Edition 117


Inductive Logic James Hawthorne

million times higher for hypotheses asserting frequency values below .999 described in terms of statements or sentences of a language. So let’s
than for hypotheses making frequency claims between .999 and 1, a follow suit throughout this discussion.
sample of 25600 black ravens will, nevertheless, pull the posterior
plausibility above .9999 that “the true frequency is over .999” for every Plausibility relations (Friedman & Halpern 1995) constitute the most
support function in the set. general of these representations. They satisfy the weakest axioms, the
weakest constraints on the logic of uncertainty. For a plausibility relation
Some Prominent Approaches to the Representation ≽ between sentences, an expression ‘A ≽ B’, says that A is at least as
plausible as B. The axioms for plausibility relations say that tautologies
of Uncertain Inference
are more plausible than contradictions, any two logically equivalent
The following figure indicates a relationship among six of the most sentences are plausibility-related to other sentence in precisely the same
prominent approaches to uncertain inference. The arrows point from more way, a sentence is no more plausible than the sentences it logically entails,
general to less general representation schemes. For example, the and the at least as plausible relation is transitive. These axioms make
Dempster-Shafer representation contains the probability functions as a plausibility relations weak partial orders on the relative plausibility of
special case. sentences. They permits some sentences to be incomparable with respect
to plausibility—neither more plausible, nor less plausible, nor equally
plausible to one another.

Qualitative probability relations are plausibility relations for which three


additional axioms are satisfied. One of these additional axioms says that
when a sentence S is logically incompatible with both sentence A and
sentence B, then A ≽ B holds just in case (A or S) ≽ (B or S) holds as
well. Now, suppose that a qualitative probability relations are on a
language satisfy one additional (two-part) axiom, they can be shown to be
representable by probability functions—i.e., given any qualitative
probability relation ≽, there is a unique probability function P such that
A ≽ B just in case P[A] ≥ P[B]. So quantitative probability may be
viewed as essentially just a way of placing a numerical measure on
REPRESENTATIONS OF UNCERTAINTY sentences that uniquely emulates the is at least as plausible relation
specified by qualitative probability (see Koopman 1940; Savage 1954;
These representations are often described as measures on events, or states,
Hawthorne & Bovens 1999; Hawthorne 2009).
or propositions, or sets of possibilities. But deductive logics are usually

118 Stanford Encyclopedia of Philosophy Spring 2021 Edition 119


Inductive Logic James Hawthorne

Probability (i.e., quantitative probability) is a measure of plausibility that 1978; Dubois & Prade 1980, 1990) are generally read as representing the
assigns a number between 0 and 1 to each sentence. Intuitively, the degree of uncertainty in a claim, where such uncertainty is often attributed
probability of a sentence S, P[S] = r, says that S is plausible to degree r, to vagueness or fuzziness. These functions are formally like probability
or that the rational degree of confidence (or belief) that S is true is r. The functions and Dempster-Shafer functions, but they subscribe to a simpler
axioms for probabilities basically require two things. First, tautologies get addition rule: the degree of uncertainty of a disjunction is the greater of
probability 1. Second, when A and B contradict each other, the probability the degrees of uncertainty of the parts. Similarly, the degree of uncertainty
of the disjunction (A or B) must be the sum of the probabilities of A and of of a conjunction is the smaller of the uncertainties of the parts.
B individually. It is primarily in regard to this second axiom that
probability differs from each of the other quantitative measures of Ranking functions (Spohn 1988) supply a measure of how surprising it
uncertainty. would be if a claim turned out to be true, rated on a scale from 0 (not at all
surprising) to infinity. Tautologies have rank 0 and contradictions are
Like probability, Dempster-Shafer belief functions (Shafer 1976, 1990) infinitely surprising. Logically equivalent claims have the same rank. The
measure appropriate belief strengths on a scale between 0 and 1, with rank of a disjunction is equal to the rank of the lower ranking disjunct.
contradictions and tautologies at the respective extremes. But whereas the These functions may be used to represent a kind of order-of-magnitude
probability of a disjunction of incompatible claims must equal the sum of reasoning about the plausibility of various claims.
the parts, Dempster-Shafer belief functions only require such disjunctions
be believed at least as strongly as the sum of the belief strengths of the See Halpern 2003 for a good comparative treatment of all of these
parts. So these functions are a generalization of probability. By simply approaches.
tightening up the Dempster-Shafer axiom about how disjunctions are
Here are the axioms for the Plausibility Relations and the Qualitative
related to their parts we get back a restricted class of Dempster-Shafer
Probability Relations. We may specify them in terms of a formal language
functions that just is the class of probability functions. Dempster-Shafer
L for either sentential logic or for predicate logic. In this formal language
functions are primarily employed as a logic of evidential support for
we employ the following expressions to represent the standard logical
terms: ‘not’, ‘∼’; ‘and’, ‘⋅’; ‘or’, ‘∨ ’. We represent the the standard
hypotheses. In that realm they are a generalization of the idea of evidential
logical entailment relation with the symbol ‘⊨’—i.e., the expression ‘
support embodied by probabilistic inductive logic. There is some
B ⊨ A’ says “B logically entails A” and the expression ‘⊨ A’ says “A is a
controversy as to whether such a generalization is useful or desirable, or
whether simple probability is too narrow to represent important evidential
tautology”.
relationships captured by some Dempster-Shafer functions.
Axioms for the Plausibility Relations
Each plausibility relation ≽ satisfies the following axioms:
There is a sense in which the other two quantitative measures of
uncertainty, possibility functions and ranking functions, are definable in
terms of formulas employing the Dempster-Shafer functions. But this is 1. if ⊨ A , then not ∼A ≽ A ;
not the best way to understand them. Possibility functions (Zadeh 1965 2. if B ⊨ A, then A ≽ B;

120 Stanford Encyclopedia of Philosophy Spring 2021 Edition 121


Inductive Logic James Hawthorne

3. if A ⊨ B and B ⊨ A (i.e., if A is logically equivalent to B), then: consisting of a disjunction of n sentences, ⊨ (S1 ∨ S2 ∨ … ∨ Sn ),
(1) if A ≽ C, then B ≽ C; and (2) if D ≽ A, then D ≽ B; where for each distinct Si and Sj we have ⊨ ∼(Si ⋅ Sj ) (each distinct
4. if A ≽ B and B ≽ C, then A ≽ C. Si and Sj are inconsistent with one another) and where Si ≈ Sj (they
are equally plausible), and such that for each of the Si , A ≻ (B ∨ Si ) .
Axioms for the Qualitative Probability Relations
To get the qualitative probability relations we add the axioms Think of each sentence Si as saying that in some fair lottery consisting of
n tickets ticket i will win (where exactly one ticket must win, and where
5. if ⊨ ∼(A ⋅ C) and ⊨ ∼(B ⋅ C), then A ≽ B if and only if each ticket has the same chance of winning).
(A ∨ C) ≽ (B ∨ C).
6. A ≽ B or B ≽ A; It can be proved that for each fine-grained qualitative probability relation
≽ there is a unique probability function P such that A ≽ B just in case
The typical axioms for quantitative probability are as follows: P[A] ≥ P[B].
i. for all sentences A, 0 ≤ P[A] ≤ 1; Now, call a qualitative probability relation ≽ properly extendable just in
ii. if ⊨ A , then P[S] = 1; case it can be extended to a fine-grained qualitative probability relation
iii. if ⊨ ∼(A ⋅ B) , then P[(A ∨ B)] = P[A] + P[B]. (perhaps defined on a larger language, i.e., a language containing
additional sentences). Then for every properly extendable qualitative
Axioms 1–6 for the qualitative probability relations are probabilistically
probability relation ≽ there is a probability function P such that A ≽ B
sound with respect to the quantitative probability functions. That is, for
just in case P[A] ≥ P[B]. In general a given properly extendable
each given probability function P, define a relation ≽ such that A ≽ B just
in case P[A] ≥ P[B]. Then it can be shown that ≽ must satisfy axioms 1–
qualitative probability relation may have many such representing
probability functions, corresponding to different ways of extending it to
6. However, not every qualitative probability relation that satisfies axioms
fine-grained qualitative probability relations.
1–6 may be represented by a probability function. To get that we must add
one further structural axiom. Thus, the quantitative probability functions may be viewed as just useful
ways of representing properly extendable qualitative probability relations
First we need to define a bit more notation. Two sentences are defined as
equally plausible, A ≈ B, just when both A ≽ B and B ≽ A. One sentence
on a convenient numerical scale.

is defined as more plausible than another, A ≻ B, just when A ≽ B but not


B ≽ A.
Likelihood Ratios, Likelihoodism, and the Law of
Likelihood
Let’s say that a qualitative probability relation ≽ is fine-grained just in
case it satisfies the following axiom: The versions of Bayes’ Theorem provided by Equations 9–11 show that
for probabilistic inductive logic the influence of empirical evidence (of the
7. if A ≻ B, then (for each natural number n) there is some tautology

122 Stanford Encyclopedia of Philosophy Spring 2021 Edition 123


Inductive Logic James Hawthorne

ind for which hypotheses express likelihoods) is completely captured by Each probabilistic support function satisfies the axioms in Section 2.
the ratios of likelihoods, According to these axioms the conditional probability of one sentence on
another is always defined. So, in the context of the inductive logic of
P[en ∣ hj ⋅ b ⋅ cn ]
. support functions the likelihoods are always defined, and the qualifying
P[en ∣ hi ⋅ b ⋅ cn ] clause about this in the General Law of Likelihood is automatically
The evidence (cn ⋅ en ) influences the posterior probabilities in no other satisfied. For inductive support functions, all of the versions of Bayes’
way. So, the following “Law” is a consequence of the logic of theorem (Equations 8–11) continue to hold even when the likelihoods are
probabilistic support functions. not objective or intersubjectively agreed on by the scientific community.
In many scientific contexts there will be general agreement on the values
General Law of Likelihood: of likelihoods; but whenever such agreement fails the subscripts α, β , etc.
Given any pair of incompatible hypotheses hi and hj , whenever the must remain attached to the support function likelihoods to indicate this.
likelihoods Pα [en ∣ hj ⋅ b ⋅ cn ] and Pα [en ∣ hi ⋅ b ⋅ cn ] are defined, the Even so, the General Law of Likelihood continues to hold for each
evidence (cn ⋅ en ) supports hi over hj , given b, if and only if support function.

Pα [en ∣ hi ⋅ b ⋅ cn ] > Pα [ en ∣ hj ⋅ b ⋅ cn ]. There is a view, or family of views, called likelihoodism that maintains
that the inductive logician or statistician should only be concerned with
The ratio of likelihoods whether the evidence provides increased or decreased support for one
Pα [en ∣ hi ⋅ b ⋅ cn ] hypothesis over another, and only in cases where this evaluation is based
Pα [en ∣ hj ⋅ b ⋅ cn ] on the ratios of completely objective likelihoods. (Prominent likelihoodists
include Edwards (1972) and Royall (1997); also see Forster & Sober
measures the strength of the evidence for hi over hj given b. (2004) and Fitelson (2007).) When the likelihoods involved are objective,
the ratios
Two features of this law require some explanation. As stated, the General
Law of Likelihood does not presuppose that likelihoods of form P[en ∣ hj ⋅ b ⋅ cn ]
Pα [en ∣ hj ⋅ b ⋅ cn ] and Pα [en ∣ hi ⋅ b ⋅ cn ] are always defined. This P[en ∣ hi ⋅ b ⋅ cn ]
qualification is introduced to accommodate a conception of evidential
provide a pure, objective measure of how strongly the evidence supports
support called Likelihoodism, which is especially influential among
statisticians. Also, the likelihoods in the law are expressed with the
hi as compared to hj , a measure that is “untainted” by prior plausibility
considerations. According to likelihoodists, only this kind of pure measure
subscript α attached to indicate that the law holds for each inductive
is scientifically appropriate for the assessment of how evidence impacts
support function Pα , even when the values of the likelihoods are not
hypotheses. (It should be noted that the classical school of statistics,
objective or agreed on by all agents in a given scientific community. These
associated with R.A. Fisher (1922) and with Neyman and Pearson (1967),
two features of the law are closely related, as we will see.

124 Stanford Encyclopedia of Philosophy Spring 2021 Edition 125


Inductive Logic James Hawthorne

reject both the Bayesian logic of evidential support and the claim about the the hypothesis (together with background) logically entails an explicit
nature of evidential support expressed by the General Law of simple statistical hypothesis that (together with experimental conditions)
Likelihood.) specifies precise statistical probabilities for each of the events that make
up the evidence.
Likelihoodists maintain that it is not appropriate for statisticians to
incorporate assumptions about prior probabilities of hypotheses into the Likelihoodists contrast simple statistical hypotheses with composite
assessment of evidential support. It is not their place to compute statistical hypotheses, which only entail vague, or imprecise, or
recommended values of posterior probabilities for the scientific directional claims about the statistical probabilities of evidential events.
community. When the results of experiments are made public, say in Whereas a simple statistical hypothesis might say, for example, “the
scientific journals, only objective likelihoods should be reported. The chance of heads on tosses of the coin is precisely .65”, a composite
evaluation of the impact of objective likelihoods on agents’ posterior statistical hypothesis might say, “the chance of heads on tosses is either
probabilities depends on each agent’s individual subjective prior .65 or .75”, or it may be a directional hypothesis that says, “the chance of
probability, which represents plausibility considerations that have nothing heads on tosses is greater than .65”. Likelihoodists maintain that composite
to do with the evidence. So, likelihoodists suggest, posterior probabilities hypotheses are not an appropriate basis for well-defined likelihoods. Such
should be left for individuals to compute, if they desire to do so. hypotheses represent a kind of disjunction of simple statistical hypotheses.
The direction hypothesis, for instance, is essentially a disjunction of the
The conditional probabilities between most pairs of sentences fail to be various simple statistical hypotheses that assign specific values above .65
objectively defined in a way that suits likelihoodists. So, for likelihoodists, to the chances of heads on tosses. Likelihoods based on such hypotheses
the general logic of evidential support functions (captured by the axioms are not appropriately objective by the lights of the likelihoodist because
in Section 2 and the forms of Bayes’ theorem discussed above) cannot they must in effect depend on factors that represent the degree to which
represent an objective logic of evidential support for hypotheses. Because the composite hypothesis supports each of the simple statistical hypotheses
they eschew the logic of support functions, likelihoodist do not have that it encompasses; and likelihoodists consider such factors too subjective
Bayes’ theorem available, and so cannot derive the Law of Likelihood to be permitted in a logic that should permit only objective likelihoods.[24]
from it. Rather, they must state the Law of Likelihood as an axiom of
their particular version of inductive logic, an axiom that applies only when Taking all of this into account, the version of the Law of Likelihood
the likelihoods have well-defined objective values. appropriate to likelihoodists may be stated as follows.

Likelihoodists tend to have a very strict conception of what it takes for Special Law of Likelihood:
likelihoods to be well-defined. They consider a likelihood to be well- Given a pair of incompatible hypotheses hi and hj that imply simple
defined only when it is (what we referred to earlier as) a direct inference statistical models regarding outcomes en given (b ⋅ cn ), the likelihoods
likelihood— i.e., only when either, (1) the hypothesis (together with P[en ∣ hj ⋅ b ⋅ cn ] and P[en ∣ hi ⋅ b ⋅ cn ] are well defined. For such
background and experimental conditions) logically entails the data, or (2)

126 Stanford Encyclopedia of Philosophy Spring 2021 Edition 127


Inductive Logic James Hawthorne

likelihoods, the evidence (cn ⋅ en ) supports hi over hj , given b, if and Thus, the General Law of Likelihood implies the following principle.
only if
General Likelihood Principle:
(P[en ∣ hi ⋅ b ⋅ cn ] > P[ en ∣ hj ⋅ b ⋅ cn ]; Suppose two different experiments or observations (or two sequences
of them) c1 and c2 produce outcomes e1 and e2 , respectively. Let
the ratio of likelihoods {h1 , h2 , …} be any set of alternative hypotheses. If there is a constant
P[en ∣ hi ⋅ b ⋅ cn ] K such that for each hypothesis hj from the set,
P[en ∣ hj ⋅ b ⋅ cn ]
Pα [e1 ∣ hj ⋅ b ⋅ c1 ] = K × Pα [e2 ∣ hj ⋅ b ⋅ c2 ],
measures the strength of the evidence for hi over hj given b.
then the evidential import of (c1 ⋅ e1 ) for distinguishing among
Notice that when either version of the Law of Likelihood holds, the hypotheses in the set (given b) is precisely the same as the evidential
absolute size of a likelihood is irrelevant to the strength of the evidence. import of (c2 ⋅ e2 ).
All that matters is the relative size of the likelihoods for one hypothesis as
Similarly, the Special Law of Likelihood implies a corresponding Special
compared to another. That is, let c1 and c2 be the conditions for two
Likelihood Principle that applies only to hypotheses that express simple
distinct experiments having outcomes e1 and e2 , respectively. Suppose
statistical models.[25]
that e1 is 1000 times more likely on hi (given b ⋅ c1 ) than is e2 on hi
(given b ⋅ c2 ); and suppose that e1 is also 1000 times more likely on hj Throughout the remainder of this article we will not assume that
(given b ⋅ c1 ) than is e2 on hj (given b ⋅ c2 )—i.e., suppose that likelihoods must be based on simple statistical hypotheses, as likelihoodist

Pα [e1 ∣ hi ⋅ b ⋅ c1 ] = 1000 × Pα [e2 ∣ hi ⋅ b ⋅ c1 ],


would have them. However, most of what will be said about likelihoods,
especially the convergence result in Section 4, applies to likelihoodist
and likelihoods as well. We will, however, continue to suppose that likelihoods
are objective in the sense that all members of the scientific community
Pα [e1 ∣ hj ⋅ b ⋅ c1 ] = 1000 × Pα [e2 ∣ hj ⋅ b ⋅ c2 ]. agree on their numerical values. In Section 5 we will see how even this
supposition may be relaxed in scientific contexts where completely
Which piece of evidence, (c1 ⋅ e1 ) or (c2 ⋅ e2 ), is stronger evidence with
objective values for likelihoods are not realistically available.
regard to the comparison of hi to hj ? The Law of Likelihood implies both
are equally strong. All that matters evidentially are the ratios of the
Proof of the Probabilistic Refutation Theorem
likelihoods, and they are the same:

Pα [e1 ∣ hi ⋅ b ⋅ c1 ] P [e ∣ h ⋅ b ⋅ c2 ] The proof of Convergence Theorem 2 requires the introduction of one


= α 2 i .
Pα [e1 ∣ hj ⋅ b ⋅ c1 ] Pα [e2 ∣ hj ⋅ b ⋅ c2 ]
more concept, that of the variance in the quality of information for a
sequence of experiments or observations, VQI[cn ∣ hi /hj ∣ b]. The quality

128 Stanford Encyclopedia of Philosophy Spring 2021 Edition 129


Inductive Logic James Hawthorne

of the information QI from a specific outcome sequence en may vary Theorem: The VQI Decomposition Theorem for Independent
somewhat from the expected quality of information for conditions cn . A Evidence on Each Hypothesis:
common statistical measure of how widely individual values tend to vary Suppose both condition independence and result-independence
from an expected value is given by the expected squared distance from the hold. Then
expected value, which is called the variance. n
VQI[ cn ∣ hi /hj ∣ b] = ∑ VQI[ck ∣ hi /hj ∣ b].
Definition: VQI—the Variance in the Quality of Information. k=1
For hj outcome-compatible with hi on ck , define
For the Proof, we employ the following abbreviations:
VQI[ck ∣ hi /hj ∣ b] =
Q[ ek ] = QI[ek ∣ hi /hj ∣ b ⋅ ck ]
∑(QI[oku ∣ hi /hj ∣ b ⋅ ck ] − EQI[ck ∣ hi /hj ∣ b])
2

u Q[ ek ] = QI[ek ∣ hi /hj ∣ b ⋅ ck ]
× P[oku ∣ hi ⋅ b ⋅ ck ]. E[ck ] = EQI[ck ∣ hi /hj ∣ b]
E[ck ] = EQI[ck ∣ hi /hj ∣ b]
For a sequence cn of observations on which hj is outcome-compatible
V[ ck ] = VQI[ck ∣ hi / hj ∣ b]
with hi , define
V[ ck ] = VQI[ck ∣ hi / hj ∣ b]
VQI[cn ∣ hi /hj ∣ b] =
The equation stated by the theorem may be derived as follows:
∑ (QI[en ∣ hi /hj ∣ b ⋅ cn ] − EQI[cn ∣ hi /hj ∣ b])2
{e }
n
V[cn ] = ∑(Q[en ] − E[cn ])2 × P[en ∣ hi ⋅ b ⋅ cn ]
× P[en ∣ hi ⋅ b ⋅ cn ]. {e }
n

Clearly VQI will be positive unless hi and hj agree on the likelihoods = ∑((Q[en ] + Q[en−1 ]) − (E[cn ] + E[cn−1 ]))2
{en }
of all possible outcome sequences in the evidence stream, in which
case both EQI[cn ∣ hi /hj ∣ b] and VQI[cn ∣ hi /hj ∣ b] equal 0. × P[en ∣ hi ⋅ b ⋅ cn ]
× P[en−1 ∣ hi ⋅ b ⋅ cn−1 ]
When both Independent Evidence Conditions hold, VQI[cn ∣ hi /hj ∣ b]
decompose into the sum of the VQI for individual experiments or = ∑ ∑((Q[en ] − E[cn ]) + (Q[en−1 ] − E[cn−1 ]))2
{en−1 } {en }
observations ck .
× P[en ∣ hi ⋅ b ⋅ cn ] × P[en−1 ∣ hi ⋅ b ⋅ cn−1]

( )
=
130 Stanford Encyclopedia of Philosophy Spring 2021 Edition 131
Inductive Logic James Hawthorne

= V[cn ] + V[cn−1 ]
( + 2 × (Q[en ] − E[cn ]) × (Q[en−1 ] − E[cn−1 ]) )
(Q[en ] − E[cn ])2 + (Q[en−1] − E[cn−1 ])2
= ∑ ∑
⎛ ⎞
⎜ ∑ ∑ Q[en ] × Q[e ]
n−1
{en−1 } {en } ⎟
⎜ {e } n ⎟
n−1 {e }
× P[en ∣ hi ⋅ b ⋅ cn ] × P[en−1 ∣ hi ⋅ b ⋅ cn−1]
⎜ × P[en ∣ hi ⋅ b ⋅ cn ] ⎟
⎜ n−1 ⎟
= ∑ ∑(Q[en ] − E[cn ])2 × P[en ∣ hi ⋅ b ⋅ cn ] ⎜ × P[en−1
∣ hi ⋅ b × c ] ⎟
⎜− ⎟
∑ Q[en ] × E[c ]
{en−1 } {en }
⎜ {∑
n−1

× P[en−1 ∣ hi ⋅ b ⋅ cn−1 ] n−1 {e } ⎟


⎜ e } n ⎟
+ ∑ ∑(Q[en−1] − E[cn−1])2 × P[en ∣ hi ⋅ b ⋅ cn ] ⎜ × P[en ∣ hi ⋅ b ⋅ cn ] ⎟
{en−1 } {en } ⎜ × P[en−1 n−1 ⎟
∣ hi ⋅ b × c ]
+2×⎜ ⎟
× P[en−1 ∣ hi ⋅ b ⋅ cn−1 ] ⎜− E[cn ] × Q[e ] ⎟
∑ ∑
n−1

+ ∑ ∑ 2 × (Q[en ] − E[cn ]) ⋅ (Q[en−1] − E[cn−1 ]) ⎜ {en−1 } {en } ⎟


⎜ ⎟
{en−1 } {en } ⎜ × P[en ∣ hi ⋅ b ⋅ cn ] ⎟
× P[en ∣ hi ⋅ b ⋅ cn ] × P[en−1 ∣ hi ⋅ b ⋅ cn−1] ⎜ × P[en−1 ∣ hi ⋅ b × cn−1] ⎟
⎜ ⎟
⎜ + ∑ ∑ E[cn ] × E[c ] ⎟
n−1
= V[cn ] + V[cn−1 ] ⎜ {en−1 } {en } ⎟
⎛ Q[en ] × Q[en−1] ⎞ ⎜ ⎟
⎜ ⎟ ⎜ × P[en ∣ hi ⋅ b ⋅ cn ] ⎟
+ 2 × ∑ ∑⎜ ⎜ −Q[en ] × E[cn−1] ⎟ ⎝ × P[en−1 ∣ hi ⋅ b × cn−1] ⎠
{en−1 } {en } ⎜
−E[cn ] × Q[en−1] ⎟⎟
⎝ +E[cn ] × E[cn−1 ] ⎠ = V[cn ] + V[cn−1 ]
× P[en ∣ hi ⋅ b ⋅ cn ] × P[en−1 ∣ hi ⋅ b × cn−1 ] ⎛ E[cn ] × E[cn−1 ] ⎞
⎜ ⎟
+2×⎜ ⎜ −E[cn ] × E[cn−1 ] ⎟
n−1 ⎟
⎜ −E[cn ] × E[c ] ⎟
⎝ +E[cn ] × E[cn−1 ] ⎠

= V[cn ] + V[cn−1 ]
=…
n
= ∑ VQI[ck ∣ hi /hj ∣ b].
k=1

By averaging the values of VQI[cn ∣ hi /hj ∣ b] over the number of


observations n we obtain a measure of the average variance in the

= V[ n ] + V[ n−1
]
132 Stanford Encyclopedia of Philosophy Spring 2021 Edition 133
Inductive Logic James Hawthorne

P[en ∣ hj ⋅ b ⋅ cn ]
[ { } ]
quality of the information due to cn . We represent this average by
P ∨ en : < ε ∣ hi ⋅ b ⋅ cn
overlining ‘VQI’. P[en ∣ hi ⋅ b ⋅ cn ]
⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯
1 VQI[cn ∣ hi /hj ∣ b]
Definition: The Average Variance in the Quality of > 1 − × ⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯
Information n (EQI[cn ∣ hi /hj ∣ b] + (log ε)/n)2

⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯ VQI[cn ∣ hi /hj ∣ b]


VQI[cn ∣ hi /hj ∣ b] = .
Thus, provided that the average expected quality of the information,
⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯
n EQI[cn ∣ hi /hj ∣ b], for the stream of experiments and observationscn
doesn’t get too small (as n increases), and provided that the average
We are now in a position to state a very general version of the second ⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯
part of the Likelihood Ratio Convergence Theorem. It applies to all variance, VQI[cn ∣ hi /hj ∣ b], doesn’t blow up (e.g., it is bounded above),
evidence streams not containing possibly falsifying outcomes for hj . hypothesis hi (together with b ⋅ cn ) says it is highly likely that outcomes of
That is, it applies to all evidence streams for which hj is fully outcome- cn will be such as to make the likelihood ratio against hj as compared to hi
compatible with hi on each ck in the evidence stream. This theorem is as small as you like, as n increases.
essentially a specialized version of Chebyshev’s Theorem, which is a
Proof: Let
Weak Law of Large Numbers.
V = VQI[cn ∣ hi / hj ∣ b]
Likelihood Ratio Convergence Theorem 2*—The Probabilistic E = EQI[cn ∣ hi /hj ∣ b]
Q[ en ] = QI[en ∣ hi /hj ∣ b ⋅ cn ]
Refutation Theorem.
Suppose the evidence stream cn contains only experiments or
P[en ∣ hi ⋅ b ⋅ cn ]
( P[en ∣ hj ⋅ b ⋅ cn ] )
observations on which hj is fully outcome-compatible with hi —i.e., = log
suppose that for each condition ck in sequence cn , for each of its
possible outcomes possible outcomes oku , either Choose any small ε > 0 , and suppose (for n large enough) that
P[oku ∣ hi ⋅ b ⋅ ck ] = 0 or P[oku ∣ hj ⋅ b ⋅ ck ] > 0. And suppose that E > −(log ε)/n . Then we have
the Independent Evidence Conditions hold for evidence stream cn with
respect to each of these hypotheses. Now, choose any positive ε < 1,
as small as you like, but large enough (for the number of observations
⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯ (log ε)
n being contemplated) that the value of EQI[cn ∣ hi /hj ∣ b] > − n
.
Then:

∣ hj ⋅ b ⋅ n]

[ { } ]
= (E − Q 2 × P[ ∣ ⋅b⋅ ]
P[ n

n n
P ∨ n
: <ε ∣ i ⋅b⋅ n i

134 Stanford Encyclopedia of Philosophy Spring 2021 Edition 135


Inductive Logic James Hawthorne

⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯ ⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯
V= ∑ (E − Q)2 × P[en ∣ hi ⋅ b ⋅ cn ] This theorem shows that when VQI is bounded above and EQI has a
{en :P[e ∣ hj ⋅b⋅cn ]>0}
n
positive lower bound, a sufficiently long stream of evidence will very
≥ ∑ (E − Q)2 × P[en ∣ hi ⋅ b ⋅ cn ] likely result in the refutation of false competitors of a true hypothesis. We
⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯
{ en :P[e ∣hj ⋅b⋅cn ]>0
n
can show that VQI will indeed be bounded above when a very simple
& Q[en ]≤−(log ε)}

≥ (E + (log ε))2 × P[en ∣ hi ⋅ b ⋅ cn ]


condition is satisfied. This gives us the version of the theorem stated in the
∑ main text.
{ en :P[e hj ⋅b⋅cn ]>0
n

& Q[en ]≤−(log ε)}


⎡ ⎧ n
e : P[e ∣ hj ⋅ b ⋅ cn ] ⎫ n ⎤ Likelihood Ratio Convergence Theorem 2—The Probabilistic
⎪ ⎪
= (E + (log ε))2 × P ⎢⎢∨ ⎨ > 0 & Q[ e ]n ⎬ ∣ h ⋅ b ⋅ cn ⎥

Refutation Theorem.
⎪ ⎪
i
⎣ ⎩ ≤ log(1/ε) ⎭ ⎦ Suppose the evidence stream cn contains only experiments or observations
n ∣ h ⋅ b ⋅ cn ] on which hj is fully outcome-compatible with hi —i.e., suppose that for
[ { } ]
P[ e
= (E + (log ε))2 × P ∨ en : ≥ ε ∣ hi ⋅ b ⋅ cn
j
each condition ck in sequence cn , for each of its possible outcomes
P[en ∣ hi ⋅ b ⋅ cn ]
possible outcomes oku , either P[oku ∣ hi ⋅ b ⋅ ck ] = 0 or
So, P[oku ∣ hj ⋅ b ⋅ ck ] > 0. In addition (as a slight strengthening of the
⎯⎯⎯⎯
previous supposition), for some γ > 0 a number smaller than 1/e2 (≈ .135;
V V
= where ‘e’ is the base of the natural logarithm), suppose that for each
⎯⎯⎯⎯ log ε 2 (E + (log ε))2
n × (E + n )
possible outcome oku of each observation condition ck in cn , either
P[oku ∣ hi ⋅ b ⋅ ck ] = 0 or P[oku ∣ hj ⋅ b ⋅ ck ]/P[oku ∣ hi ⋅ b ⋅ ck ] ≥ γ. And
P[en ∣ hj ⋅ b ⋅ cn ]
[ { } ]
≥ P ∨ en : ≥ ε ∣ hi ⋅ b ⋅ cn
suppose that the Independent Evidence Conditions hold for evidence
P[en ∣ hi ⋅ b ⋅ cn ] stream cn with respect to each of these hypotheses. Now, choose any
P[en ∣ hj ⋅ b ⋅ cn ] positive ε < 1 , as small as you like, but large enough (for the number of
[ { } ]
= 1 − P ∨ en : < ε ∣ hi ⋅ b ⋅ cn
P[en ∣ hi ⋅ b ⋅ cn ] observations n being contemplated) that the value of
⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯
EQI[cn ∣ hi /hj ∣ b] > −(log ε)/n. Then:
Thus, for any small ε > 0,
P[en ∣ hj ⋅ b ⋅ cn ]
[ { } ]
∣ hj ⋅ b ⋅ cn ] P ∨ en : < ε ∣ hi ⋅ b ⋅ cn
[ { } ]
P[en P[en ∣ hi ⋅ b ⋅ cn ]
P ∨ en : < ε ∣ hi ⋅ b ⋅ cn
P[e ∣ hi ⋅ b ⋅ cn ]
n
1 (log γ)2
⎯⎯⎯⎯ >1− × ⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯
V n (EQI[cn ∣ hi /hj ∣ b] + (log ε)/n)2
≥1−
⎯⎯⎯⎯ (log ε) 2
n × (E + n ) Proof: This follows from Theorem 2* together with the following
observation:
(End of Proof)
⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯ ⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯

136 Stanford Encyclopedia of Philosophy Spring 2021 Edition 137


Inductive Logic James Hawthorne

If for each ck in cn , for each of its possible outcomes oku , either So, for 0 < P[ek ∣ hi ⋅ b ⋅ ck ] < P[ek ∣ hj ⋅ b ⋅ ck ] we have
P[oku ∣ hj ⋅ b ⋅ ck ] = 0 or P[oku ∣ hj ⋅ b ⋅ ck ]/P[oku ∣ hi ⋅ b ⋅ ck ] ≥ γ > 0,
2
P[ek ∣ hi ⋅ b ⋅ ck ]
( ( P[ek ∣ hj ⋅ b ⋅ ck ] ))
for some lower bound γ < 1/e2 (≈ .135 ; where ‘e’ is the base of the log × P[ ek ∣ hi ⋅ b ⋅ ck ]
⎯⎯⎯⎯ ⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯
natural logarithm), then V = VQI[cn ∣ hi /hj ∣ b] ≤ (log γ)2 .
≤ (log(1/ e2 ))2 × P[ek ∣ hj ⋅ b ⋅ ck ]
To see that this observation holds, assume its antecedent.
≤ (log γ)2 × P[ek ∣ hj ⋅ b ⋅ ck ]
1. First notice that when 0 < P[ek ∣ hj ⋅ b ⋅ ck ] < P[ek ∣ hi ⋅ b ⋅ ck ] we
(since, for γ ≤ 1/e2 we have log γ ≤ log(1/e2 ) < 0; so
have
(log γ)2 ≥ (log(1/e2 ))2 > 0).
2
P[ek ∣ hi ⋅ b ⋅ ck ]
( [ P[ek ∣ hj ⋅ b ⋅ ck ] ])
log × P[ ek ∣ hi ⋅ b ⋅ ck ] 3. Now (assuming the antecedent of the theorem), for each ck ,

≤ (log γ)2 × P[ek ∣ hi ⋅ b ⋅ ck ].

So we only need establish that when


P[ek ∣ hj ⋅ b ⋅ ck ] > P[ek ∣ hi ⋅ b ⋅ ck ] > 0 , we will also have this
relationship—i.e., we will also have

2
P[ek ∣ hi ⋅ b ⋅ ck ]
( [ P[ek ∣ hj ⋅ b ⋅ ck ] ])
log × P[ ek ∣ hi ⋅ b ⋅ ck ]

≤ (log γ)2 × P[ek ∣ hi ⋅ b ⋅ ck ].


⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯
(Then it will follow easily that VQI[cn ∣ hi /hj ∣ b] ≤ (log γ)2 , and
we’ll be done.)

2. To establish the needed relationship, suppose that


P[ek ∣ hj ⋅ b ⋅ ck ] > P[ek ∣ hi ⋅ b ⋅ ck ] > 0 . Notice that for all p ≤ q,
p and q between 0 and 1, the function g(p) = (log(p/q))2 × p has a
minimum at p = q , where g(p) = 0, and (for p < q) has a maximum
value at p = q/e2 —i.e., at p/q = 1/e2 . (To get this, take the
derivative of g(p) with respect to p and set it equal to 0; this gives a
maximum for g(p) at p = q/e2 .)

= ∑ (EQI[ k ] − QI[ k ] 2

138 Stanford Encyclopedia of Philosophy Spring 2021 Edition 139


Inductive Logic James Hawthorne

VQI[ck ∣ hi /hj ∣ b] = ∑ (EQI[ck ] − QI[ck ])2 Immediate Consequences of the Independent


{oku :P[ oku ∣ hj ⋅b⋅ck ]>0}
Evidence Conditions
× P[ oku ∣ hi ⋅ b ⋅ ck ]
⎛ EQI[ck ]2 ⎞ When neither independence condition holds, we at least have:
= ⎜ ⎟
∑ ⎜ − 2 × QI[ ck ] × EQI[ck ] ⎟ P[en ∣ hj ⋅ b ⋅ cn ] = P[en ∣ hj ⋅ b ⋅ cn ⋅ en−1 ] × P[en−1 ∣ hj ⋅ b ⋅ cn ]
{oku :P[ oku ∣ hj ⋅b⋅ck ]>0} ⎝ + QI[c ]2 ⎠
k
=…
× P[ oku ∣ hi ⋅ b ⋅ ck ] n
= ∏ P[ ek ∣ hj ⋅ b ⋅ cn ⋅ ek−1 ]
= ∑ EQI[ ck ]2 × P[oku ∣ hi ⋅ b ⋅ ck ]
k=1
{oku :P[ oku ∣ hj ⋅b⋅ck ]>0}

− 2 × EQI[ ck ] When condition-independence holds we have:

( × P[ oku ∣ hi ⋅ b ⋅ ck ] )
QI[ ck ]
× ∑ P[en ∣ hj ⋅ b ⋅ cn ] = P[ en ∣ hj ⋅ b ⋅ cn ⋅ ( cn−1 ⋅ en−1)]
{oku :P[ oku ∣ h ⋅b⋅c ]>0}
× P[ en−1 ∣ hj ⋅ b ⋅ cn ⋅ cn−1]
j k

( × P[ oku ∣ hi ⋅ b ⋅ ck ] )
QI[ck ]2
+ ∑ = P[ en ∣ hj ⋅ b ⋅ cn ⋅ ( cn−1 ⋅ en−1)]
{oku :P[ oku ∣ h ⋅b⋅c ]>0}
j k
× P[ en−1 ∣ hj ⋅ b ⋅ cn−1]
= ∑ QI[ ck ] × P[oku ∣ hi ⋅ b ⋅ ck ]
2
=…
{oku :P[ oku ∣ hj ⋅b⋅ck ]>0} n
− EQI[ck ] 2 = ∏ P[ ek ∣ hj ⋅ b ⋅ ck ⋅ ( ck−1 ⋅ ek−1 )]
≤ QI[ ck ]2 × P[oku ∣ hi ⋅ b ⋅ ck ]
k=1

{oku :P[ oku ∣ hj ⋅b⋅ck ]>0} If we add result-independence to condition-independence, the occurrences
≤ ∑ (log γ)2 × P[oku ∣ hi ⋅ b ⋅ ck ] of ‘(ck−1 ⋅ ek−1 )’ may be removed from the previous formula, giving:
{oku :P[ oku ∣ hj ⋅b⋅ck ]>0}
n
≤ (log γ)2 . P[en ∣ hj ⋅ b ⋅ cn ] = ∏ P[ ek ∣ hj ⋅ b ⋅ ck ]
k=1
So,

⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯
n Proof of the Falsification Theorem
VQI[ck ∣ hi /hj ∣ b] = (1/n) × ∑ VQI[ck ∣ hi /hj ∣ b] ≤ (log γ)2 .
k=1 Likelihood Ratio Convergence Theorem 1—The Falsification
Theorem:
Suppose the evidence stream cn contains precisely m experiments or

140 Stanford Encyclopedia of Philosophy Spring 2021 Edition 141


Inductive Logic James Hawthorne

observations on which hj is not fully outcome-compatible with hi . And Then, we may iteratively decompose
suppose that the Independent Evidence Conditions hold for evidence
stream cn with respect to each of these hypotheses. Furthermore, suppose P [∨ {en : P[en ∣ hj ⋅ b ⋅ cn ] > 0} ∣ hi ⋅ b ⋅ cn ]
there is a lower bound δ > 0 such that for each ck on which hj is not fully
into its components as follows:
outcome-compatible with hi ,
P[∨{en : P[en ∣ hj ⋅ b ⋅ cn ] > 0} ∣ hi ⋅ b ⋅ cn ]
P[∨{oku : P[oku ∣ hj ⋅ b ⋅ ck ] = 0} ∣ hi ⋅ b ⋅ ck ] ≥ δ
= ∑ P[ en ∣ hi ⋅ b ⋅ cn ]
—i.e., hi (together with b ⋅ ck ) says, via a likelihood with value no smaller {en :P[en ∣ hj ⋅b⋅cn ]>0}

than δ , that one of the outcomes will occur that hj says cannot occur). = ∑ P[ en ∣ hj ⋅ b ⋅ cn ⋅ cn−1 ⋅ en−1]
en :P[en ∣hj ⋅b⋅cn ⋅ cn−1 ⋅en−1 ]
{ ×P[en−1 ∣h ⋅b⋅c ⋅cn−1 ]>0 }
Then,
i n

P[en ∣ hj ⋅ b ⋅ cn ]
[ { } ]
P ∨ en : = 0 ∣ hi ⋅ b ⋅ cn × P[ en−1 ∣ hi ⋅ b ⋅ cn ⋅ cn−1]
P[en ∣ hi ⋅ b ⋅ cn ]
= ∑ P[en ∣ hj ⋅ b ⋅ cn ] × P[en−1 ∣ hi ⋅ b ⋅ cn−1]
= P ∨ {en : P[en ∣ hj ⋅ b ⋅ cn ] = 0} ∣ hi ⋅ b ⋅ cn ]
en :P[ en ∣ hj ⋅b⋅ cn ]
{ ×P[en−1 ∣h ⋅b⋅cn−1 ]>0 }
≥ 1 − (1 − δ) , m
i

= ∑ P[en ∣ hj ⋅ b ⋅ cn ] × P[en−1 ∣ hi ⋅ b ⋅ cn−1]


which approaches 1 for large m.
en :P[en ∣ hj ⋅b⋅cn ]>0
{ &P[en−1 ∣h ⋅b⋅cn−1 ]>0 }
i

= P[ onu ∣ hi ⋅ b ⋅ cn ]
Proof
∑ ∑
{en−1 :P[en−1 ∣hj ⋅b⋅cn−1 ]>0} {onu ∈ On :P[onu ∣hj ⋅b⋅cn ]>0}
First notice that according to the supposition of the theorem, for each of
the m experiments or observations ck on which hj is not fully outcome- × P[ en−1 ∣ hi ⋅ b ⋅ cn−1]
compatible with hi we have
= ∑ P[∨{onu : P[onu ∣ hj ⋅ b ⋅ cn ] > 0} ∣ hi ⋅ b ⋅ cn ]
(1 − δ) ≥ P ∨ {oku : P[oku ∣ hj ⋅ b ⋅ ck ] > 0} ∣ hi ⋅ b ⋅ ck ] {en−1 :P[en−1 ∣ hj ⋅b⋅cn−1 ]>0}

= P[oku ∣ hi ⋅ b ⋅ ck ]. × P[ en−1 ∣ hi ⋅ b ⋅ cn−1]



{oku ∈ Ok :P[oku ∣hj ⋅b⋅ck ]>0}
≤ (1 − γ) × ∑ P[en−1 ∣ hi ⋅ b ⋅ cn−1 ],
And for each of the other ck in the evidence stream cn —i.e., for each of {en−1 :P[en−1 ∣ hj ⋅b⋅cn−1 ]>0}
the ck on which hj is fully outcome-compatible with hi ,
if cn is an observation on which hj is not fully outcome-compatible with hi
P [∨ {oku : P[oku ∣ hj ⋅ b ⋅ ck ] > 0} ∣ hi ⋅ b ⋅ ck ] = 1.

142 Stanford Encyclopedia of Philosophy Spring 2021 Edition 143


Inductive Logic James Hawthorne

P[en ∣ hj ⋅ b ⋅ cn ]
[ { } ]
P ∨ en : > 0 ∣ hi ⋅ b ⋅ cn
or, alternatively,
P[en ∣ hi ⋅ b ⋅ cn ]
= ∑ P[en−1 ∣ hi ⋅ b ⋅ cn−1],
{e
n−1
:P[e n−1
∣ hj ⋅b⋅c
n−1
]>0} = ∑ P[ en ∣ hi ⋅ b ⋅ cn ]
{en :P[ en ∣ hj ⋅b⋅ c ]/P[en ∣hi ⋅b⋅cn ]>0}
n

if cn is an observation on which hj is fully outcome-compatible with hi . = ∑ P[ en ∣ hi ⋅ b ⋅ cn ]


{ e :P[e ∣hj ⋅b⋅c ]>0&P[ e ∣ hi ⋅b⋅ c ]>0}
n n n n n

Now, continuing this process of decomposing terms of form


= ∑ P[ en ∣ hi ⋅ b ⋅ cn ]
{en :P[e ∣ hj ⋅b⋅cn ]>0}
P[ek ∣ hi ⋅ b ⋅ ck ]
n

∑ = P ∨ { e : P[e ∣ hj ⋅ b ⋅ cn ] > 0} ∣ hi ⋅ b ⋅ cn ].
n n
{ek :P[ek ∣ hj ⋅b⋅ck ]>0}

(in each disjunct of the ‘or’ above, using the same decomposition process Proof that the EQI for cn is the sum of EQI for the
shown in the six lines preceding that disjunction), and realizing that individual ck
according to the supposition of the theorem, this decomposition leads to
terms of the form of the first disjunct exactly m times, we get Theorem: The EQI Decomposition Theorem:

⋮ When the Independent Evidence Conditions are satisfied,


m
≤ ∏(1 − γ) = (1 − γ)m . n

k=1 EQI[cn ∣ hi / hj ∣ b] = ∑ EQI[ck ∣ hi /hj ∣ b].


k=1
So,
Proof:
P [∨{en : P[en ∣ hj ⋅ b ⋅ cn ] = 0} ∣ hi ⋅ b ⋅ cn ]
EQI[cn ∣ hi /hj ∣ b]
= 1 − P[∨{ en : P[en ∣ hj ⋅ b ⋅ cn ] > 0} ∣ hi ⋅ b ⋅ cn ]
= ∑ QI[en ∣ hi /hj ∣ b ⋅ cn ] × P[ en ∣ hi ⋅ b ⋅ cn ]
≥ 1 − (1 − γ)m . n
{e }
P[en ∣ hi ⋅ b ⋅ cn ]
[ P[en ∣ hj ⋅ b ⋅ cn ] ]
= ∑ log × P[en ∣ hi ⋅ b ⋅ cn ]
We also have,
{en }
P[en ∣ hj ⋅ b ⋅ cn ]
[ { } ]
P ∨ e : n
= 0 ∣ hi ⋅ b ⋅ cn
P[en ∣ hi ⋅ b ⋅ cn ]
P [∨ {en : P[en ∣ hj ⋅ b ⋅ cn ] = 0} ∣ hi ⋅ b ⋅ cn ] ,

because
⎛ P[ n ∣ i ⋅ b ⋅ n ⋅ ( ⋅ n−1)] ⎞
[ { } ] ⎜ log[ ] ⎟⎟
n−1
P ∨ n
: >0 ∣ ⋅b⋅ n

i
144 Stanford Encyclopedia of Philosophy = ⎜
Spring 2021 Edition

145
Inductive Logic James Hawthorne

⎛ P[en ∣ hi ⋅ b ⋅ cn ⋅ (cn−1 ⋅ en−1)] ⎞


⎜ log[
P[en ∣ hj ⋅ b ⋅ cn ⋅ (cn−1 ⋅ en−1) ] ⎟

The Effect on EQI of Partitioning the Outcome
= ∑ ∑⎜ ⎜ Space More Finely—Including Proof of the
P[en−1 ∣ hi ⋅ b ⋅ cn ⋅ cn−1] ⎟
[ P[en−1 ∣ hj ⋅ b ⋅ cn ⋅ cn−1] ] ⎠
{en−1 } {en } ⎜ + log ⎟ Nonnegativity of EQI

× P[ en ∣ hi ⋅ b ⋅ cn ⋅ ( cn−1 ⋅ en−1)] Given some experiment or observation (or series of them) c, is there any
× P[ en−1 ∣ hi ⋅ b ⋅ cn ⋅ cn−1] special advantage to parsing the space of possible outcomes O into more
⎛ P[en ∣ hi ⋅ b ⋅ cn ] ⎞
⎜ log[
P[en ∣ hj ⋅ b ⋅ cn ] ]
alternatives rather than fewer alternatives? Couldn’t we do as well at

= ∑ ∑ ⎜⎜ ⎟
evidentially evaluating hypotheses by parsing the space of outcomes into
n−1 ∣ h ⋅ b ⋅ cn−1] ⎟ just a few alternatives—e.g., one possible outcome that hi says is very
[ P[en−1 ∣ hj ⋅ b ⋅ cn−1] ] ⎠
{en−1 } {en } ⎜ + log ⎟
P[ e i
likely and hj says is rather unlikely, one that hi says is rather unlikely and

hj says is very likely, and perhaps a third outcome on which hi and hj
× P[ en ∣ hi ⋅ b ⋅ cn ]
pretty much agree? The answer is “No!”. Parsing the space of outcomes
× P[ en−1 ∣ hi ⋅ b ⋅ cn−1] into a larger number of empirically distinct possible outcomes always
⎛ P[en ∣ hi ⋅ b ⋅ cn ] ⎞
⎜ ∑ log[
P[en ∣ hj ⋅ b ⋅ cn ] ] ⎟

provides a better measure of evidential support.
⎜ {en }
= ⎜⎜ × P[ en ∣ hi ⋅ b ⋅ cn ] ⎟ To see this intuitively, suppose some outcome description o can be parsed

⎜ n−1 ⎟
into two distinct outcome descriptions, o1 and o2 , where o is equivalent to
⎜ × ∑ P[ e ∣ hi ⋅ b ⋅ c ] ⎟ (o1 ∨ o2 ) , and suppose that hi differs from hj much more on the likelihood
n−1

⎝ {en−1 } ⎠ of o1 than on the likelihood of o2 . Then, intuitively, when o is found to be


⎛ P[e ∣ hi ⋅ b ⋅ cn−1] ⎞
⎜ {en−1 } [ P[en−1 ∣ hj ⋅ b ⋅ cn−1] ] ⎟
n−1
⎜ ∑ log ⎟ true, whichever of the more precise descriptions, o1 or o2 , is true should

+ ⎜⎜ × P[ en−1 ∣ h ⋅ b ⋅ cn−1] ⎟
make a difference as to how strong the comparative support for the two
⎟ hypotheses turns out to be. Reporting whichever of o1 or o2 occurs will be
⎜ ⎟
i

⎜ × ∑ P[ en ∣ hi ⋅ b ⋅ cn ] ⎟
more informative than simply reporting o. That is, if the outcome of the
⎝ {en } ⎠ experiment is only described as o, relevant information is lost.

It turns out that EQI measures how well possible outcomes can distinguish
= EQI[cn ∣ hi /hj ∣ b] + EQI[cn−1 ∣ hi /hj ∣ b] between hypotheses in a way that reflects the intuition that a finer partition
= … (iterating this decomposition process) of the possible outcomes is more informative. The numerical value of EQI
n
= ∑ EQI[ck ∣ hi /hj ∣ b]. is always made larger by parsing the outcome space more finely, provided
k=1 that the likelihoods for outcomes in the finer parsing differ at least a bit
from some of the likelihoods for outcomes of the less refined parsing. This

146 Stanford Encyclopedia of Philosophy Spring 2021 Edition 147


Inductive Logic James Hawthorne

is important for our main convergence result because in that theorem we


want the average value of EQI for the whole sequence of experiments and To prove this theorem first notice that
observations to be positive, and the larger the better. r1 (r + r2 )
= 1 iff r1 s1 + r1 s2 = s1 r1 + s1 r2
s1 (s1 + s2 )
The following Partition Theorem implies the Nonnegativity of EQI r r
result as well. It show that each EQI[ck ∣ hi /hj ∣ b] must be non-negative; iff 1 = 2 .
s1 s2
and it will be positive just in case for at least one possible outcome
We’ll draw on this little result immediately below. It is clearly relevant to
oku , P[oku ∣ hj ⋅ b ⋅ ck ] ≠ P[ oku ∣ hi ⋅ b ⋅ ck ]. the antecedent of case (2) of the theorem we want to prove.

This theorem will also show that EQI[ck ∣ hi /hj ∣ b] generally becomes We establish case (2) first. Suppose the antecedent of case (2) holds. Then,
larger whenever the outcome space is partitioned more finely. It follows from the little result just proved, we have
immediately that the average value of EQI for a sequence of experiments
⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯ (r + r2 ) (r + r2 )
r1 log[ ] + r2 log[ 2 ] = r1 log[ 1 ] + r2 log[ 1
(s1 + s2 ) ]
r1 r
or observations, EQI[cn ∣ hi /hj ∣ b], averaged over the sequence of
observations cn, is non-negative, and must be positive if for even one of
s1 s2 (s1 + s2 )
(r + r2 )
= (r1 + r2 ) log[ 1
(s1 + s2 ) ]
the ck that contribute to it, at least one possible outcome oku distinguishes .
between the two hypotheses by making
That establishes case (2).
P[oku ∣ hj ⋅ b ⋅ ck ] ≠ P[oku ∣ hi ⋅ b ⋅ ck ].
To get case (1), consider the following function of p:
Partition Theorem:
For any positive real numbers r1 , r2 , s1 , s2 : (1 − p)
f (p) = p log[ ] + (1 − p) log[
v ]
p
,
u
(1) if r1 /s1 > (r1 + r2 )/(s1 + s2 ), then
where we only assume that u > 0, v > 0, and 0 < p < 1 .
(r + r2 )
(r1 + r2 ) × log[ 1 ] < r1 × log[ 1 ] + r2 × log[ 2 ];
r r
(s1 + s2 ) s1 s2 This function has its minimum value when p = u/(u + v) . (This is easily
verified by setting the derivative of f (p) with respect to p equal to 0 to find
the minimum value of f (p) ; and it is easy to verified that this is a
and
minimum rather than a maximum value.) At this minimum, where
(2) if r1 /s1 = (r1 + r2 )/(s1 + s2 ), then p = u/(u + v), we have

(r + r2 )
r1 × log[ ] + r2 × log[ 2 ] = ( r1 + r2 ) × log[ 1
(s1 + s2 ) ]
r1 r
.
s1 s2
u v
=− log[u + v] − log[u + v]
148 Stanford Encyclopedia of Philosophy Spring 2021 Edition 149
Inductive Logic James Hawthorne

u v
f (p) = − log[u + v] − log[u + v] if
(u + v) (u + v)
r1 (r + r2 )
= − log[u + v]. ≠ 1 ,
s1 (s1 + s2 )
Thus, for all values of p other than u/(u + v),
then
− log[u + v] < f (p)
(r1 + r2 )
(r1 + r2 ) log[ ] < r1 log[ 1 ] + r2 log[ 2 ].
r r
(1 − p)
= p log[ ] + (1 − p) log[
v ]
p
. (s1 + s2 ) s1 s2
u
That is, if p ≠ u/(u + v) ,
This completes the proof of the theorem.

(1 − p) To apply this result to EQI[ck ∣ hi /hj ∣ b] recall that


− log[u + v] < p log[ ] + (1 − p) log[
v ]
p
.
u
P[oku ∣ hi ⋅ b ⋅ ck ]
[ P[oku ∣ hj ⋅ b ⋅ ck ] ]
EQI[ck ∣ hi /hj ∣ b] = ∑ log
Now, let p = r1 /(r1 + r2 ), let u = s1 /(r1 + r2 ), and let v = s2 /(r1 + r2 ). {u:P[oku ∣hj ⋅b⋅ ck ]>0}

× P[oku ∣ hi ⋅ b ⋅ ck ].
Plugging into the previous formula, and multiplying both sides by
(r1 + r2 ), we get:
Suppose ck has m alternative outcomes oku on which both
if
r1 s1 P[oku ∣ hj ⋅ b ⋅ ck ] > 0 and P[ oku ∣ hi ⋅ b ⋅ ck ] > 0.

(r1 + r2 ) (s1 + s2 ) Let’s label their likelihoods relative to hi (i.e., their likelihoods
(i.e., equivalently, if r1 /s1 ≠ (r1 + r2 )/(s1 + s2 )), P[oku ∣ hi ⋅ b ⋅ ck ]) as r1 , r2 , …, rm . And let’s label their likelihoods
relative to hj as s1 , s2 ,… ,sm . In terms of this notation,
then
m
EQI[ck ∣ hi /hj ∣ b] = ∑ ru × log[ u ].
r
(r1 + r2 )
log[
(s1 + s2 ) ]
u=1
su

<[ ] log[ 1 ] + (1 − [ ]) log[ 2 ]


r1 r r1 r Notice also that
(r1 + r2 ) s1 (r1 + r2 ) s2
(r1 + r2 + r3 + … + rm ) = 1
(r +r )
(i.e., equivalently, (r1 + r2 ) log[ (s1 +s2 ) ] < r1 log[ s1 ] + r2 log[ s2 ].)
r r
1 2 1 2
and
Thus, from the two equivalents, we’ve proved case 2:

150 Stanford Encyclopedia of Philosophy Spring 2021 Edition 151


Inductive Logic James Hawthorne

(s1 + s2 + s3 + … + sm ) = 1. (r1 + r2 + r3 + … + rm )
1=
(s1 + s2 + s3 + … + sm )
Now, think of EQI[ck ∣ hi /hj ∣ b] as generated by applying the theorem in r
= 1
successive steps: s1
(r + r3 + … + rm )
1 = 2
0 = 1 × log[ ] (s2 + s3 + … + sm )
1 r2
=
(r1 + r2 + r3 + … + rm )
= ( r1 + r2 + r3 + … + rm ) × log[
(s1 + s2 + s3 + … + sm ) ]
s2
(r3 + … + rm )
=
(r + r3 + … + rm ) (s3 + … + sm )
≤ r1 × log[ 1 ] + (r2 + r3 + … + rm ) × log[ 2
(s2 + s3 + … + sm ) ]
r
s1 = r3 / s3
≤ r1 × log[ 1 ] + r2 × log[ 2 ]
r r =…
s1 s2 r
= m.
(r + … + rm )
+ ( r3 + … + rm ) × log[ 3
(s3 + … + sm ) ]
sm

≤… That is,
m
≤ ∑ ru × log[ u ] EQI[ck ∣ hi /hj ∣ b] = 0
r
u=1
su
= EQI[ck ∣ hi /hj ∣ b]. just in case for all oku such that P[oku ∣ hj ⋅ b] > 0 and P[oku ∣ hi ⋅ b] > 0,

P[oku ∣ hi ⋅ b ⋅ ck ]
The theorem also says that at each step equality holds just in case = 1.
P[oku ∣ hj ⋅ b ⋅ ck ]
ru (r + ru+1 + … + rm )
= u ,
su (su + su+1 + … + sm ) Otherwise,

which itself holds just in case EQI[ck ∣ hi /hj ∣ b] > 0;

ru (r + … + rm )
= u+1 .
and for each successive step in partitioning the outcome space to generate
su (su+1 + … + sm ) EQI[ck ∣ hi /hj ∣ b], if

(ru + ru+1 + … + rm )
ru /su ≠ ,
So,
(su + su+1 + … + sm )
EQI[ck ∣ hi /hj ∣ b] = 0
we have the strict inequality:
just in case
( + + +…+ m) ( + +…+ m)
= 1 2 3
( u + u+1 +…+ m) × log[ u u+1
]
152 Stanford Encyclopedia of Philosophy Spring 2021 Edition 153
Inductive Logic James Hawthorne

(ru + ru+1 + … + rm )
(ru + ru+1 + … + rm ) × log[
(su + su+1 + … + sm ) ]
A support function is a function Pα from pairs of sentences of L to
real numbers between 0 and 1 (inclusive) that satisfies the
(r + … + rm )
< ru × log[ u ] + (ru+1 + … + rm ) × log[ u+1
(su+1 + … + sm ) ]
r
.
following axioms:
su
(1) Pα [D ∣ E] ≠ 1 for at least one pair of sentences D and E.
So each such division of (oku ∨ oku+1 ∨ … ∨ okm ) into two separate
statements, oku and (oku+1 ∨ … ∨ okm ), adds a strictly positive For all sentence A, B, and C,
contribution to the size of EQI[ck ∣ hi /hj ∣ b] just when
(2) If B ⊨ A, then Pα [A ∣ B] = 1 ;
P[oku ∣ hi ⋅ b ⋅ ck ] P[(oku+1 ∨ … ∨ okm ) ∣ hi ⋅ b ⋅ ck ] (3) If B ⊨ C and C ⊨ B, then Pα [A ∣ B] = Pα [A ∣ C] ;
≠ .
P[oku ∣ hj ⋅ b ⋅ ck ] P[(oku+1 ∨ … ∨ okm ) ∣ hj ⋅ b ⋅ ck ] (4) If C ⊨ ∼(B ⋅ A) , then either

Notes to Inductive Logic Pα [(A ∨ B) ∣ C] = Pα [A ∣ C] + Pα [B ∣ C]

1. Although enumerative inductive arguments may seem to be similar to or else Pα [D ∣ C] = 1 for every D;
what classical statisticians call estimation, it is not really the same thing. (5) Pα [(A ⋅ B) ∣ C] = Pα [A ∣ (B ⋅ C)] × Pα [B ∣ C].
As classical statisticians are quick to point out, estimation does not use the
The results stated in the main text derive from the axioms in the main text
sample to inductively support a conclusion about the whole population.
as follows:
Estimation is not supposed to be a kind of inductive inference. Rather,
estimation is a decision strategy. The sample frequency will be within two a. If B ⊨ A, then Pα [A ∣ B] = 1 .
standard deviations of the population frequency in about 95% of all
samples. So, if one adopts the strategy of accepting as true the claim that Proof: First notice that (by axiom 2) all logical entailments must
the population frequency is within two standard deviations of the sample have the same support value: for, when B ⊨ A and D ⊨ C we have
frequency, and if one uses this strategy repeatedly for various samples, one Pα [A ∣ B] ≥ Pα [C ∣ D] ≥ Pα [A ∣ B]; so Pα [C ∣ D] = Pα [A ∣ B] .
should be right about 95% of the time. I discuss enumerative induction in Thus, all logical entailments B ⊨ A must have the same real number
much more detail in the supplement on Enumerative Inductions: Bayesian support value r : Pα [A ∣ B] = r whenever B ⊨ A. To see that r must
Estimation and Convergence, which treats Bayesian convergence and the equal either 1 or 0, use axiom 5 as follows (employing the various
satisfaction of the CoA for the special case of enumerative inductions. logical entailments involved):

2. Here are the more usual axioms for conditional probability functions. r = Pα [A ⋅ A ∣ A] = Pα [A ∣ A ⋅ A] × Pα [A ∣ A] = r × r;
These axioms are provably equivalent to the axioms presented in the main
text. so r = r × r; so r = 1 or r = 0 . To see that r ≠ 0 , let us suppose
r = 0 and derive a contradiction. Axiom 1 together with axiom 2

154 Stanford Encyclopedia of Philosophy Spring 2021 Edition 155


Inductive Logic James Hawthorne

requires that for some A and B, Pα [A ∣ B] < Pα [A ∣ A] = r = 0 . Proof: Result 1 (above) together with axiom 2 implies that
Then, for this A and B, since B ⊨ (A ∨ ∼A) and B ⊨ ∼(A ⋅ ∼A), 1 ≥ Pα [A ∣ B] , for all A, B. Suppose that for some A and B,
from axiom 4 we have Pα [A ∣ B] < 0 ; then

0 = r = Pα [(A ∨ ∼A) ∣ B] = Pα [A ∣ B] + Pα [∼A ∣ B] Pα [∼A ∣ B] = 1 − Pα [A ∣ B] > 1,

(or else Pα [C ∣ B] = Pα [B ∣ B] = 0 for every sentence C, which contradiction! So we must have Pα [A ∣ B] ≥ 0 . Thus, for all
contradicts our supposition that Pα [A ∣ B] < 0). But then we must sentences A and B, 1 ≥ Pα [A ∣ B] ≥ 0.
have Pα [∼A ∣ B] > 0 = r = Pα [A ∣ A], which contradicts axiom 2.
e. If B ⊨ A, then Pα [A ∣ C] ≥ Pα [B ∣ C].
b. If C ⊨ ∼(B ⋅ A) , then either
Proof: Suppose B ⊨ A. Then ⊨ ∼(B ⋅ ∼A), so
Pα [(A ∨ B) ∣ C] = Pα [A ∣ C] + Pα [B ∣ C]
1 ≥ Pα [(B ∨ ∼A) ∣ C]
or else Pα [E ∣ C] = 1 for every sentence E. = Pα [B ∣ C] + Pα [∼A ∣ C]
= Pα [B ∣ C] + 1 − Pα [A ∣ C].
Proof: Since E ⊨ E, Pα [E ∣ E] = 1 . Thus, axiom 4 becomes: If
C ⊨ ∼(B ⋅ A), then either Thus, Pα [A ∣ C] ≥ Pα [B ∣ C] .

Pα [(A ∨ B) ∣ C] = Pα [A ∣ C] + Pα [B ∣ C] f. If B ⊨ A and A ⊨ B, then Pα [A ∣ C] = Pα [B ∣ C].

or else Pα [E ∣ C] = 1 for every sentence E. Proof: Suppose B ⊨ A and A ⊨ B. Then, from the previous result we
have Pα [A ∣ C] = Pα [B ∣ C].
c. Pα [∼A ∣ B] = 1 − Pα [A ∣ B] or else Pα [C ∣ B] = 1 for every
sentence C. g. If C ⊨ B, then Pα [(A ⋅ B) ∣ C] = Pα [(B ⋅ A) ∣ C] = Pα [A ∣ C].

Proof: B ⊨ (A ∨ ∼A) and B ⊨ ∼(A ⋅ ∼A), so Proof: Suppose C ⊨ B. Then, (A ⋅ C) ⊨ B , so Pα [B ∣ (A ⋅ C)] = 1 ,


by result 1 above. Then, by axiom 5,
1 = Pα [(A ∨ ∼A) ∣ B] = Pα [A ∣ B] + Pα [∼A ∣ B]
Pα [(B ⋅ A) ∣ C] = Pα [B ∣ (A ⋅ C)] × Pα [A ∣ C]
or else Pα [C ∣ B] = 1 for every sentence C. Thus, = Pα [A ∣ C].
Pα [∼A ∣ B] = 1 − Pα [A ∣ B] or else Pα [C ∣ B] = 1 for every
Thus, since (A ⋅ B) ⊨ (B ⋅ A) and (B ⋅ A) ⊨ (A ⋅ B), result 6 above
sentence C.
yields
d. 1 ≥ Pα [A ∣ B] ≥ 0.

156 Stanford Encyclopedia of Philosophy Spring 2021 Edition 157


Inductive Logic James Hawthorne

Pα [(A ⋅ B) ∣ C] = Pα [(B ⋅ A) ∣ C] Proof: If Pα [D ∣ C] = 1 for every D, then each term in the result is
= Pα [A ∣ C]. 1, so the result holds. So, let’s suppose that Pα [D ∣ C] ≠ 1 for some

h. If C ⊨ B and B ⊨ C, then Pα [A ∣ B] = Pα [A ∣ C] .
D. Notice that

(A ∨ B) ⊨ (A ∨ (∼A ⋅ B))
Proof: Suppose C ⊨ B and B ⊨ C. Then, from previous results
together with axioms 5 and 3 we have and
Pα [A ∣ B] = Pα [(A ⋅ C) ∣ B] (A ∨ (∼A ⋅ B)) ⊨ (A ∨ B),
= Pα [A ∣ (C ⋅ B)] × Pα [C ∣ B]
= Pα [A ∣ (C ⋅ B)] and also
= Pα [A ∣ (B ⋅ C)]
= Pα [A ∣ (B ⋅ C)] × Pα [B ∣ C] C ⊨ ∼(A ⋅ (∼A ⋅ B)).
= Pα [(A ⋅ B) ∣ C]
= Pα [A ∣ C]
We’ll also be using the fact that

B ⊨ ((A ⋅ B) ∨ (∼A ⋅ B))


i. If Pα [B ∣ C] > 0 , then

Pα [A ∣ C] and
Pα [A ∣ (B ⋅ C)] = Pα [B ∣ (A ⋅ C)] ×
Pα [B ∣ C]
((A ⋅ B) ∨ (∼A ⋅ B)) ⊨ B,
(simple form of Bayes’ theorem).
and also
Proof: Axioms 5 together with result 6 yields
C ⊨ ∼((A ⋅ B) ⋅ (∼A ⋅ B)).
Pα [A ∣ (B ⋅ C)] × Pα [B ∣ C] = Pα [(A ⋅ B) ∣ C]
= Pα [(B ⋅ A) ∣ C] Then
= Pα [B ∣ (A ⋅ C)] × Pα [A ∣ C].

Thus, if Pα [B ∣ C] > 0 , then

Pα [A ∣ C]
Pα [A ∣ (B ⋅ C)] = Pα [B ∣ (A ⋅ C)] × .
Pα [B ∣ C]

j. Pα [(A ∨ B) ∣ C] = Pα [A ∣ C] + Pα [B ∣ C] − Pα [(A ⋅ B) ∣ C].

158 Stanford Encyclopedia of Philosophy Spring 2021 Edition 159


Inductive Logic James Hawthorne

Pα [(A ∨ B) ∣ C] = Pα [(A ∨ (∼A ⋅ B)) ∣ C] Pα [(((B1 ∨ B2 ) ∨ … ∨ Bn−1 ) ∨ Bn ) ∣ C]


= Pα [A ∣ C] + Pα [(∼A ⋅ B) ∣ C] = Pα [((B1 ∨ B2 ) ∨ … ∨ Bn−1 ) ∣ C] + Pα [Bn ∣ C]
= Pα [A ∣ C] + Pα [(∼A ⋅ B) ∣ C] + Pα [B ∣ C] =…
− Pα [B ∣ C] n

= Pα [A ∣ C] + Pα [(∼A ⋅ B) ∣ C] + Pα [B ∣ C] = ∑ Pα [Bi ∣ C].


− Pα [(A ⋅ B) ∨ (∼A ⋅ B) ∣ C]
i=1

= Pα [A ∣ C] + Pα [(∼A ⋅ B) ∣ C] + Pα [B ∣ C] 3. This result is not the rule commonly known as countable additivity.
− Pα [(A ⋅ B) ∣ C] − Pα [(∼A ⋅ B) ∣ C] Countable additivity requires a language in which infinite disjunctions are
= Pα [A ∣ C] + Pα [B ∣ C] − Pα [(A ⋅ B) ∣ C]. defined.

k. If {B1 , … , Bn } is any finite set of sentences such that for each pair Bi The present result follows directly from the previous result (without
and Bj , C ⊨ ∼(Bi ⋅ Bj ) (i.e., the members of the set are mutually appealing to countable additivity), since, by definition
exclusive, given C), then either Pα [D ∣ C] = 1 for every sentence D,
∞ n

∑ Pα [Bi ∣ C] = lim Pα [Bi ∣ C].


or
n ∑
n i=1 i=1
Pα [((B1 ∨ B2 ) ∨ … ∨ Bn ) ∣ C] = ∑ Pα [Bi ∣ C].
i=1 So, under the conditions stated in the main text,

Proof: For each distinct i and j, let C ⊨ ∼(Bi ⋅ Bj ); and suppose that
lim Pα [((B1 ∨ B2 ) ∨ … ∨ Bn ) ∣ C] = ∑ Pα [Bi ∣ C].
Pα [D ∣ C] < 1 for at least one sentence D. First notice that we have, n
i=1
for each i greater than 1 and less than n,
Given a language in which infinite disjunction is defined, countable
C ⊨ (∼(B1 ⋅ Bi+1 ) ⋅ … ⋅ ∼(Bi ⋅ Bi+1 )); additivity would then result from the following rule (or axiom):

so Pα [((B1 ∨ B2 ) ∨ …) ∣ C] = limnPα [((B1 ∨ B2 ) ∨ … ∨ Bn ) ∣ C].

C ⊨ ∼(((B1 ∨ B2 ) ∨ … ∨ Bi ) ⋅ Bi+1 ). 4. Here are the usual axioms when unconditional probability is taken as
basic:
Then, for any finite list of the first n of the Bi (for each value of n),
Pα is a function from statements to real numbers between 0 and 1 that
satisfies the following rules:

i. if ⊨ A (i.e., if A is a logical truth), then Pα [A] = 1;


ii. if ⊨ ∼(A ⋅ B) (i.e., if A and B are logically incompatible), then

160 Stanford Encyclopedia of Philosophy Spring 2021 Edition 161


Inductive Logic James Hawthorne

Pα [(A ∨ B)] = Pα [A] + Pα [B] ; same theory. The view suggested here, however, is not positivism, but its
inverse, which should be much less controversial: different likelihoods
Definition: if Pα [B] > 0, then implies different theories. That is, given that all of the relevant background
Pα [(A ⋅ B)] and auxiliaries are made explicit (represented in ‘b’), if two scientists
Pα [A ∣ B] = .
Pα [B] disagree significantly about the likelihoods of important evidence claims
on a given hypothesis, they must understand the empirical content of that
5. Bayesians often refer to the probability of an evidence statement on a hypothesis quite differently. To that extent, though they may employ the
hypothesis, P[e ∣ h ⋅ b ⋅ c], as the likelihood of the hypothesis. This can be same syntactic expressions, they use them to express empirically distinct
a somewhat confusing convention since it is clearly the evidence that is hypotheses.
made likely to whatever degree by the hypothesis. So, I will disregard the
usual convention here. Also, presentations of probabilistic inductive logic 8. Call an object grue at a given time just in case either the time is earlier
often suppress c and b, and simply write ‘P[e ∣ h]’. But c and b are than the first second of the year 2030 and the object is green or the time is
important parts of the logic of the likelihoods. So I will continue to make not earlier than the first second of 2030 and the object is blue. Now the
them explicit. statement ‘All emeralds are green (at all times)’ has the same syntactic
structure as ‘All emeralds are grue (at all times)’. So, if syntactic structure
6. These attempts have not been wholly satisfactory thus far, but research determines priors, then these two hypotheses should have the same prior
continues. For an illuminating discussion of the logic of direct inference probabilities. Indeed, both should have prior probabilities approaching 0.
and the difficulties involved in providing a formal account, see the series For, there are an infinite number of competitors of these two hypotheses,
of papers: Levi 1977; Kyburg 1978; Levi 1978. Levi 1980 develops a very each sharing the same syntactic structure: consider the hypotheses ‘All
sophisticated approach. emeralds are gruen (at all times)’, where an object is gruen at a given time
just in case either the time is earlier than the first second of the nth day
Kyburg has developed a logic of statistical inference based solely on
after January 1, 2030, and the object is green or the time is not earlier than
logical direct inference probabilities (Kyburg 1974). Kyburg’s logical
the first second of the nth day after January 1, 2030, and the object is blue.
probabilities do not satisfy the usual axioms of probability theory. The
A purely syntactic specification of the priors should assign all of these
series of papers cited above compares Kyburg’s approach to a kind of
hypotheses the same prior probability. But these are mutually exclusive
Bayesian inductive logic championed by Levi (e.g., in Levi 1967).
hypotheses; so their prior probabilities must sum to a value no greater than
7. This idea should not be confused with logical positivism or logical 1. The only way this can happen is for ‘All emeralds are green’ and each
empiricism. A version of logical positivism applied to likelihoods would of its gruen competitors to have prior probability values either equal to 0
hold that if two theories assign the same likelihood values to all possible or infinitesimally close to it.
evidence claims, then they are essentially the same theory, though they
9. This assumption may be substantially relaxed without affecting the
may be couched in different words. In short: same likelihoods implies
analysis below; we might instead only suppose that the ratios

162 Stanford Encyclopedia of Philosophy Spring 2021 Edition 163


Inductive Logic James Hawthorne

Pα [cn ∣ hj ⋅ b]/Pα [cn ∣ hi ⋅ b] are bounded so as not to get exceptionally possibly do better?—could somehow employ evidence to confirm the true
far from 1. If that supposition were to fail, then the mere occurrence of the hypothesis over evidentially equivalent rivals?) If the true hypothesis has
experimental conditions would count as very strong evidence for or evidentially equivalent rivals, then the convergence result implies that the
against hypotheses—a highly implausible effect. Our analysis could odds against the disjunction of the true hypothesis with these rivals very
include such bounded condition-ratios, but this would only add inessential probably goes to 0, so the posterior probability of this disjunction goes to
complexity to our treatment. 1. Among evidentially equivalent hypotheses the ratio of their posterior
probabilities equals the ratio of their priors:
10. For example, when a new disease is discovered, a new hypothesis hu+1
Pα [hj ∣ b ⋅ cn ⋅ en ] Pα [hj ∣ b]
= .
about that disease being a possible cause of patients’ symptoms is made
explicit. The old catch-all was, “the symptoms are caused by some Pα [hi ∣ b ⋅ cn ⋅ en ] Pα [hi ∣ b]
unknown disease—some disease other than h1 , … , hu”. So the new catch-
So the true hypothesis will have a posterior probability near 1 (after
all hypothesis must now state that “the symptoms are caused by one of the
remaining unknown diseases—some disease other than h1 , … , hu , hu+1”.
evidence drives the posteriors of evidentially distinguishable rivals near to
0) just in case plausibility arguments and considerations (expressed in b)
And, clearly,
make each evidentially indistinguishable rival so much less plausible by
Pα [hK ∣ b] = Pα [∼h1 ⋅ … ⋅ ∼ hu ∣ b] comparison that the sum of each of their comparative plausibilities (as
= Pα [∼h1 ⋅ … ⋅ ∼hu ⋅ (hu+1 ∨ ∼hu+1) ∣ b] compared to the true hypothesis) remains very small.
= Pα [∼h1 ⋅ … ⋅ ∼ hu ⋅ ∼hu+1 ∣ b] + Pα [hu+1 ∣ b]
= Pα [hK∗ ∣ b] + Pα [hu+1 ∣ b].
One more comment about this. It is tempting to identify evidential
distinguishability (via the evidential likelihoods) with empirical
Thus, the new hypothesis hu+1 is “peeled off” of the old catch-all distinguishability. But many plausibility arguments in the sciences, such
hypothesis hK , leaving a new catch-all hypothesis hK∗ with a prior as thought experiments, draw on broadly empirical considerations, on
probability value equal to that of the old catch-all minus the prior of the what we know or strongly suspect about how the world works based on
new hypothesis. our experience of the world. Although this kind of “evidence” may not be
representable via evidential likelihoods (because the hypotheses it bears on
11. This claim depends, of course, on hi being evidentially distinct from don’t deductively or probabilistically imply it), it often plays an important
each alternative hj . I.e., there must be conditions ck with possible role in scientific assessments of hypotheses—in assessments of whether a
outcomes oku on which the likelihoods differ: hypothesis is so extraordinary that only really extraordinary likelihood
P[oku ∣ hi ⋅ b ⋅ ck ] ≠ P[oku ∣ hj ⋅ b ⋅ ck ].
evidence could rescue it. It is (arguably) a distinct virtue of the Bayesian
logic of evidential support that it permits such considerations to be figured
Otherwise hi and hj are empirically equivalent, and no amount of evidence into the net support for hypotheses.
can support one over the other. (Did you think a confirmation theory could

164 Stanford Encyclopedia of Philosophy Spring 2021 Edition 165


Inductive Logic James Hawthorne

12. This is a good place to describe one reason for thinking that inductive Here is the problem. If the agent is already certain of an evidence
support functions must be distinct from subjectivist or personalist degree- statement e, then her belief-function likelihoods for that statement must be
of-belief functions. Although likelihoods have a high degree of objectivity 1 on every hypothesis. I.e., if Qγ is her belief function and Qγ [e] = 1 , then
in many scientific contexts, it is difficult for belief functions to properly it follows from the axioms of probability theory that Qγ [e ∣ hi ⋅ b ⋅ c] = 1,
represent objective likelihoods. This is an aspect of the problem of old regardless of what hi says—even if hi implies that e is quite unlikely
evidence. (given b ⋅ c). But the problem goes even deeper. It not only applies to
evidence that the agent knows with certainty. It turns out that almost
Belief functions are supposed to provide an idealized model of belief anything the agent learns that can change how strongly she believes e will
strengths for agents. They extend the notion of ideally consistent belief to also influence the value of her belief-function likelihood for e, because
a probabilistic notion of ideally coherent belief strengths. There is no harm Qγ [e ∣ hi ⋅ b ⋅ c] represents the agent’s belief strength given everything she
in this kind of idealization. It is supposed to supply a normative guide for knows.
real decision making. An agent is supposed to make decisions based on
her belief-strengths about the state of the world, her belief strengths about To see the difficulty with less-than-certain evidence, consider the
possible consequences of actions, and her assessment of the desirability following example. Let e be any statement that is statistically implied to
(or utility) of these consequences. But the very role that belief functions degree r by a hypothesis h together with experimental conditions c (e.g., e
are supposed to play in decision making makes them ill-suited to inductive says “the coin lands heads on the next toss” and h ⋅ c says “the coin is fair
inferences where the likelihoods are often supposed to be objective, or at and is tossed in the usual way on the next toss”). Then the correct
least possess inter-subjectively agreed values that represent the empirical objective likelihood value is just P[e ∣ h ⋅ c] = r (e.g., for r = 1/2). Let d
import of hypotheses. For the purposes of decision making, degree-of- be a statement that is intuitively not relevant in any way to how likely e
belief functions should represent the agent’s belief strengths based on should be on h ⋅ c (e.g., let d say “Jim will be really pleased with the
everything she presently knows. So, degree-of-belief likelihoods must outcome of that next toss”). Suppose some rational agent has a degree-of-
represent how strongly the agent would believe the evidence if the belief function Q for which the likelihood for e due to h ⋅ c agrees with the
hypothesis were added to everything else she presently knows. However, objective value: Q[e ∣ h ⋅ c] = r (e.g., with r = 1/2) .
support-function likelihoods are supposed to represent what the hypothesis
(together with explicit background and experimental conditions) says or Our analysis will show that this agent’s belief-strength for d given
implies about the evidence. As a result, degree-of-belief likelihoods are ∼e ⋅ h ⋅ c will be a relevant factor; so suppose that her degree-of-belief in
saddled with a version of the problem of old evidence, a problem not that regard has any value s other than 1: Q[d ∣ ∼e ⋅ h ⋅ c] = s < 1 (e.g.,
shared by support function likelihoods. Furthermore, it turns out that the suppose s = 1/2) . This is a very weak supposition. It only says that adding
old evidence problem for likelihoods is much worse than is usually ∼e ⋅ h ⋅ c to everything else the agent currently knows leaves her less than
recognized. certain that d is true.

166 Stanford Encyclopedia of Philosophy Spring 2021 Edition 167


Inductive Logic James Hawthorne

Qnew [e ⋅ h ⋅ c]
Now, suppose this agent learns the following bit of new information in a Qnew[e ∣ h ⋅ c] =
completely convincing way (e.g., I seriously tell her so, and she believes Qnew[h ⋅ c]
me completely): (d ∨ e) (i.e., Jim will be really pleased with the outcome Q[e ⋅ h ⋅ c ∣ (d ∨ e)]
=
of the next toss unless it comes up heads). Q[h ⋅ c ∣ (d ∨ e)]
Q[(d ∨ e) ⋅ (e ⋅ h ⋅ c)]
Thus, on the usual Bayesian degree-of-belief account the agent is =
supposed to update her belief function Q to arrive at a new belief function Q[(d ∨ e) ⋅ (h ⋅ c)]
Qnew by the updating rule: Q[(d ∨ e) ⋅ e ∣ h ⋅ c]
=
Q[(d ∨ e) ∣ h ⋅ c]
Qnew [S] = Q[S ∣ (d ∨ e)], for each statement S.
Q[e ∣ h ⋅ c]
=
However, this update of the agent’s belief function has to screw up the Q[((d ⋅ ∼e) ∨ e) ∣ h ⋅ c]
objectivity of her new belief-function likelihood for e on h ⋅ c, because she Q[e ∣ h ⋅ c]
=
now should have: Q[e ∣ h ⋅ c] + Q[d ⋅ ∼e ∣ h ⋅ c]]
Q[e ∣ h ⋅ c]
=
[Q[e ∣ h ⋅ c] + Q[d ∣ ∼e ⋅ h ⋅ c] × Q[∼e ∣ h ⋅ c]]
r
=
[r + s × (1 − r)]
1
= .
(1−r)
[1 + s × r
]

Thus, the updated belief function likelihood must have value

1
Qnew [e ∣ h ⋅ c] = .
(1−r)
1+s× r

This factor can be equal to the correct likelihood value r just in case s = 1 .
For example, for r = 1/2 and s = 1/2 we get Qnew [e ∣ h ⋅ c] = 2/3.

The point is that even the most trivial knowledge of disjunctive claims
involving e may completely upset the value of the likelihood for an agent’s
belief function. And an agent will almost always have some such trivial

Qnew [e ⋅ h ⋅ c]
=
168 Stanford Encyclopedia of Philosophy Spring 2021 Edition 169
Inductive Logic James Hawthorne

knowledge. Updating on such conditionals can force the agent’s belief Pα [hj ∣ b ⋅ cn ⋅ en ]
functions to deviate widely from the evidentially relevant objective values Pα [hi ∣ b ⋅ cn ⋅ en ]
of likelihoods on which scientific hypotheses should be tested.
goes to 0; and that must occur if the likelihood ratios
More generally, it can be shown that the incorporation into a belief
P[en ∣ hj ⋅ b ⋅ cn ]
P[en ∣ hi ⋅ b ⋅ cn ]
function Q of almost any kind of evidence for or against the truth of a
prospective evidence claim e—even uncertain evidence for e, as may
come through Jeffrey updating—completely undermines the objective or approach 0, provided that and the prior probability Pα [hi ∣ b] is greater
inter-subjectively agreed likelihoods that a belief function might have than 0. The Likelihood Ratio Convergence Theorem will show that when
expressed before updating. This should be no surprise. The agent’s belief hi ⋅ b is true, it is very likely that the evidence will indeed be such as to
function likelihoods reflect her total degree-of-belief in e, based on a drive the likelihood ratios as near to 0 as you please, for a long enough (or
hypothesis h together with everything else she knows about e. So the strong enough) evidence stream. (If the stream is strong in that the
agent’s present belief function may capture appropriate public likelihoods likelihood ratios of individual bits of evidence are small, then to bring
for e only if e is completely isolated from the agents other beliefs. And about a very small cumulative likelihood ratio, the evidence stream need
this will rarely be the case. not be as long.) As likelihood ratios head towards 0, the only way a
Bayesian agent can avoid having her inductive support function(s) yield
One Bayesian subjectivist response to this kind of problem is that the posterior probabilities for hj that approach 0 (as n gets large) is to
belief functions employed in scientific inductive inferences should often be continually change her prior probability assessments. That means either
“counterfactual” belief functions, which represent what the agent would continually finding and adding new plausibility arguments (i.e., adding to
believe if e were subtracted (in some suitable way) from everything else or modifying b) that on balance favor hj over hi , or continually reassessing
she knows (see, e.g., Howson & Urbach 1993). However, our examples the support strength due to plausibility arguments already available, or
show that merely subtracting e won’t do. One must also subtract any both.
disjunctive statements containing e. And it can be shown that one must
subtract any uncertain evidence for or against e as well. So the Technically, continual reassessments of support strengths that favor hj
counterfactual belief function idea needs a lot of working out if it is to over hi based on already extant arguments (in b) means switching to new
rescue the idea that subjectivist Bayesian belief functions can provide a support functions (or new vagueness sets of them) that assign hj ever
viable account of the likelihoods employed by the sciences in inductive higher prior probabilities as compared to hi based on the same arguments
inferences. in b. In any case, such revisions of argument strengths may avoid the
convergence towards 0 of the posterior probability of hj only if it proceeds
13. That is, for each inductive support function Pα , the posterior at a rate that keeps ahead of the rate at which the evidence drives the
Pα [hj ∣ b ⋅ cn ⋅ en ] must go to 0 as the ratio likelihood ratios towards 0.

Pα [hj ∣ b ⋅ n ⋅ n]

170 Stanford Encyclopedia of Philosophy Spring 2021 Edition 171


Inductive Logic James Hawthorne

For a thorough presentation of the most prominent Bayesian convergence parameter value (and because of that, failed to be result-independent of the
results and a discussion of their weaknesses see Earman 1992: Ch. 6. parameter value evidence relative to the original hypothesis) is result-
However, Earman does not discuss the convergence theorems under independent of the parameter value evidence relative to each of these more
consideration here (due to the fact that the convergence results discussed precise hypotheses—because each of the precise hypotheses already
here first appeared in Hawthorne 1993, just after Earman’s book came identifies precisely what (it claims) the value of the parameter is. Thus,
out). wherever the workings of the logic of evidential support is made more
perspicuous by treating evidence as composed of result-independent
14. In scientific contexts all of the most important kinds of cases where chunks, one may treat hypotheses whose unspecified parameter values
large components of the evidence fail to be result-independent of one interfere with result-independence as disjunctively composite hypotheses,
another are cases where some part of the total evidence helps to tie down and apply the evidential logic to these more specific disjuncts, and thereby
the numerical value of a parameter that plays an important role in the regain result-independence.
likelihood values the hypothesis specifies for other large parts of the total
evidence. In cases where this only happens rather locally, where the 15. Technically, suppose that Ok can be further “subdivided” into more
evidence for a parameter value influences the likelihoods of only a very outcome-descriptions by replacing okv with two mutually exclusive parts,
small part of the total evidence that bears on the hypothesis, we can treat o∗kv and o#kv , to produce new outcome space
the conjunction of the evidence for the parameter value with the evidential
outcomes whose likelihood the parameter value influences as a single Ok$ = {ok1 , … , o∗kv , o#kv , … , okw },
chunk of evidence, which is then result-independent of the rest of the
where
evidence (on each alternative hypothesis). This is the sort of chuncking of
the evidence into result-independent parts suggested in the main text. P[o∗kv ⋅ o#kv ∣ hi ⋅ b ⋅ ck ] = 0

However, in cases where the value of a parameter left unspecified by the and
hypothesis has a wide-ranging influence on many of the likelihood values
the hypothesis specifies, another strategy for obtaining result- P[o∗kv ∣ hi ⋅ b ⋅ ck ] + P[o#kv ∣ hi ⋅ b ⋅ ck ] = P[okv ∣ hi ⋅ b ⋅ ck ];
independence among these components of the evidence will do the job. A
and suppose similar relationships hold for hj . Then the new EQI* (based
hypothesis that has an unspecified parameter value is in effect equivalent
on O∗k ) is greater than or equal to EQI (based on Ok ); and EQI∗ > EQI
to a disjunction of more specific hypotheses, where each disjunct consists
just in case at least one of the new likelihood ratios, e.g.,
of a more precise version of the original hypothesis, a version in which the
value for the parameter has been “filled in”. Relative to each of these more P[o∗kv ∣ hi ⋅ b ⋅ ck ]
,
precise hypotheses, any evidence for or against the parameter value that P[o∗kv ∣ hj ⋅ b ⋅ ck ]
hypothesis specifies is evidence for or against that more precise hypothesis
itself. Furthermore, the evidence whose likelihood values depend on the differs in value from the “undivided” outcome’s likelihood ratio,
P[ ∣ hi ⋅ b ⋅ k]
kv
.
172 Stanford Encyclopedia of Philosophy Spring 2021 Edition 173
Inductive Logic James Hawthorne

P[okv ∣ hi ⋅ b ⋅ ck ]
. and we want this likelihood ratio to have a value less than ε. A bit of
P[okv ∣ hi ⋅ b ⋅ ck ] algebraic manipulation shows that to get this likelihood ratio value to be
below ε, the percentage of “heads” needs to be k/n > 12 − 12 (log ε)/n.
The supplement on the effect on EQI of partitioning the outcome space
Using the normal approximation to the binomial distribution (with mean
proves this claim.
= 2/3 and variance = (2/3)(1/3)/n) the actual likelihood of obtaining an
16. The likely rate of convergence will almost always be much faster than outcome sequence having more than 12 − 12 (log ε)/n “heads” (which we
the worst case bound provided by Theorem 2. To see the point more just saw corresponds to getting a likelihood ratio less than ε, thus
clearly, let’s look at a very simple example. Suppose hi says that a certain disfavoring the 1/3 propensity hypothesis as compared to the 2/3
bent coin has a propensity for “heads” of 2/3 and hj says the propensity is propensity hypothesis by that much) when the true propensity for “heads”
1/3. Let the evidence stream consist of outcomes of tosses. In this case the is 2/3 is given by the formula

⎡ mean − 1 − 1 (log ε) ⎤
average EQI equals the EQI of each toss, which is 1/3; and the smallest
possible likelihood ratio occurs for “heads”, which yields the value γ = 12 . ⎢( ( 2 2 n )) ⎥ 3(log ε)
⎥ = Φ [(1/8) 2 n 2 (1 +
n )]
Φ⎢
1 1

So, the value of the lower bound given by Theorem 2 for the likelihood of ⎢ (variance) 2
1

getting an outcome sequences with a likelihood ratio below ε (for hj over ⎣ ⎦
hi ) is
(where Φ[x] gives the value of the standard normal distribution from −∞
(1/n)(log )
1 2 to x). Now let ε = 1/16 (= .0625), as before. So the actual likelihood of
2 9
1− =1− . obtaining a stream of outcomes with likelihood ratio this small when hi is
((1/3) + (log ε)/n)2 (n × (1 + 3(log ε)/n)2
true and the number of tosses is n = 52 is Φ[1.96] > .975, whereas the
Thus, according to the theorem, the likelihood of getting an outcome lower bound given by Theorem 2 was .70. And if the number of tosses is
sequence with a likelihood ratio less than ε = 1/16 (=.06) when hi is true increased to n = 204 , the likelihood of obtaining an outcome sequence
and the number of tosses is n = 52 is at least .70; and for n = 204 tosses with a likelihood ratio this small (i.e., ε = 1/16) is Φ[4.75] > .999999,
the likelihood is at least .95. whereas the lower bound from Theorem 2 for this likelihood is .95.
Indeed, to actually get a likelihood of .95 that the evidence stream will
To see the amount by which the lower bound provided by the theorem is in produce a likelihood ratio less than ε >.06, the number of tosses needed is
fact overly cautious, consider what the usual binomial distribution for the only n = 43, rather than the 204 tosses the bound given by the theorem
coin tosses in this example implies about the likely values of the requires in order to get up to the value .95. (Note: These examples employ
likelihood ratios. The likelihood ratio for exactly k “heads” in n tosses is “identically distributed” trials—repeated tosses of a coin—as an

((1/3)k (2/3)n−k )
illustration. But Convergence Theorem 2 applies much more generally. It
= 2n−2k ;
((2/3)k (1/3)n−k )
applies to any evidence sequence, no matter how diverse the probability
distributions for the various experiments or observations in the sequence.)

174 Stanford Encyclopedia of Philosophy Spring 2021 Edition 175


Inductive Logic James Hawthorne

17. It should now be clear why the boundedness of EQI above 0 is 18. In many scientific contexts this is the best we can hope for. But it still
important. Convergence Theorem 2 applies only when provides a very reasonable representation of inductive support. Consider,

−(log ε)
for example, the hypothesis that the land masses of Africa and South
⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯
EQI[cn ∣ hi /hj ∣ b] > . America separated and drifted apart over the eons, the drift hypothesis, as
n
opposed to the hypothesis that the continents have fixed positions acquired
But this requirement is not a strong assumption. For, the Nonnegativity of when the earth first formed and cooled and contracted, the contraction
EQI Theorem shows that the empirical distinctness of two hypotheses on hypothesis. One may not be able to determine anything like precise
a single possible outcome suffices to make the average EQI positive for the likelihoods, on each hypothesis, for the evidence that: (1) the shape of the
whole sequence of experiments. So, given any small fraction ε > 0, the east coast of South America matches the shape of the west coast of Africa
value of −(log ε)/n (which is always greater than 0 when ε < 0) will as closely as it in fact does; (2) the geology of the two coasts match up so
⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯
eventually become smaller than EQI, provided that the degree to which closely when they are “fitted together” in the obvious way; (3) the plant
the hypotheses are empirical distinct for the various observations ck does and animal species on these distant continents should be as similar as they
not on average degrade too much as the length n of the evidence stream are, as compared to how similar species are among other distant
increases. continents. Although neither the drift hypothesis nor the contraction
hypothesis supplies anything like precise likelihoods for these evidential
When the possible outcomes for the sequence of observations are claims, experts readily agree that each of these observations is much more
independent and identically distributed, Theorems 1 and 2 effectively likely on the drift hypothesis than on the contraction hypothesis. That is,
reduce to L. J. Savage’s Bayesian Convergence Theorem [Savage, pg. 52– the likelihood ratio for this evidence on the contraction hypothesis as
54], although Savage’s theorem doesn’t supply explicit lower bounds on compared to the drift hypothesis is very small. Thus, jointly these
the probability that the likelihood ratio will be small. Independent, observations constitute very strong evidence for drift over contraction.
identically distributed outcomes most commonly result from the repetition
of identical statistical experiments (e.g., repeated tosses of a coin, or Historically, the case of continental drift is more complicated. Geologists
repeated measurements of quantum systems prepared in identical states). tended to largely dismiss this evidence until the 1960s. This was not
In such experiments a hypothesis will specify the same likelihoods for the because the evidence wasn’t strong in its own right. Rather, this evidence
⎯⎯⎯⎯⎯⎯⎯⎯⎯⎯
same kinds of outcomes from one observation to the next. So EQI will was found unconvincing because it was not sufficient to overcome prior
remain constant as the number of experiments, n, increases. However, plausibility considerations that made the drift hypothesis extremely
Theorems 1 and 2 are much more general. They continue to hold when the implausible—much less plausible than the contraction hypothesis. The
sequence of observations encompasses completely unrelated experiments problem was that there seemed to be no plausible mechanism by which
that have different distributions on outcomes—experiments that have drift might occur. It was argued, quite plausibly, that no known force could
nothing in common but their connection to the hypotheses they test. push or pull the continents apart, and that the less dense continental
material could not push through the denser material that makes up the
ocean floor. These plausibility objections were overcome when a plausible

176 Stanford Encyclopedia of Philosophy Spring 2021 Edition 177


Inductive Logic James Hawthorne

mechanism was articulated—i.e., the continental crust floats atop molten 19. To see the point of the third clause, suppose it were violated. That is,
material and moves apart as convection currents in the molten material suppose there are possible outcomes for which the likelihood ratio is very
carry it along. The case was pretty well clinched when evidence for this near 1 for just one of the two support functions. Then, even a very long
mechanism was found in the form of “spreading zones” containing sequence of such outcomes might leave the likelihood ratio for one
alternating strips of magnetized material at regular distances from mid- support function almost equal to 1, while the likelihood ratio for the other
ocean ridges. The magnetic alignments of materials in these strips support function goes to an extreme value. If that can happen for support
correspond closely to the magnetic alignments found in magnetic functions in a class that represent likelihoods for various scientists in the
materials in dateable sedimentary layers at other locations on the earth. community, then the empirical contents of the hypotheses are either too
These magnetic alignments indicate time periods when the direction of vague or too much in dispute for meaningful empirical evaluation to occur.
earth’s magnetic field has reversed. And this gave geologists a way of
measuring the rate at which the sea floor might spread and the continents 20. If there are a few directionally controversial likelihood ratios, where
move apart. Although geologists may not be able to determine anything Pα says the ratio is somewhat greater than 1, while and Pβ assigns a value
like precise values for the likelihoods of any of this evidence on each of somewhat less than 1, these may not greatly effect the trend of Pα and Pβ
the alternative hypotheses, the evidence is universally agreed to be much towards agreement on the refutation and support of hypotheses provided
more likely on the drift hypothesis than on the alternative contraction that the controversial ratios are not so extreme as to overwhelm the stream
hypothesis. The likelihood ratio for this evidence on the contraction of other evidence on which the likelihood ratios do directionally agree.
hypothesis as compared to the drift hypothesis is somewhat vague, but Even so, researches will want to get straight on what each hypothesis says
extremely small. The vagueness is only in regard how extremely small the or implies about such cases. While that remains in dispute, the empirical
likelihood ratio is. Furthermore, with the emergence of a plausible content of each hypothesis remains unsettlingly vague.
mechanism, the drift hypothesis is no longer so overwhelmingly
implausible (logically) prior to taking the likelihood evidence into Notes to Supplement on Enumerative Inductions:
account. Thus, even when precise values for individual likelihoods are not Bayesian Estimation and Convergence
available, the value of a likelihood ratio range may be objective enough to
strongly refute one hypothesis as compared to another. Indeed, the drift 21. What it means for a sample to be randomly selected from a population
hypothesis is itself strongly supported by the evidence; for, no alternative is philosophically controversial. Various analyses of the concept have been
hypothesis that has the slightest amount of comparative plausibility can proposed, and disputed. For our purposes an account of the following sort
account for the available evidence nearly so well. That is, no plausible will suffice. To say
alternative makes the evidence anywhere near so likely. Given the
S is a random sample of population B with respect to attribute A
currently available evidence, the only issues left open (for now) involve
comparing various alternative versions of the drift hypothesis (involving means that
differences of detail) against one another.

178 Stanford Encyclopedia of Philosophy Spring 2021 Edition 179


Inductive Logic James Hawthorne

the selection set S is generated by a process that has an objective representative subpopulation C of B—e.g., from registered voters who
chance (or propensity) r of choosing individual objects that have answered the telephone between the hours of 7 PM and 9 PM in the
attribute A from among the objects in population B, where on each middle of the week. Of course, the claim that a given subpopulation C is
selection the chance value r agrees with the value r of the frequency of representative is itself a hypothesis that is open to inductive support by
As among the Bs, F[A, B]. evidence. Professional polling organizations do a lot of research to
calibrate their sampling technique, to find out what sort of subpopulations
Defined this way, randomness implies probabilistic independence among C they may draw on as highly representative. For example, one way to see
the outcomes of selections with regard to whether they exhibit attribute A, if registered voters who answer the phone during the evening, mid-week,
on any given hypothesis about the true value of the frequency r of As are likely to constitute a representative sample is to conduct a large poll of
among the Bs. such voters immediately after an election, when the result is known, to see
how representative of the actual vote count the count from of the
The tricky part of generating a randomly selected set from the population
subpopulation turns out to be.
is to find a selection process for which the chance of selecting an A each
time matches the true frequency without already knowing what the true Notice that although the selection set S is selected from B, S cannot be a
frequency value is—i.e., without already knowing what the value of r is. subset of B, not if S can be generated by sampling with replacement. For, a
However, there clearly are ways to do this. Here is one way: specific member of B may be randomly selected into S more than once. If
S were a subset of B, any specific member of B could only occur once in S.
the sample S is generated by a process that on each selection gives
That is, consider the case where S consists of n selections from B, but
each member of B an equal chance of being selected into S (like
where the process happens to select the same member b of B twice. Then,
drawing balls from a well-shaken urn).
were S a subset of B, although b is selected into S twice, S can only
Here, schematically, is another way: possess b as a member once, so S has at most n − 1 members after all
(even fewer if other members of B are selected more than once). So, rather
find a subclass of B, call it C, from which S can be generated by a than being members of B, the members of S must be representations of
process that gives every member of C an equal chance of being members of B, like names, where the same member of B may be
selected into S, where C is representative of B with respect to A in the represented by different names. However, the representations (or names)
sense that the frequency of A in C is almost precisely the same as the in S technically may not be the sorts of things that can possess attribute A.
frequency of A in B. So, technically, on this way of handling the problem, when we say that a
member of S exhibits A, this is shorthand for the referent of S in B
Pollsters use a process of this kind. Ideally a poll of registered voters,
possesses attribute A.
population B, should select a sample S in a way that gives every registered
voter the same chance of getting selected into S. But that may be 22. This is closely analogous to the Stable-Estimation Theorem of
impractical. However, it suffices if the sample is selected from a Edwards, Lindman, & Savage (1993). Here is a proof of Case 1, i.e.,

180 Stanford Encyclopedia of Philosophy Spring 2021 Edition 181


Inductive Logic James Hawthorne

where the number of members of the reference class B is finite and where K ≥ Pα [F[A, B] = s ∣ b]/L for each value of Pα [F[A, B] = s ∣ b] for
for some integer u at least as large as the size of B there is a specific which s is outside of R; and 1 ≤ Pα [F[A, B] = r ∣ b]/L for each value
(perhaps very large) integer K such that the prior probability of a of Pα [F[A, B] = r ∣ b] for which r is inside of R.
hypothesis stating a frequency outside region R is never more than K times vi. It follows that:
as large as a hypothesis stating a frequency within region R. (The proof is
Pα [F[A, B] = s ∣ b]
Case 2 is almost exactly the same, but draws on integrals wherever the
∑ s × (1 − s)
m n−m
×( )
L
present proof draws on sums using the ‘∑ ’ expression.) s∉R

≤ ∑ sm × (1 − s)n−m × Pα [F[A, B] = s ∣ b] × K
A few observations before proceeding to the main derivation:
s∉R

i. The hypotheses under consideration consist of all expressions of form and


F[A, B] = k/u, where u is as described above and k is a non-negative
∑ r × (1 − r) × (Pα [F[A, B] = r ∣ b]/L)
m n−m
integer between 0 and u.
ii. R is some set of fractions of form k/u for a contiguous sequence of s∉R

non-negative integers k that includes the sample frequency m/n . ≥ ∑ rm × (1 − r)n−m × Pα [F[A, B] = r ∣ b].
iii. In the following derivation all sums over values r in R are r∈R
abbreviations for sums over integers k such that k/u is in R; similarly,
vii. For β[R, m + 1, n − m + 1] defined as
all sums over values s not in R are abbreviations for sums over
integers k such that k/u is not in R. The sum over {s ∣ s = k/u} ∫R rm (1 − r)n−m dr
1
,
∫0 rm (1 − r)n−m dr
represents the sum over all integers k from 0 through u.

iv. Define L to be the smallest value of a prior probability


Pα [F[A, B] = r ∣ b] for r a fraction in R. Notice that L > 0 because,
when u is large, its an established mathematical fact that

by supposition, finite ∑r∈R rm × (1 − r)n−m


Pα [F[A, B] = s ∣ b] ∑s∈{s∣s=k/u} sm × (1 − s)n−m
K≥
Pα [F[A, B] = r ∣ b]
is extremely close to the value of β[R, m + 1, n − m + 1].
for the largest value of Pα [F[A, B] = s ∣ b] for which s is outside of
R and the smallest value of Pα [F[A, B] = r ∣ b] for which r is outside
We now proceed to the main part of the derivation.
of region R. From the Odds Form of Bayes’ Theorem (Equation 10) we have,
v. Thus, from the definition of L and of K, it follows that:

⎡ ⎤
⎢ ⎥
⎢ ⎥
182 Stanford Encyclopedia of Philosophy α⎢ ⎥
Spring 2021 Edition 183
Inductive Logic James Hawthorne

⎡ F[A, B] ∉ R ⎤ ∑s∈{s∣s=k/u} sm × (1 − s)n−m − ∑r∈R r m × (1 − r)n−m


⎢ ⎥ =K×
∣ F[A, S] = m/n ⋅ Rnd[S, B, A]
Ωα ⎢⎢ ⎥

∑r∈R r m × (1 − r)n−m
⎢ ⋅ Size[S] = n ⎥
⎣ ⋅b ⎦ ∑s∈{s∣s=k/u} sm × (1 − s)n−m
=K× −1
∑r∈R r m × (1 − r)n−m
⎡ F[A, B] = s ⎤
⎢ ⎥
∣ F[A, S] = m/n ⋅ Rnd[S, B, A]
∑s∉R Pα ⎢⎢ ⎥

≈ K × [(1/β[R, m + 1, n − m + 1]) − 1].
⎢ ⋅ Size[S] = n ⎥
⎣ ⋅b ⎦ Thus,
=
⎡ F[A, B] = r ⎤ ⎡ F[A, B] ∉ R ⎤
⎢ ⎥ ⎢ ⎥
∣ F[A, S] = m/n ⋅ Rnd[S, B, A]
∑r∈R Pα ⎢⎢ ⎥
⎥ Ωα ⎢⎢
∣F[A, S] = m/n ⋅ Rnd[S, B, A] ⎥
⎢ ⋅ Size[S] = n ⎥ ⋅ Size[S] = n ⎥
⎣ ⋅b ⎦ ⎢ ⎥
⎣ ⋅b ⎦
⎡ F[A, S] = m/n ⎤ 1
⎢ ⎥ ≤ K × [(
β[R, m + 1, n − m + 1] )
∣ F[A, B] = s ⋅ Rnd[S, B, A] − 1] .
∑s∉R P ⎢⎢ ⎥ × P [F[A, B] = s ∣ b]

⋅ Size[S] = n
α
⎢ ⎥
⎣ ⋅b ⎦ Then by equation (11), which expresses the relationship between posterior
=
⎡ F[A, S] = m/n ⎤ probability and posterior odds against,
⎢ ⎥
∣ F[A, B] = r ⋅ Rnd[S, B, A]
∑r∈R P ⎢⎢ ⎥ × P [F[A, B] = r ∣ b]
⎥ ⎡ F[A, B] ∈ R ⎤
⋅ Size[S] = n ⎢ ⎥
α
⎢ ⎥ ∣ F[A, S] = m/n ⋅ Rnd[S, B, A] ⎥
⎣ ⋅b ⎦ Pα ⎢⎢
⎢ ⋅ Size[S] = n ⎥⎥
∑s∉R sm × (1 − s)n−m × Pα [F[A, B] = s ∣ b] ⎣ ⋅b ⎦
=
∑r∈R r m × (1 − r)n−m × Pα [F[A, B] = r ∣ b] ⎛ ⎡ F[A, B] ∉ R ⎤⎞
⎜ ⎢ ⎥⎟
∣ = ⋅ Rnd[S,
= 1/ ⎜⎜1 + Ωα ⎢⎢ ⎥⎟
F[A, S] m/n B, A]
∑s∉R sm × (1 − s)n−m × (Pα [F[A, B] = s ∣ b]/L) ⎥⎟
= ⎜ ⎢ ⋅ Size[S] = n ⎥⎟
∑r∈R r m × (1 − r)n−m × (Pα [F[A, B] = r ∣ b]/L) ⎝ ⎣ ⋅b ⎦⎠
∑s∉R sm × (1 − s)n−m × K 1
≥ .

( 1 + × [( β[R,m+1,n−m+1]) − 1 ])
1
∑r∈R r m × (1 − r)n−m K

∑s∈{s∣s=k/u} m × (1 − s n−m − ∑r∈R m × (1 − r n−m


=K×
184 Stanford Encyclopedia of Philosophy Spring 2021 Edition 185
Inductive Logic James Hawthorne

⎡ r − .01 < F[A, S] < r + .01 ⎤


P ⎢⎢ ∣ F[A, B] = r ⋅ Rnd[S, B, A] ⎥ ≥ .9999.
23. To get a better idea of the import of this theorem, let’s consider some
specific values. First notice that the factor r × (1 − r) can never be larger ⎥
than (1/2)× (1/2) = 1/4 ; and the closer r is to 1 or 0, the smaller ⎣ ⋅ Size[S] = 38000 ⎦
r × (1 − r) becomes. So, whatever the value of r, the factor q/
As the sample size n becomes larger, it becomes extremely likely that the
((r × (1 − r)/n) 2 ≤ 2× q×n 2 . Thus, for any chosen value of q,
1 1

sample frequency will come to within any specified region close to the
⎡ r − q < F[A, S] < r + q ⎤ true frequency r, as close as you wish.
P ⎢⎢ ∣ F[A, B] = r ⋅ Rnd[S, B, A] ⎥ ≥ 1 − 2 × Φ[−2 × q × n 12 ].

⎣ ⋅ Size[S] = n ⎦ Notes to Supplement on Likelihood Ratios,
Likelihoodism, and the Law of Likelihood
For example, if q = .05 and n = 400, then we have (for any value of r),

⎡ r − .05 < F[A, S] < r + .05 ⎤


24. To see the point more clearly, consider an example. To keep things
P ⎢⎢ ∣ F[A, B] = r ⋅ Rnd[S, B, A] ⎥ ≥ .95. simple, let’s suppose our background b says that the chances of heads for

⎣ ⋅ Size[S] = 400 ⎦
tosses of this coin is some whole percentage between 0% and 100%. Let c
say that the coin is tossed in the usual random way; let e say that the coin
For n = 900 (and margin q = .05) this lower bound raises to .997: comes up heads; and for each r that is a whole fraction of 100 between 0
and 1, let h[r] be the simple statistical hypothesis asserting that the chance
⎡ r − .05 < F[A, S] < r + .05 ⎤
P ⎢⎢ ∣ F[A, B] = r ⋅ Rnd[S, B, A] ⎥ ≥ .997. of heads on each random toss of this coin is r. Now consider the composite
⎥ statistical hypothesis h[>.65] , which asserts that the chance of heads on
⎣ ⋅ Size[S] = 900 ⎦
each random (independent) toss is greater than .65. From the axioms of
If we are interested in a smaller margin of error q, we can keep the same probability we derive the following relationship:

Pα [e ∣ h[>.65] ⋅ c ⋅ b] = P[e ∣ h[.66] ⋅ c ⋅ b] × Pα [h[.66] ∣ h[>.65] ⋅ c ⋅ b]


sample size and find the value of the lower bound for that value of q. For

+ P[e ∣ h[.67] ⋅ c ⋅ b] × Pα [h[.67] ∣ h[>.65] ⋅ c ⋅ b]


example,

⎡ r − .03 < F[A, S] < r + .03 ⎤ +…


P ⎢⎢ ∣ F[A, B] = r ⋅ Rnd[S, B, A] ⎥ ≥ .928.

+ P[e ∣ h[1] ⋅ c ⋅ b] × Pα [h[1] ∣ h[>.65] ⋅ c ⋅ b].
⎣ ⋅ Size[S] = 900 ⎦
The issue for the likelihoodist is that the values of the terms of form
By increasing the sample size the bound on the likelihood can be made as Pα [h[r] ∣ h[>.65] ⋅ c ⋅ b] are not objectively specified by the composite
close to 1 as we want, for any margin q we choose. For example: hypothesis h[>.65] (together with c ⋅ b), but the value of the likelihood
Pα [e ∣ h[>.65] ⋅ c ⋅ b] depends essentially on these non-objective factors.

⎡ ⎤
P⎢ ⎥ ≥ .9999.
186 ⎢ Stanford Encyclopedia⎥of Philosophy
⎣ ⎦
Spring 2021 Edition 187
Inductive Logic

So, likelihoods based on composite statistical hypotheses fail to possess


the kind of objectivity that likelihoodists require.

25. The Law of Likelihood and the Likelihood Principle have been
formulated in slightly different ways by various logicians and statisticians.
The Law of Likelihood was first identified by that name in Hacking 1965,
and has been invoked more recently by the likelihoodist statisticians
A.F.W. Edwards (1972) and R. Royall (1997). R.A. Fisher (1922) argued
for the Likelihood Principle early in the 20th century, although he didn’t
call it that. One of the first places it is discussed under that name is Savage
et al. 1962.

Copyright © 2021 by the author


James Hawthorne

188 Stanford Encyclopedia of Philosophy

You might also like