<hide main text><show main text>
<show this volume only><show all volumes>
cover
The Collected Papers of Charles Sanders Peirce. Electronic edition.
Volume 3: Exact Logic
Volume 3: Exact Logic. (Previously Published Papers)
Paper 1: On an Improvement in Boole's Calculus of Logic
<hide milestones><show milestones><hide table of contents><show table of contents>

Paper 1: On an Improvement in Boole's Calculus of Logic †1

1. The principal use of Boole's Calculus of Logic lies in its application to problems concerning probability. It consists, essentially, in a system of signs to denote the logical relations of classes. The data of any problem may be expressed by means of these signs, if the letters of the alphabet are allowed to stand for the classes themselves. From such expressions, by means of certain rules for transformation, expressions can be obtained for the classes (of events or things) whose frequency is sought in terms of those whose frequency is known. Lastly, if certain relations are known between the logical relations and arithmetical operations, these expressions for events can be converted into expressions for their probability.

It is proposed, first, to exhibit Boole's system in a modified form, and second, to examine the difference between this form and that given by Boole himself.

2. Let the letters of the alphabet denote classes whether of things or of occurrences. It is obvious that an event may either be singular, as "this sunrise," or general, as "all sunrises." Let the sign of equality with a comma beneath it express numerical identity. Thus a inline image b is to mean that a and b denote the same class—the same collection of individuals.

3. Let a inline image b denote all the individuals contained under a and b together. †2 The operation here performed will differ from arithmetical addition in two respects: first, that it has reference to identity, not to equality; and second, that what is common to a and b is not taken into account twice over, as it would be in arithmetic. The first of these differences, however, amounts to nothing, inasmuch as the sign of identity would indicate the distinction in which it is founded; and therefore we may say that

(1) If No a is b a inline image b inline image a + b. †1

It is plain that

(2) a inline image a inline image a †2

and also, that the process denoted by inline image , and which I shall call the process of logical addition, is both commutative and associative. That is to say

(3) a inline image b inline image b inline image a

and

(4) (a inline image b) inline image c inline image a inline image (b inline image c).

4. Let a, b denote the individuals contained at once under the classes a and b; those of which a and b are the common species. If a and b were independent events, a, b would denote the event whose probability is the product of the probabilities of each. On the strength of this analogy (to speak of no other), the operation indicated by the comma may be called logical multiplication. It is plain that

(5) a, a inline image a. †2

Logical multiplication is evidently a commutative and associative process. That is,

(6) a,b inline image b,a

(7) (a,b),c inline image a,(b,c).

Logical addition and logical multiplication are doubly distributive, so that

(8) (a inline image b),c inline image a,c inline image b,c
and
(9) a,b inline image c inline image (a inline image c),(b inline image c).
Proof. Let a inline image a'+x+y+o
b inline image b'+x+z+o
c inline image c'+y+z+o

where any of these letters may vanish. These formulæ comprehend every possible relation of a, b and c; and it follows from them, that

a inline image b inline image a'+b'+x+y+z+o (a inline image b),c inline image y+z+o.

But

a,c inline image y+o b,c inline image z+o a,c inline image b,c inline image y+z+o ∴ (8).

So

a,b inline image x+o a,b inline image c inline image c'+x+y+z+o.

But

(a inline image c) = a'+c'+x+y+z+o (b inline image c) inline image b'+c'+x+y+z+o
(a inline image c),(b inline image c) inline image c'+x+y+z+o     ∴ (9).

5. Let inline image be the sign of logical subtraction; so defined that

(10) If b inline image x inline image a x inline image a inline image b.

Here it will be observed that x is not completely determinate. It may vary from a to a with b taken away. This minimum may be denoted by a-b. †1 It is also to be observed that if the sphere of b reaches at all beyond a, the expression a inline image b is uninterpretable. †2 If then we denote the contradictory negative of a class by the letter which denotes the class itself, with a line above it, †P1 if we denote by v a wholly indeterminate class, and if we allow (0 inline image 1) to be a wholly uninterpretable symbol, we have

(11) a inline image b inline image v,a,b+a,+[0 inline image 1],ā,b †3

which is uninterpretable unless

ā,b inline image 0.

If we define zero by the following identities, in which x may be any class whatever,

(12) 0 inline image x inline image x inline image x - x

then, zero denotes the class which does not go beyond any class, †1 that is nothing or nonentity.

6. Let a;b be read a logically divided by b, and be defined by the condition that

(13) If b,x inline image a x inline image a;b

x is not fully determined by this condition. It will vary from a to a + and will be uninterpretable if a is not wholly contained under b. †2 Hence, allowing (1;0) to be some uninterpretable symbol,

(14) a;b inline image a,b+v,ā,+(1;0)a, †3

which is uninterpretable unless

a, inline image 0.

7. Unity may be defined by the following identities in which x may be any class whatever.

(15) 1 inline image x;x inline image x:x. †4

Then unity denotes the class of which any class is a part; that is, what is or ens.

8. It is plain that if for the moment we allow a:b to denote the maximum value of a;b, then

(16) inline image 1-x inline image 0:x. †5

So that

(17) x,(1-x) inline image 0 x inline image 0:x inline image 1.

9. The rules for the transformation of expressions involving logical subtraction and division would be very complicated. The following method is, therefore, resorted to. †1

It is plain that any operations consisting solely of logical addition and multiplication, being performed upon interpretable symbols, can result in nothing uninterpretable. Hence, if φ+✕x signifies such an operation performed upon symbols of which x is one, we have

φ+✕x inline image a,x+b,(1-x) †2

where a and b are interpretable.

It is plain, also, that all four operations being performed in any way upon any symbols, will, in general, give a result of which one term is interpretable and another not; although either of these terms may disappear. We have then

φx inline image i,x+j,(1-x). †P1

We have seen that if either of these coefficients i and j is uninterpretable, the other factor of the same term is equal to nothing, or else the whole expression is uninterpretable. But

φ(1) inline image i and φ(0) inline image j.

Hence

(18) φx inline image φ(1),x + φ(0),(1-x)
φ(x and y) inline image φ(1 and 1),x,y + φ(1 and 0),x, + φ(0 and 1),,y + φ(0 and 0),,.
(18') φx inline image (φ(1)inline image),(φ(0) inline image x) †P2
φ(x and y) inline image (φ(1 and 1) inline image inline image ),(φ(1 and 0) inline image , inline image y),
(φ(0 and 1) inline image x inline image ),(φ(0 and 0) inline image x inline image y).

Developing by (18) x inline image y, we have,

x inline image y inline image (1 inline image 1),x,y+(1 inline image 0),x,+(0 inline image 1),,y+(0 inline image 0),,.

So that, by (11),

(19) (1 inline image 1) inline image v. 1 inline image 0 inline image 1. 0 inline image 1 inline image (0 inline image 1). 0 inline image 0 inline image 0.

10. Developing x;y in the same way, we have †P1

x;y inline image 1;1,x,y+1;0,x,+0;1,,y+0;0,,.

So that, by (14),

(20) 1;1 inline image1 1;0 inline image (1;0) 0;1 inline image 0 0;0 inline image v.

Boole gives (20), †1 but not (19).

In solving identities we must remember that

(21) (a inline image b) - b inline image a
(22) (a inline image b) inline image b inline image a.

From a inline image b the value of b cannot be obtained.

(23) (a,b) ÷ b inline image a
(24) a;b,b inline image a.

From a;b the value of b cannot be determined.

11. Given the identity φx inline image 0.

Required to eliminate x.

φ(1) inline image x,φ(1)+(1-x),φ(1)

φ(0) inline image x,φ(0)+(1-x),φ(0).

Logically multiplying these identities, we get

φ(1),φ(0) inline image x,φ(1),φ(0)+(1-x),φ(1),φ(0).

For two terms disappear because of (17).

But we have, by (18),

φ(1),x+φ(0),(1-x) inline image φx inline image 0.

Multiplying logically by x we get

φ(1),x inline image 0

and by (1-x) we get

φ(0),(1-x) inline image 0.

Substituting these values above, we have

(25) φ(1),φ(0) inline image 0 when φx inline image 0.

12. Given φx inline image 1.

Required to eliminate x.

Let φ'x inline image 1-φx inline image 0

φ'(1),φ'(0) inline image (1- φ(1)),(1-φ(0)) inline image 0

1-(1-φ(1)),(1-φ(0)) inline image 1.

Now, developing as in (18), only in reference to φ(1) and φ(0) instead of to x and y,

1-(1-φ(1)),(1-φ(0)) inline image φ(1),φ(0)+φ(1),(1-φ(0)) + φ(0),(1-φ(1)).

But by (18) we have also,

φ(1) inline image φ(0) inline image φ(1),φ(0)+φ(1),(1-φ(0))+φ(0),(1-φ(1)).

So that

(26) φ(1) inline image φ(0) inline image 1 when φx inline image 1.

Boole gives (25), †1 but not (26).

13. We pass now from the consideration of identities to that of equations. †2

Let every expression for a class have a second meaning, which is its meaning in an equation. Namely, let it denote the proportion of individuals of that class to be found among all the individuals examined in the long run.

Then we have

(27) If a inline image b a = b
(28) a + b = (a inline image b) + (a,b).

14. Let ba denote the frequency of b's among the a's. Then considered as a class, if a and b are events, ba denotes the fact that if a happens b happens.

(29) aba = a,b. †3

It will be convenient to set down some obvious and fundamental properties of the function ba.

(30) aba = bab
(31) φ(ba and ca) = (φ(b and c))a
(32) (1-b)a = 1-ba
(33) ba = b/a + b(1-a)(1-1/a)
(34) ab = 1 - (1-a/b) b(1-a)
(35) a)a = (φ(1))a.

The application of the system to probabilities may best be exhibited in a few simple examples, some of which I shall select from Boole's work, in order that the solutions here given may be compared with his.

15. Example 1. Given the proportion of days upon which it hails, and the proportion of days upon which it thunders. Required the proportion of days upon which it does both.

Let 1 inline image days,

p inline image days when it hails,

q inline image days when it thunders,

r inline image days when it hails and thunders.

p,q inline image r

Then by (29), r inline image p,q = p qp=q pq.

Answer. The required proportion is an unknown fraction of the least of the two proportions given.

By p might have been denoted the probability of the major, and by q that of the minor premiss of a hypothetical syllogism of the following form:

If a noise is heard, an explosion always takes place;

If a match is applied to a barrel of gunpowder, a noise is heard;

If a match is applied to a barrel of gunpowder, an explosion always takes place.

In this case, the value given for r would have represented the probability of the conclusion. Now Boole (page 284) solves this problem by his unmodified method, and obtains the following answer:

r = p q+a(1-q)

where a is an arbitrary constant. Here, if q=1 and p=0, r=0. That is, his answer implies that if the major premiss be false and the minor be true, the conclusion must be false. That this is not really so is shown by the above example. Boole (page 286) is forced to the conclusion that "propositions which, when true, are equivalent, are not necessarily equivalent when regarded only as probable." This is absurd, because probability belongs to the events denoted, and not to forms of expression. The probability of an event is not altered by translation from one language to another.

Boole, in fact, puts the problem into equations wrongly (an error which it is the chief purpose of a calculus of logic to prevent), and proceeds as if the problem were as follows:

It being known what would be the probability of Y, if X were to happen, and what would be the probability of Z, if Y were to happen; what would be the probability of Z, if X were to happen?

But even this problem has been wrongly solved by him. For, according to his solution, where

p = YX q = ZY r = ZX,

r must be at least as large as the product of p and q. But if X be the event that a certain man is a negro, Y the event that he is born in Massachusetts, and Z the event that he is a white man, then neither p nor q is zero, and yet r vanishes.

This problem may be rightly solved as follows:

Let p' inline image Y[p] inline image X,Y
q' inline image Z[q] inline image X,Z
r' inline image Z[r] inline image X,Z.
Then, r' inline image p',q';p' inline image p',q';q'.

Developing these expressions by (18) we have

r' inline image p',q'+r'p',q̄'(p',')+r'p̄',q̄'(',')
inline image p',q'+r'p̄',q'(',q')+r'p̄',q̄'(p',q')

The comparison of these two identities shows that

r' inline image p',q'+r'p̄',q̄'(',').
Let V inline image r'p̄',q̄' inline image ((x,,z)/(,y,+))
Now p',q' inline image p'- p',' inline image q'- q','
', inline image ' - p',' inline image ' - q','
And p',' inline image p' - p'q'q' inline image '- p̄''
',q' inline image q' - q'p'p' inline image ' - 'q̄''

Then let

A inline image p'q' inline image (x,y,z)/(y,z)

B inline image 'p̄' inline image (,y,+,,z+x,,z+,,)/(1 - x,y)

C inline image 'q̄' inline image (,y,+,,z+x,,z+,,)/(1 - y,z)

D inline image q'p' inline image (x,y,z)/(x,y)

And we have r = (Y/Z)p+ V(1/Z - q)-(1 + V)((Y/Z)p - A q)

= (Y/Z)p+ V(1/Z - q)-(1 + V)(1/Z - q - B((1-Y p)/Z))
= q+V((1 - Y p)/Z) - (1 + V)((1-Y p)/Z - C((1/Z) - q))
= q+V(1-Y p/Z)-(1 + V)(q - D(Y/Z)p)

16. Example 2. (See Boole, page 276.) Given r and q; to find p.

p inline image r;q inline image r+v,(1-q) because p is interpretable.

Answer. The required proportion lies somewhere between the proportion of days upon which it both hails and thunders, and that added to one minus the proportion of days when it thunders.

17. Example 3. (See Boole, page 279.) Given, out of the number of questions put to two witnesses, and answered by yes or no, the proportion that each answers truly, and the proportion of those their answers to which disagree. Required, out of those wherein they agree, the proportion they answer truly and the proportion they answer falsely. †1

Let 1 inline image the questions put to both witnesses,

p inline image those which the first answers truly,

q inline image those which the second answers truly,

r inline image those wherein they disagree,

w inline image those which both answer truly,

w' inline image those which both answer falsely.

w inline image p,q w' inline image , r inline image p inline image q - w inline image inline image - w'.

Now by (28)

p inline image q = p + q - w inline image = p - p + 1 - q - w'.

Substituting and transposing,

2w = p + q - r 2w' = 2 - p - q - r.
Now w1-r = (w,(1-r))/(1-r) but w(1-r) inline image w.
w'1-r = (w',(1-r))/(1-r) but w'(1-r) inline image w'.
w1-r = (p+q-r)/(2(1-r)) w'1-r = (2-p-q-r)/(2(1-r)).

18. The differences of Boole's system, as given by himself, from the modification of it given here, are three.

First. Boole does not make use of the operations here termed logical addition and subtraction. The advantages obtained by the introduction of them are three, viz., they give unity to the system; they greatly abbreviate the labor of working with it; and they enable us to express particular propositions. This last point requires illustration. Let i be a class only determined to be such that only some one individual of the class a comes under it. Then a inline image i, a is the expression for some a. Boole cannot properly express some a.

Second. Boole uses the ordinary sign of multiplication for logical multiplication. This debars him from converting every logical identity into an equality of probabilities. Before the transformation can be made the equation has to be brought into a particular form, and much labor is wasted in bringing it to that form.

Third. Boole has no such function as ab. This involves him in two difficulties. When the probability of such a function is required, he can only obtain it by a departure from the strictness of his system. And on account of the absence of that symbol, he is led to declare that, without adopting the principle that simple, unconditioned events whose probabilities are given are independent, a calculus of logic applicable to probabilities would be impossible.

19. The question as to the adoption of this principle is certainly not one of words merely. The manner in which it is answered, however, partly determines the sense in which the term "probability" is taken.

In the propriety of language, the probability of a fact either is, or solely depends upon, the strength of the argument in its favor, supposing all relevant relations of all known facts to constitute that argument. Now, the strength of an argument is only the frequency with which such an argument will yield a true conclusion when its premisses are true. Hence probability depends solely upon the relative frequency of a specific event (namely, that a certain kind of argument yields a true conclusion from true premisses) to a generic event (namely, that that kind of argument occurs with true premisses). Thus, when an ordinary man says that it is highly probable that it will rain, he has reference to certain indications of rain—that is, to a certain kind of argument that it will rain—and means to say that there is an argument that it will rain, which is of a kind of which but a small proportion fail. "Probability," in the untechnical sense, is therefore a vague word, inasmuch as it does not indicate what one, of the numerous subordinated and coordinated genera to which every argument belongs, is the one the relative frequency of the truth of which is expressed. It is usually the case, that there is a tacit understanding upon this point, based perhaps on the notion of an infima species of argument. But an infima s pecies is a mere fiction in logic. And very often the reference is to a very wide genus.

The sense in which the term should be made a technical one is that which will best subserve the purposes of the calculus in
question. Now, the only possible use of a calculation of a probability is security in the long run. But there can be no question that an insurance company, for example, which assumed that events were independent without any reason to think that they really were so, would be subjected to great hazard. Suppose, says Mr. Venn, †1 that an insurance company knew that nine tenths of the Englishmen who go to Madeira die, and that nine tenths of the consumptives who go there get well. How should they treat a consumptive Englishman? Mr. Venn has made an error in answering the question, but the illustration puts in a clear light the advantage of ceasing to speak of probability, and of speaking only of the relative frequency of this event to that. †P1


Paper 2: Upon the Logic of Mathematics †1

PART I †2

§1. The Boolean CalculusE

20. The object of the present paper is to show that there are certain general propositions from which the truths of mathematics follow syllogistically, and that these propositions may be taken as definitions of the objects under the consideration of the mathematician without involving any assumption in reference to experience or intuition. That there actually are such objects in experience or pure intuition is not in itself a part of pure mathematics.

21. Let us first turn our attention to the logical calculus of Boole. I have shown in a previous communication to the Academy, †3 that this calculus involves eight operations, viz., Logical Addition, Arithmetical Addition, Logical Multiplication, Arithmetical Multiplication, and the processes inverse to these.

Definitions

1. Identity. a inline image b expresses the two facts that any a is b and any b is a.

2. Logical Addition. a inline image b denotes a member of the class which contains under it all the a's and all the b's, and nothing else.

3. Logical Multiplication. a,b denotes only whatever is both a and b.

4. Zero denotes nothing, or the class without extent, by which we mean that if a is any member of any class, a inline image 0 is a.

5. Unity denotes being, or the class without content, by which we mean that, if a is a member of any class, a is a, 1.

6. Arithmetical Addition. a+b, if a,b inline image 0, is the same as a inline image b, but, if a and b are classes which have any extent in common, it is not a class.

7. Arithmetical Multiplication. ab represents an event when a and b are events only if these events are independent of each other, in which case ab inline image a,b. By the events being independent is meant that it is possible to take two series of terms, A[1], A[2], A[3], etc., and B[1], B[2], B[3], etc., such that the following conditions will be satisfied. (Here x denotes any individual or class, not nothing; A[m], A[n], B[m], B[n], any members of the two series of terms, and ΣA, ΣB, Σ(A,B) logical sums of some of the A[n]'s, the B[n]'s, and the (A[n],B[n])'s respectively.)

Condition 1. No A[m] is A[n].

Condition 2. No B[m] is B[n].

Condition 3. x inline image Σ(A,B)

Condition 4. a inline image ΣA.

Condition 5. b inline image ΣB.

Condition 6. Some A[m] is B[n]. †1

22. From these definitions a series of theorems follow syllogistically, the proofs of most of which are omitted on account of their ease and want of interest.

Theorems

I

23. If a inline image b, then b inline image a.

II

24. If a inline image b, and b inline image c, then a inline image c.

III

25. If a inline image b inline image c, then b inline image a inline image c.

IV

26. If a inline image b inline image m and b inline image c inline image n and a inline image n inline image x, then m inline image c inline image x.

Corollary. These last two theorems hold good also for arithmetical addition.

V

27. If a + b inline image c and a'+b inline image c, then a inline image a', or else there is nothing not b.

This theorem does not hold with logical addition. But from definition 6 it follows that

No a is b (supposing there is any a)

No a' is b (supposing there is any a')

neither of which propositions would be implied in the corresponding formulæ of logical addition. Now from definitions 2 and 6,

Any a is c

∴ Any a is c not b

But again from definitions 2 and 6 we have

Any c not b is a' (if there is any not b)

∴ Any a is a' (if there is any not b)

And in a similar way it could be shown that any a' is a (under the same supposition). Hence by definition 1,

a inline image a' if there is anything not b.

Scholium. In arithmetic this proposition is limited by the supposition that b is finite. †1 The supposition here though similar to that is not quite the same.

VI

28. If a,b inline image c, then b,a inline image c.

VII

29. If a,b inline image m and b,c inline image n and a,n inline image x, then m,c inline image x.

VIII

30. If m,n inline image b and a inline image m inline image u and a inline image n inline image v and a inline image b inline image x, then u,v inline image x.

IX

31. If m inline image n inline image b and a,m inline image u and a,n inline image v and a,b inline image x, then u inline image v inline image x.

The proof of this theorem may be given as an example of the proofs of the rest.

It is required then (by definition 3) to prove three propositions, viz.

First. That any u is x.

Second. That any v is x.

Third. That any x not u is v.

FIRST PROPOSITION

Since u inline image a,m, by definition 3

Any u is m,

and since m inline image n inline image b, by definition 2

Any m is b,

whence

Any u is b,

But since u inline image a,m, by definition 3

Any u is a,

whence

Any u is both a and b,

But since a,b inline image x, by definition 3

Whatever is both a and b is x

whence

Any u is x.

SECOND PROPOSITION

This is proved like the first.

THIRD PROPOSITION

Since a,m inline image u, by definition 3,

Whatever is both a and m is u.
or Whatever is not u is not both a and m.
or Whatever is not u is either not a or not m.
or Whatever is not u and is a is not m.

But since a,b inline image x, by definition 3

Any x is a,
whence Any x not u is not u and is a,
whence Any x not u is not m.

But since a,b inline image x, by definition 3

Any x is b,
whence Any x not u is b,
Any x not u is b, not m.

But since m inline image n inline image b, by definition 2

Any b not m is n,
whence Any x not u is n,
and therefore Any x not u is both a and n. †1

But since a,n inline image v, by definition 3

Whatever is both a and n †2 is v,
whence Any x not u is v.

32. Corollary 1. This proposition readily extends itself to arithmetical addition.

Corollary 2. The converse propositions produced by transposing the last two identities of theorems VIII and IX are also true.

Corollary 3. Theorems VI, VII, and IX hold also with arithmetical multiplication. This is sufficiently evident in the case of theorem VI, because by definition 7 we have an additional premiss, namely, that a and b are independent, and an additional conclusion which is the same as that premiss.

33. In order to show the extension of the other theorems, I shall begin with the following lemma. If a and b are independent, then corresponding to every pair of individuals, one of which is both a and b, there is just one pair of individuals one of which is a and the other b; and conversely, if the pairs of individuals so correspond, a and b are independent. For, suppose a and b independent, then, by definition 7, condition 3, every class (Am,Bn) is an individual. If then Aa denotes any Am which is a, and Bb any Bm which is b, by condition 6 (Aa,Bn) and (Am,Bb) both exist, and by conditions 4 and 5 the former is any individual a, and the latter any individual b. But given this pair of individuals, both of the pair (Aa,Bb) and (Am,Bn) exist by condition 6. But one individual of this pair is both a and b. Hence the pairs correspond, as stated above. Next, suppose a and b to be any two classes. Let the series of Am's be a and not-a; and let the series of Bm's be all individuals separately. Then the first five conditions can always be satisfied. Let us suppose, then, that the sixth alone cannot be satisfied. Then Ap and Bq may be taken such that (Ap,Bq) is nothing. Since Ap and Bq are supposed both to exist, there must be two individuals (Ap,Bn) and (Am,Bq) which exist. But there is no corresponding pair (Am,Bn) and (Ap,Bq). Hence, no case in which the sixth condition cannot be satisfied simultaneously with the first five is a case in which the pairs rightly correspond; or, in other words, every case in which the pairs correspond rightly is a case in which the sixth condition can be satisfied, provided the first five can be satisfied. But the first five can always be satisfied. Hence, if the pairs correspond as stated, the classes are independent.

34. In order to show that theorem VII may be extended to arithmetical multiplication, we have to prove that if a and b, b and c, and a and (b,c), are independent, then (a,b) and c are independent. Let s denote any individual. Corresponding to every s with (a,b,c), there is an a and (b,c). Hence, corresponding to every s with s and with (a,b,c) (which is a particular case of that pair), there is an s with a and with (b,c). But for every s with (b,c) there is a b with c; hence, corresponding to every a with s and with (b,c), there is an a with b and with c. Hence, for every s with s and with (a,b,c) there is an a with b and with c. For every a with b there is an s with (a,b); hence, for every a with b and with c, there is an s with (a,b) and c. Hence, for every s with s and with (a,b,c) there is an s with (a,b) and with c. Hence, for every s with (a,b,c) there is an (a,b) with c. The converse could be proved in the same way. Hence, etc.

35. Theorem IX holds with arithmetical addition of whichever sort the multiplication is. For we have the additional premiss that "No m is n"; whence since "any u is m" and "any v is n," "no u is v," which is the additional conclusion.

Corollary 2, so far as it relates to theorem IX, holds with arithmetical addition and multiplication. For, since no m is n, every pair, one of which is a and either m or n, is either a pair, one of which is a and m, or a pair, one of which is a and n, and is not both. Hence, since for every pair one of which is a and m, there is a pair one of which is a and the other m, and since for every pair one of which is a,n there is a pair one of which is a and the other n; for every pair one of which is a and either m or n, there is either a pair one of which is a and the other m, or a pair one of which is a and the other n, and not both; or, in other words, there is a pair one of which is a and the other either m or n.

(It would perhaps have been better to give this complicated proof in its full syllogistic form. But as my principal object is merely to show that the various theorems could be so proved, and as there can be little doubt that if this is true of those which relate to arithmetical addition it is true also of those which relate to arithmetical multiplication, I have thought the above proof (which is quite apodeictic) to be sufficient. The reader should be careful not to confound a proof which needs itself to be experienced with one which requires experience of the object of proof.)

X

36. If ab inline image c and a'b inline image c, then a inline image a', or no b exists.

This does not hold with logical, but does with arithmetical multiplication.

For if a is not identical with a', it may be divided thus

a inline image a,a'+a,ā'

if ā' denotes not a'. Then

a,b inline image (a,a'),b + (a,ā'),b

and by the definition of independence the last term does not vanish unless (a,ā') inline image 0, or all a is a'; but since a,b inline image a',b inline image (a,a'),b+(ā,a'),b, this term does vanish, and, therefore, only a is a', and in a similar way it could be shown that only a' is a.

XI

37. 1 inline image a inline image 1.

This is not true of arithmetical addition, for since by definition 7,

1x,1 inline image x1

by theorem IX

x,(1+a) inline image x(1+a) inline image x1 + x a inline image x + x a.

Whence x a inline image 0, while neither x nor a is zero, which, as will appear directly, is impossible.

XII

38. 0,a inline image 0.

Proof. For call 0, a inline image x. Then by definition 3

x belongs to the class zero.

∴ by definition 4 x inline image 0.

Corollary 1. The same reasoning applies to arithmetical multiplication.

Corollary 2. From theorem x and the last corollary it follows that if ab inline image 0, either a inline image 0 or b inline image 0.

XIII

39. a,a inline image a. †1

XIV

40. a inline image a inline image a. †1

These do not hold with arithmetical operations.

41. General Scholium. This concludes the theorems relating to the direct operations. As the inverse operations have no peculiar logical interest, they are passed over here.
In order to prevent misapprehension, I will remark that I do not undertake to demonstrate the principles of logic themselves. Indeed, as I have shown in a previous paper, these principles considered as speculative truths are absolutely empty and indistinguishable. †2 But what has been proved is the maxims of logical procedure, a certain system of signs being given.

The definitions given above for the processes which I have termed arithmetical plainly leave the functions of these operations in many cases uninterpreted. Thus if we write

a+b inline image b+a

a+(b+c) inline image (a+b)+c

bc inline image cb

(ab)c inline image a(bc)

a(m+n) inline image a m+a n

we have a series of identities whose truth or falsity is entirely undeterminable. In order, therefore, fully to define those operations, we will say that all propositions, equations, and identities which are in the general case left by the former definitions undetermined as to truth, shall be true, provided they are so in all interpretable cases.

§2. On Arithmetic. †1

42. Equality is a relation of which identity is a species.

If we were to leave equality without further defining it, then by the last scholium all the formal rules of arithmetic would follow from it. And this completes the central design of this paper, as far as arithmetic is concerned.

43. Still it may be well to consider the matter a little further. Imagine, then, a particular case under Boole's calculus, in which the letters are no longer terms of first intention, but terms of second intention, and that of a special kind. Genus, species, difference, property, and accident, are the well-known terms of second intention. These relate particularly to the comprehension †2 of first intentions; that is, they refer to different sorts of predication. Genus and species, however, have at least a secondary reference to the extension †2 of first intentions. Now let the letters, in the particular application of Boole's calculus now supposed, be terms of second intention which relate exclusively to the extension of first intentions. †3 Let the differences of the characters of things and events be disregarded, and let the letters signify only the differences of classes as wider or narrower. In other words, the only logical comprehension which the letters considered as terms will have is the greater or less divisibility of the classes. Thus, n in another case of Boole's calculus might, for example, denote "New England States"; but in the case now supposed, all the characters which make these States what they are being neglected, it would signify only what essentially belongs to a class which has the same relations to higher and lower classes which the class of New England States has, — that is, a collection of six.

44. In this case, the sign of identity will receive a special meaning. For, if m denotes what essentially belongs to a class of the rank of "sides of a cube," then m inline image n will imply, not that every New England State is a side of a cube, and conversely, but that whatever essentially belongs to a class of the numerical rank of "New England States" essentially belongs to a class of the rank of "sides of a cube," and conversely. Identity of this particular sort may be termed equality, and be denoted by the sign =. †P1 Moreover, since the numerical rank of a logical sum depends on the identity or diversity (in first intention) of the integrant parts, and since the numerical rank of a logical product depends on the identity or diversity (in first intention) of parts of the factors, logical addition and multiplication can have no place in this system. Arithmetical addition and multiplication, however, will not be destroyed. ab = c will imply that whatever essentially belongs at once to a class of the rank of a, and to another independent class of the rank of b belongs essentially to a class of the rank of c, and conversely. †1 a + b = c implies that whatever belongs essentially to a class which is the logical sum of two mutually exclusive classes of the ranks of a and b belongs essentially to a class of the rank of c, and conversely. †1 It is plain that from these definitions the same theorems follow as from those given above. Zero and unity will, as before, denote the classes which have respectively no extension and no comprehension; only the comprehension here spoken of is, of course, that comprehension which alone belongs to letters in the system now considered, that is, this or that degree of divisibility; and therefore unity will be what belongs essentially to a class of any rank independent of its divisibility. These two classes alone are common to the two systems, because the first intentions of these alone determine, and are determined by, their second intentions. Finally, the laws of the Boolian calculus, in its ordinary form, are identical with those of this other so far as the latter apply to zero and unity, because every class, in its first intention, is either without any extension (that is, is nothing), or belongs essentially to that rank to which every class belongs, whether divisible or not.

These considerations, together with those advanced [in 1.556], will, I hope, put the relations of logic and arithmetic in a somewhat clearer light than heretofore.



Paper 3: Description of a Notation for the Logic of Relatives, Resulting from an Amplification of the Coneptions of Boole's Calculus of Logic †1

§1. De Morgan's NotationE

45. Relative terms usually receive some slight treatment in works upon logic, but the only considerable investigation into the formal laws which govern them is contained in a valuable paper by Mr. De Morgan in the tenth volume of the Cambridge Philosophical Transactions. †2 He there uses a convenient algebraic notation, which is formed by adding to the well-known spiculæ of that writer the signs used in the following examples.

X . . LY signifies that X is some one of the objects of thought which stand to Y in the relation L, or is one of the L's of Y.

X . LMY signifies that X is not an L of an M of Y.

X . . (L,M)Y signifies that X is either an L or an M of Y.

LM' an L of every M. L[,]M an L of none but M's.

L[-1]Y something to which Y is L. l (small L) non-L.

This system still leaves something to be desired. Moreover, Boole's logical algebra has such singular beauty, so far as it goes, that it is interesting to inquire whether it cannot be extended over the whole realm of formal logic, instead of being restricted to that simplest and least useful part of the subject, the logic of absolute terms, which, when he wrote, was the only formal logic known. The object of this paper is to show that an affirmative answer can be given to this question. I think there can be no doubt that a calculus, or art of drawing inferences, based upon the notation I am to describe, would be perfectly possible and even practically useful in some difficult cases, and particularly in the investigation of logic. I regret that I am not in a situation to be able to perform this labor, but the account here given of the notation itself will afford the ground of a judgment concerning its probable utility.

46. In extending the use of old symbols to new subjects, we must of course be guided by certain principles of analogy, which, when formulated, become new and wider definitions of these symbols. As we are to employ the usual algebraic signs as far as possible, it is proper to begin by laying down definitions of the various algebraic relations and operations. The following will, perhaps, not be objected to.

§2. General Definitions of the Algebraic Signs

47. Inclusion in or being as small as is a transitive relation. The consequence holds that †P1

If xy,
and yz,
then xz.

48. Equality is the conjunction of being as small as and its converse. To say that x = y is to say that xy and yx.

49. Being less than is being as small as with the exclusion of its converse. To say that x < y is to say that xy, and that it is not true that yx.

50. Being greater than is the converse of being less than. To say that x > y is to say that y < x.

51. Addition is an associative operation. That is to say, †P1

(x inline image y) inline image z = x inline image (y inline image z).

Addition is a commutative operation. That is,

x inline image y = y inline image x.

52. Invertible †1 addition is addition the corresponding inverse of which is determinative. The last two formulæ hold good for it, and also the consequence that

If x + y = z,
and x + y' = z,
then y = y'. †2

53. Multiplication is an operation which is doubly distributive with reference to addition. That is,

x(y inline image z) = xy inline image xz,

(x inline image y)z = xz inline image yz.

Multiplication is almost invariably an associative operation. †3

(xy)z = x(yz).

Multiplication is not generally commutative. If we write commutative †4 multiplication with a comma, †5 we have

x,y = y,x.

54. Invertible †1 multiplication is multiplication whose corresponding inverse operation (division) is determinative. We may indicate this by a dot; †2 and then the consequence holds that

If x.y = z,
and x.y'= z,
then y = y'. †3

55. Functional multiplication †4 is the application of an operation to a function. It may be written like ordinary multiplication; but then there will generally be certain points where the associative principle does not hold. Thus, if we write (sin abc) def, there is one such point. If we write (log (base abc) def) ghi, there are two such points. The number of such points depends on the nature of the symbol of operation, and is necessarily finite. If there were many such points, in any case, it would be necessary to adopt a different mode of writing such functions from that now usually employed. We might, for example, give to "log" such a meaning that what followed it up to a certain point indicated by a † should denote the base of the system, what followed that to the point indicated by a ‡ should be the function operated on, and what followed that should be beyond the influence of the sign "log." Thus log abcdefghi would be (log abc) ghi, the base being def. In this paper I shall adopt a notation very similar to this, which will be more conveniently described further on.

56. The operation of involution obeys the formula †P1

(xy)z = x(yz).

Involution, also, follows the indexical principle.

xy inline image z = xy,xz.

Involution, also, satisfies the binomial theorem. †1

(x inline image y)z = xz inline image Σpxz-p,yp inline image yz,

where Σp denotes that p is to have every value less than z, and is to be taken out of z in all possible ways, and that the sum of all the terms so obtained of the form xz-p,yp is to be taken.

57. Subtraction is the operation inverse to addition. We may write indeterminative †2 subtraction with a comma below the usual sign. Then we shall have that

(x inline image y) inline image y = x,

(x - y) + y = x,

(x + y) - y = x.

58. Division is the operation inverse to multiplication. Since multiplication is not generally commutative it is necessary to have two signs for division. I shall take

(x:y)y = x,

x y/x = y.

59. Division inverse to that multiplication which is indicated by a comma may be indicated by a semicolon. So that

(x;y),y = x. †3

60. Evolution and taking the logarithm are the operations inverse to involution.

(xy)x = y,

xlogxy = y.

61. These conditions are to be regarded as imperative. But in addition to them there are certain other characters which it is highly desirable that relations and operations should possess, if the ordinary signs of algebra are to be applied to them. These I will here endeavour to enumerate.

1. It is an additional motive for using a mathematical sign to signify a certain operation or relation that the general conception of this operation or relation should resemble that of the operation or relation usually signified by the same sign. In particular, it will be well that the relation expressed by ⤙ should involve the conception of one member being in the other; addition, that of taking together; multiplication, that of one factor's being taken relatively to the other (as we write 3 X 2 for a triplet of pairs, and Dφ for the derivative of φ); and involution, that of the base being taken for every unit of the exponent.

2. In the second place, it is desirable that, in certain general circumstances, determinate numbers should be capable of being substituted for the letters operated upon, and that when so substituted the equations should hold good when interpreted in accordance with the ordinary definitions of the signs, so that arithmetical algebra should be included under the notation employed as a special case of it. For this end, there ought to be a number known or unknown, which is appropriately substituted in certain cases, for each one of, at least, some class of letters.

3. In the third place, it is almost essential to the applicability of the signs for addition and multiplication, that a zero and a unity should be possible. By a zero I mean a term such that

x inline image 0 = x,

whatever the signification of x; and by a unity a term for which the corresponding general formula

x1 = x

holds good. On the other hand, there ought to be no term a such that ax=x, independently of the value of x.

4. It will also be a strong motive for the adoption of an algebraic notation, if other formulæ which hold good in arithmetic, such as

xz,yz = (x,y)z,

1x = x,

x1 = x,

x0 = 0,

continue to hold good; if, for instance, the conception of a differential is possible, and Taylor's Theorem holds, and inline image †1 or (1+i)1/i plays an important part in the system, if there should be a term having the properties of inline image †1 (3.14159), or properties similar to those of space should otherwise be brought out by the notation, or if there should be an absurd expression having the properties and uses of inline image †1 or the square root of the negative.

§3. Application of the Algebraic Signs to Logic

62. While holding ourselves free to use the signs of algebra in any sense conformable to the above absolute conditions, we shall find it convenient to restrict ourselves to one particular interpretation except where another is indicated. I proceed to describe the special notation which is adopted in this paper.

Use of the Letters

63. The letters of the alphabet will denote logical signs. Now logical terms are of three grand classes. The first embraces those whose logical form involves only the conception of quality, and which therefore represent a thing simply as "a —." These discriminate objects in the most rudimentary way, which does not involve any consciousness of discrimination. They regard an object as it is in itself as such (quale); for example, as horse, tree, or man. These are absolute terms. The second class embraces terms whose logical form involves the conception of relation, and which require the addition of another term to complete the denotation. These discriminate objects with a distinct consciousness of discrimination. They regard an object as over against another, that is as relative; as father of, lover of, or servant of. These are simple relative terms. The third class embraces terms whose logical form involves the conception of bringing things into relation, and which require the addition of more than one term to complete the denotation. They discriminate not only with consciousness of discrimination, but with consciousness of its origin. They regard an object as medium or third between two others, that is as conjugative; as giver of — to —, or buyer of — for — from —. These may be termed conjugative terms. The conjugative term involves the conception of third, the relative that of second or other, the absolute term simply considers an object. †1 No fourth class of terms exists involving the conception of fourth, because when that of third is introduced, since it involves the conception of bringing objects into relation, all higher numbers are given at once, inasmuch as the conception of bringing objects into relation is independent of the number of members of the relationship. †2 Whether this reason for the fact that there is no fourth class of terms fundamentally different from the third is satisfactory or not, the fact itself is made perfectly evident by the study of the logic of relatives. I shall denote absolute terms by the Roman alphabet, a, b, c, d, etc.; relative terms by italics, a, b, c, d, etc.; and conjugative terms by a kind of type called Kennerly, 𝐚, 𝐛, 𝐜, 𝐝, etc.

I shall commonly denote individuals by capitals, and generals †3 by small letters. General symbols for numbers will be printed in black-letter, thus, 𝖆, 𝖇, 𝖈, 𝖉, etc. The Greek letters will denote operations.

64. To avoid repetitions, I give here a catalogue of the letters I shall use in examples in this paper, with the significations I attach to them.

a. animal. p. President of the United States Senate.
b. black. r. rich person.
f. Frenchman. u. violinist.
h. horse. v. Vice-President of the United States.
m. man. w. woman.
a. enemy. h. husband. o. owner.
b. benefactor. l. lover. s. servant.
c. conqueror. m. mother. w. wife.
e. emperor. n. not.
𝗴. giver to—of—. 𝐛. betrayer to—of—.
𝘄. winner over of—to—from—. 𝘁. transferrer from—to—.

Numbers Corresponding to Letters

65. I propose to use the term "universe" to denote that class of individuals about which alone the whole discourse is understood to run. The universe, therefore, in this sense, as in Mr. De Morgan's, †1 is different on different occasions. In this sense, moreover, discourse may run upon something which is not a subjective part of the universe; for instance, upon the qualities or collections of the individuals it contains. †2

I propose to assign to all logical terms, numbers; to an absolute term, the number of individuals it denotes; to a relative term, the average number of things so related to one individual. Thus in a universe of perfect men (men), the number of "tooth of" would be 32. The number of a relative with two correlates would be the average number of things so related to a pair of individuals; and so on for relatives of higher numbers of correlates. I propose to denote the number of a logical term by enclosing the term in square brackets, thus [t].

The Signs of Inclusion, Equality, etc.

66. I shall follow Boole †3 in taking the sign of equality to signify identity. Thus, if v denotes the Vice-President of the United States, and p the President of the Senate of the United States,

v = p

means that every Vice-President of the United States is President of the Senate, and every President of the United States Senate is Vice-President. The sign "less than" is to be so taken that

f < m

means every Frenchman is a man, but there are men besides Frenchmen. Drobisch has used this sign in the same sense. †P1 It will follow from these significations of = and < that the sign ⤙ (or ≦, "as small as") will mean "is." Thus,

f ⤙ m

means "every Frenchman is a man," without saying whether there are any other men or not. So,

ml

will mean that every mother of anything is a lover of the same thing; although this interpretation in some degree anticipates a convention to be made further on. These significations of = and < plainly conform to the indispensable conditions. Upon the transitive character of these relations the syllogism depends, for by virtue of it, from

f ⤙ m
and m ⤙ a,
we can infer that f ⤙ a;

that is, from every Frenchman being a man and every man being an animal, that every Frenchman is an animal. But not only do the significations of = and < here adopted fulfill all absolute requirements, but they have the supererogatory virtue of being very nearly the same as the common significations. Equality is, in fact, nothing but the identity of two numbers; numbers that are equal are those which are predicable of the same collections, just as terms that are identical are those which are predicable of the same classes. †1 So, to write 5 < 7 is to say that 5 is part of 7, just as to write f < m is to say that Frenchmen are part of men. Indeed, if f < m, then the number of Frenchmen is less than the number of men, and if v = p, then the number of Vice-Presidents is equal to the number of Presidents of the Senate; so that the numbers may always be substituted for the terms themselves, in case no signs of operation occur in the equations or inequalities.

The Signs for Addition

67. The sign of addition is taken by Boole, †2 so that

x + y

denotes everything denoted by x, and, besides, everything denoted by y. Thus

m + w

denotes all men, and, besides, all women. This signification for this sign is needed for connecting the notation of logic with that of the theory of probabilities. But if there is anything which is denoted by both the terms of the sum, the latter no longer stands for any logical term on account of its implying that the objects denoted by one term are to be taken besides the objects denoted by the other. For example,

f + u

means all Frenchmen besides all violinists, and, therefore, considered as a logical term, implies that all French violinists are besides themselves. For this reason alone, in a paper which is published in the Proceedings of the Academy for March 17, 1867, †1 I preferred to take as the regular addition of logic a noninvertible process, such that

m inline image b

stands for all men and black things, without any implication that the black things are to be taken besides the men; and the study of the logic of relatives has supplied me with other weighty reasons for the same determination. Since the publication of that paper, I have found that Mr. W. Stanley Jevons, in a tract called Pure Logic, or the Logic of Quality, [1864] †2 had anticipated me in substituting the same operation for Boole's addition, although he rejects Boole's operation entirely and writes the new one with a + sign while withholding from it the name of addition. †P1 It is plain that both the regular non-invertible addition and the invertible addition satisfy the absolute conditions. But the notation has other recommendations. The conception of taking together involved in these processes is strongly analogous to that of summation, the sum of 2 and 5, for example, being the number of a collection which consists of a collection of two and a collection of five. Any logical equation or inequality in which no operation but addition is involved may be converted into a numerical equation or inequality by substituting the numbers of the several terms for the terms themselves—provided all the terms summed are mutually exclusive. Addition being taken in this sense, nothing is to be denoted by zero, for then

x inline image 0 = x,

whatever is denoted by x; and this is the definition of zero. †1 This interpretation is given by Boole, and is very neat, on account of the resemblance between the ordinary conception of zero and that of nothing, and because we shall thus have

[0] = 0

The Signs for Multiplication

68. I shall adopt for the conception of multiplication the application of a relation, in such a way that, for example, lw shall denote whatever is lover of a woman. This notation is the same as that used by Mr. De Morgan, although he appears not to have had multiplication in his mind. s(m inline image w) will, then, denote whatever is servant of anything of the class composed of men and women taken together. So that

s(m inline image w) = sm inline image sw.

(l inline image s)w will denote whatever is lover or servant to a woman, and

(l inline image s)w = lw inline image sw.

(sl)w will denote whatever stands to a woman in the relation of servant of a lover, and

(sl)w = s(lw).

Thus all the absolute conditions of multiplication are satisfied.

The term "identical with—" is a unity for this multiplication. That is to say, if we denote "identical with—" by 1 we have

x1 = x,

whatever relative term x may be. For what is a lover of something identical with anything, is the same as a lover of that thing.

69. A conjugative term like giver naturally requires two correlates, one denoting the thing given, the other the recipient of the gift. We must be able to distinguish, in our notation, the giver of A to B from the giver to A of B, and, therefore, I suppose the signification of the letter equivalent to such a relative to distinguish the correlates as first, second, third, etc., so that "giver of — to —" and "giver to — of —" will be expressed by different letters. Let g denote the latter of these conjugative terms. Then, the correlates or multiplicands of this multiplier cannot all stand directly after it, as is usual in multiplication, but may be ranged after it in regular order, so that

gxy

will denote a giver to x of y. But according to the notation, x here multiplies y, so that if we put for x owner (o), and for y horse (h),

goh

appears to denote the giver of a horse to an owner of a horse. But let the individual horses be H, H', H'', etc. Then

h †1 = H inline image H' inline image H'' inline image etc.

goh = go(H inline image H' inline image H'' inline image etc.) = goH inline image goH' inline image goH'' inline image etc.

Now this last member must be interpreted as a giver of a horse to the owner of that horse, and this, therefore, must be the interpretation of goh. This is always very important. A term multiplied by two relatives shows that the same individual is in the two relations. If we attempt to express the giver of a horse to a lover of a woman, and for that purpose write

glwh,

we have written giver of a woman to a lover of her, and if we add brackets, thus,

g(lw)h,

we abandon the associative principle of multiplication. A little reflection will show that the associative principle must in some form or other be abandoned at this point. But while this principle is sometimes falsified, it oftener holds, and a notation must be adopted which will show of itself when it holds. We already see that we cannot express multiplication by writing the multiplicand directly after the multiplier; let us then affix subjacent numbers after letters to show where their correlates are to be found. The first number shall denote how many factors must be counted from left to right to reach the first correlate, the second how many more must be counted to reach the second, and so on. Then, the giver of a horse to a lover of a woman may be written

g12l1wh = g11l2hw = g2-1hl1w.

70. Of course a negative number indicates that the former correlate follows the latter by the corresponding positive number. A subjacent zero makes the term itself the correlate. Thus,

l0

denotes the lover of that lover or the lover of himself, just as goh denotes that the horse is given to the owner of itself, for to make a term doubly a correlate is, by the distributive principle, to make each individual doubly a correlate, so that

l0 = L0 inline image L0' inline image L0'' inline image etc.

A subjacent sign of infinity may indicate that the correlate is indeterminate, so that

l

will denote a lover of something. We shall have some confirmation of this presently. †1

If the last subjacent number is a one it may be omitted. Thus we shall have

l1 = l,

g11 = g1 = g.

This enables us to retain our former expressions lw, goh, etc.

71. The associative principle does not hold in this counting of factors. Because it does not hold, these subjacent numbers are frequently inconvenient in practice, and I therefore use also another mode of showing where the correlate of a term is to be found. This is by means of the marks of reference, † ‡ || § ¶, which are placed subjacent to the relative term and before and above the correlate. Thus, giver of a horse to a lover of a woman may be written

g†‡l||||w‡h.

The asterisk I use exclusively to refer to the last correlate of the last relative of the algebraic term.

72. Now, considering the order of multiplication to be: — a term, a correlate of it, a correlate of that correlate, etc., — there is no violation of the associative principle. The only violations of it in this mode of notation are that in thus passing from relative to correlate, we skip about among the factors in an irregular manner, and that we cannot substitute in such an expression as goh a single letter for oh. I would suggest that such a notation may be found useful in treating other cases of non-associative multiplication. By comparing this with what was said above †1 concerning functional multiplication, it appears that multiplication by a conjugative term is functional, and that the letter denoting such a term is a symbol of operation. I am therefore using two alphabets, the Greek and Kennerly, where only one was necessary. But it is convenient to use both.

73. Thus far, we have considered the multiplication of relative terms only. Since our conception of multiplication is the application of a relation, we can only multiply absolute terms by considering them as relatives. Now the absolute term "man" is really exactly equivalent to the relative term "man that is —," and so with any other. I shall write a comma after any absolute term to show that it is so regarded as a relative term. Then man that is black will be written

m,b.

But not only may any absolute term be thus regarded as a relative term, but any relative term may in the same way be regarded as a relative with one correlate more. It is convenient to take this additional correlate as the first one. Then

l,sw †2

will denote a lover of a woman that is a servant of that woman. The comma here after l should not be considered as altering at all the meaning of l, but as only a subjacent sign, serving to alter the arrangement of the correlates. In point of fact, since a comma may be added in this way to any relative term, it may be added to one of these very relatives formed by a comma, and thus by the addition of two commas an absolute term becomes a relative of two correlates. So

m,,b,r,

interpreted like goh,

means a man that is a rich individual and is a black that is that rich individual. But this has no other meaning than

m,b,r ,

or a man that is a black that is rich. Thus we see that, after one comma is added, the addition of another does not change the meaning at all, so that whatever has one comma after it must be regarded as having an infinite number. If, therefore, l,,sw is not the same as l,sw (as it plainly is not, because the latter means a lover and servant of a woman, and the former a lover of and servant of and same as a woman), this is simply because the writing of the comma alters the arrangement of the correlates. And if we are to suppose that absolute terms are multipliers at all (as mathematical generality demands that we should), we must regard every term as being a relative requiring an infinite number of correlates to its virtual infinite series "that is—and is—and is—etc." Now a relative formed by a comma of course receives its subjacent numbers like any relative, but the question is, What are to be the implied subjacent numbers for these implied correlates? Any term may be regarded as having an infinite number of factors, those at the end being ones, thus,

l,sw = l,sw,1,1,1,1,1,1,1, etc.

A subjacent number may therefore be as great as we please. But all these ones denote the same identical individual denoted by w; what then can be the subjacent numbers to be applied to s, for instance, on account of its infinite "that is" 's? What numbers can separate it from being identical with w? There are only two. The first is zero, which plainly neutralizes a comma completely, since

s,0w = sw, †1

and the other is infinity; for as 1 is indeterminate in ordinary algebra, so it will be shown hereafter to be here, so that to remove the correlate by the product of an infinite series of ones is to leave it indeterminate. Accordingly,

m,

should be regarded as expressing some man. Any term, then, is properly to be regarded as having an infinite number of commas, all or some of which are neutralized by zeros.

"Something" may then be expressed by

1. †1

I shall for brevity frequently express this by an antique figure one (𝟏).

"Anything" by

10. †2

I shall often also write a straight 1 for anything.

74. It is obvious that multiplication into a multiplicand indicated by a comma is commutative, †P1 that is,

s,l = l,s.

This multiplication is effectively the same as that of Boole in his logical calculus. Boole's unity is my 1, that is, it denotes whatever is.

75. The sum x + x generally denotes no logical term. But x, + x, may be considered as denoting some two x's. It is natural to write

x+x = 2.x,
and x,+x, = 2.x,,

where the dot shows that this multiplication is invertible. We may also use the antique figures so that

2.x, = 𝟐x,
just as 1 = 𝟏.

Then 𝟐 alone will denote some two things. But this multiplication is not in general commutative, and only becomes so when it affects a relative which imparts a relation such that a thing only bears it to one thing, and one thing alone bears it to a thing. For instance, the lovers of two women are not the same as two lovers of women, that is,

l2.w and 2.lw

are unequal; but the husbands of two women are the same as two husbands of women, that is,

h2.w = 2.hw,
and in general, x,2.y = 2.x,y.

76. The conception of multiplication we have adopted is that of the application of one relation to another. So, a quaternion being the relation of one vector to another, the multiplication of quaternions is the application of one such relation to a second. Even ordinary numerical multiplication involves the same idea, for 2 X 3 is a pair of triplets, and 3 X 2 is a triplet of pairs, where "triplet of" and "pair of" are evidently relatives.

If we have an equation of the form

xy = z,

and there are just as many x's per y as there are per things, things of the universe, then we have also the arithmetical equation,

[x][y] = [z].

For instance, if our universe is perfect men, and there are as many teeth to a Frenchman (perfect understood) as there are to any one of the universe, then

[t][f] = [t f]

holds arithmetically. So if men are just as apt to be black as things in general,

[m,][b] = [m,b],

where the difference between [m] and [m,] must not be overlooked. It is to be observed that

[1] = 𝟏.

Boole was the first to show this connection between logic and probabilities. †1 He was restricted, however, to absolute terms. I do not remember having seen any extension of probability to relatives, except the ordinary theory of expectation.

Our logical multiplication, then, satisfies the essential conditions of multiplication, has a unity, has a conception similar to that of admitted multiplications, and contains numerical multiplication as a case under it.

The Sign of Involution

77. I shall take involution in such a sense that xy will denote everything which is an x for every individual of y. Thus lw will be a lover of every woman. Then (sl)w will denote whatever stands to every woman in the relation of servant of every lover of hers; and s(lw) will denote whatever is a servant of everything that is lover of a woman. So that

(sl)w = s(lw).

A servant of every man and woman will be denoted by sm inline image w †1, and sm, sw will denote a servant of every man that is a servant of every woman. So that

sm inline image w = sm,sw.

That which is emperor or conqueror of every Frenchman will be denoted by (e inline image c)f, and ef inline image Σpef-p, cp inline image cf will denote whatever is emperor of every Frenchman or emperor of some Frenchmen and conqueror of all the rest, or conqueror of every Frenchman. Consequently,

(e inline image c)f = ef inline image Σpef-p,cp inline image cf.

Indeed, we may write the binomial theorem so as to preserve all its usual coefficients; for we have

(e inline image c)f = efinline image[f].ef-†1,c𝟏†inline image(([f].([f]-𝟏))/𝟐).ef-‡𝟐,c𝟐‡inline image etc. †2

That is to say, those things each of which is emperor or conqueror of every Frenchman consist, first, of all those individuals each of which is a conqueror [emperor!] of every Frenchman; second, of a number of classes equal to the number of Frenchmen, each class consisting of everything which is an emperor of every Frenchman but some one and is a conqueror of that one; third, of a number of classes equal to half the product of the number of Frenchmen by one less than that number, each of these classes consisting of every individual which is an emperor of every Frenchman except a certain two, and is conqueror of those two, etc. This theorem holds, also, equally well with invertible addition, and either term of the binomial may be negative provided we assume

(—x)y = (—)[y].xy.

In addition to the above equations which are required to hold good by the definition of involution, the following also holds,

(s,l)w = sw,lw, †1

just as it does in arithmetic.

78. The application of involution to conjugative terms presents little difficulty after the explanations which have been given under the head of multiplication. It is obvious that betrayer to every enemy should be written

𝐛a,

just as lover of every woman is written

lw,

but 𝐛 = 𝐛11 and therefore, in counting forward as the subjacent numbers direct, we should count the exponents, as well as the factors, of the letter to which the subjacent numbers are attached. Then we shall have, in the case of a relative of two correlates, six different ways of affixing the correlates to it, thus:

𝐛am betrayer of a man to an †P1 enemy of him;
(𝐛a)m betrayer of every man to some enemy of him;
𝐛am betrayer of each man to an enemy of every man;
𝐛am betrayer of a †P1 man to all †P1 enemies of all men;
𝐛am betrayer of a man to every enemy of him;
𝐛am betrayer of every man to every enemy of him.

If both correlates are absolute terms, the cases are

𝐛mw betrayer of a woman to a man;
(𝐛m)w betrayer of each woman to some man;
𝐛mw betrayer of all women to a man;
𝐛mw betrayer of a woman to every man; †P1
𝐛mw betrayer of a woman to all men;
𝐛mw betrayer of every woman to every man.

These interpretations are by no means obvious, but I shall show that they are correct further on. †1

79. It will be perceived that the rule still holds here that

(𝐛a)m = 𝐛(am)

that is to say, that those individuals each of which stand to every man in the relation of betrayer to every enemy of his are identical with those individuals each of which is a betrayer to every enemy of a man of that man.

80. If the proportion of lovers of each woman among lovers of other women is equal to the average number of lovers which single individuals of the whole universe have, then

[lw] = [lW'] [lW''] [lW'''] etc.=[l][w].

Thus arithmetical involution appears as a special case of logical involution.

§4. General Formulæ

81. The formulæ which we have thus far obtained, exclusive of mere explanations of signs and of formulæ relating to the numbers of classes, are:

(1) If xy and yz, then xz.
(2) (x inline image y) inline image z = x inline image (y inline image z). (Jevons)
(3) x inline image y = y inline image x. (Jevons)
(4) (x inline image y)z = xz inline image yz.
(5) x(y inline image z) = xy inline image xz.
(6) (xy)z = x(yz).
(7) x,(y inline image z) = x,y inline image x,z. (Jevons)
(8) (x,y),z = x,(y,z). (Boole)
(9) x,y, = y,x. (Boole)
(10) (xy)z = x(yz).
(11) xy inline image z = xy,xz.
(12) (x inline image y)z = xz inline image Σp(xx-p,yp) inline image yp
= xzinline image[z].xz-†𝟏,y†𝟏 inline image (([z].[z-𝟏])/𝟐).xz-‡𝟐,y‡𝟐
inline image (([z].[z-𝟏].[z-𝟐])/(𝟐.𝟑)).xz-||𝟑,y||𝟑 inline image etc.
(13) (x,y)z = xz,yz.
(14) x + 0 = x. (Boole)
(15) x1 = x.
(16) (x + y) + z = x + (y + z). (Boole)
(17) x + y = y + x. (Boole)
(18) x + y - y = x. (Boole)
(19) x,(y + z) = x,y + x,z. (Boole)
(20) (x + y)z = x + [z].xz-†1,y†𝟏 + etc.

We have also the following, which are involved implicitly in the explanations which have been given.

(21) xx inline image y. †1

This, I suppose, is the principle of identity, for it follows from this that x = x. †2

(22) x inline image x = x. (Jevons)
(23) x,x = x. (Boole)
(24) x inline image y = x + y - x,y.

The principle of contradiction is

(25) x,nx = 0

where n stands for "not." The principle of excluded middle is

(26) x inline image nx = 1.

It is an identical proposition, that, if φ be determinative, we have

(27) If x = y φx = φy.

The six following are derivable from the formulæ already given:

(28) (x inline image y),(x inline image z) = x inline image y,z.
(29) (x - y) inline image (z - w) = (x inline image z)-(y inline image w) + y,z,(1-w) + x,(1 - y),w.

In the following, φ is a function involving only the commutative operations and the operations inverse to them.

(30) φx = (φ1),x + (φ0),(1 - x). (Boole)
(31) φx = (φ1inline image(1-x)),(φ0inline imagex).
(32) If φx = 0 (φ1),(φ0) = 0. (Boole)
(33) If φx = 1 φ1 inline image φ0 = 1.

The reader may wish information concerning the proofs of formulæ (30) to (33). When involution is not involved in a function nor any multiplication except that for which x,x=x, it is plain that φx is of the first degree, and therefore, since all the rules of ordinary algebra hold, we have as in that

φx = φ0 + (φ1 - φ0),x.

We shall find, hereafter, that when y has a still more general character, we have,

φx = φ0 + (φ1 - φ0)x.

The former of these equations by a simple transformation gives (30).

If we regard (φ1), (φ0) as a function of x and develop it by (30), we have

(φ1),(φ0) = x,(φ1),(φ0) + (φ1),(φ0),(1-x).

Comparing these terms separately with the terms of the second member of (30), we see that

(φ1),(φ0) ⤙ x.

This gives at once (32), and it gives (31) after performing the multiplication indicated in the second member of that equation and equating φx to its value as given in (30). If (φ1 inline image φ0) is developed as a function of x by (31), and the factors of the second member are compared with those of the second member of (31), we get

φx ⤙ φ1 inline image φ0,

from which (33) follows immediately.

Properties of Zero and Unity

82. The symbolical definition of zero is

x+0 = x,

so that by (19) x,a = x,(a+0) = x,a+x,0.

Hence, from the invertible character of this addition, and the generality of (14), we have

x,0 = 0.

By (24) we have in general,

x inline image 0 = x + 0 - x,0 = x,
or x inline image 0 = x.
By (4) we have a x = (a inline image 0)x = a x inline image 0x.
But if a is an absurd relation, a x = 0,
so that 0x = 0,
which must hold invariably.

From (12) we have ax = (a inline image 0)x = ax inline image 0x inline image etc.,
whence by (21) 0xax. †1

But if a is an absurd relation, and x is not zero,

ax = 0.

And therefore, unless x=0, 0x = 0.

83. Any relative x may be conceived as a sum of relatives X, X', X'', etc., such that there is but one individual to which anything is X, but one to which anything is X', etc. Thus, if x denote "cause of," X,X',X'' would denote different kinds of causes, the causes being divided according to the differences of the things they are causes of. Then we have

X y = X(y inline image 0) = X y inline image X0,

whatever y may be. Hence, since y may be taken so that

X y = 0,
we have X0 = 0;

and in a similar way,

X'0 = 0, X''0 = 0, X'''0 = 0, etc.

We have, then,

x0 = (X inline image X' inline image X'' inline image X''' inline image etc.)0
= X0 inline image X'0 inline image X''0 inline image X'''0 inline image etc. = 0.

84. If the relative x be divided in this way into X,X',X'', X''', etc., so that x is that which is either X or X' or X'' or X''', etc., then non-x is that which is at once non-X and non-X' and non-X'', etc.; that is to say,

non-x = non-X, non-X', non-X'', non-X''', etc.;

where non-X is such that there is something (Z) such that everything is non-X to Z; and so with non-X', non-X'', etc. Now, non-x may be any relative whatever. Substitute for it, then, y; and for non-X, non-X', etc., Y,Y', etc. Then we have

y = Y,Y',Y'',Y''', etc.;
and Y'Z' = 1, Y''Z'' = 1, Y'''Z''' = 1, etc.,

where Z',Z'',Z''' are individual terms which depend for what they denote on Y',Y'',Y'''.

Then we have

1 = Y'Z' = Y'Z' = Y'(Z' inline image 0) = Y'Z',Y'0 = Y'Z',Y'0,
or Y'0 = 1, Y''0 = 1, Y'''0= 1, etc.
Then y0 = (Y', Y'', Y''', etc.)0 = Y'0, Y''0, Y'''0, etc. = 1.
We have by definition, x1 = x.
Hence, by (6), a x = (a1)x = a(1x).

Now a may express any relation whatever, but things the same way related to everything are the same. Hence,

x = 1x.
We have by definition, 1 = 1[0].
Then if X is any individual X,1 = X,1[0] = X,1X.
But 1X = X.
Hence X,1 = X,X;
and by (23) X,1 = X;
whence if we take x = X + X' + X'' + X''' + etc.,

where X,X' etc, denote individuals (and by the very meaning of a general term this can always be done, whatever x may be)

x,1 = (X + X' + X'' + etc.),1 = X,1 + X',1 + X'',1 + etc.
= X + X' + X'' + etc. = x,
or x,1 = x.
We have by (24) x inline image 1 = x + 1 - x,1 = x + 1 - x = 1,
or x inline image 1 = 1.

85. We may divide all relatives into limited and unlimited. Limited relatives express such relations as nothing has to everything. For example, nothing is knower of everything. Unlimited relatives express relations such as something has to everything. For example, something is as good as anything. For limited relatives, then, we may write

p1 = 0.

The converse of an unlimited relative expresses a relation which everything has to something. Thus, everything is as bad as something. Denoting such a relative by q,

q1 = 1.

These formulæ remind one a little of the logical algebra of Boole; because one of them holds good in arithmetic only for zero, and the other only for unity.

We have by (10) 1x = (q0)x = q(0x) = q0 = 1,
or 1x = 1.
We have by (4) 1x = (a inline image 1)x = a x inline image 1x,
or by (21) a x ⤙ 1x.

But everything is somehow related to x unless x is 0; hence unless x is 0,

1x = 1.

If a denotes "what possesses," and y "character of what is denoted by x,"

x = ay = a(y1) = (ay)1 = x1,
or x1 = x.

Since 1 means "identical with," l,1w denotes whatever is both a lover of and identical with a woman, or a woman who is a lover of herself. And thus, in general,

x,1 = x[0],.

86. Nothing is identical with every one of a class; and therefore 1x is zero, unless x denotes only an individual when 1x becomes equal to x. But equations founded on interpretation may not hold in cases in which the symbols have no rational interpretation.

Collecting together all the formulæ relating to zero and
unity, we have

(34) x inline image 0 = x. (Jevons)
(35) x inline image 1 = 1. (Jevons)
(36) x0 = 0.
(37) 0x = 0.
(38) x,0 = 0. (Boole)
(39) x0 = 1.
(40) 0x = 0, provided x > 0. †1
(41) 1x = x.
(42) x,1 = x0,.
(43) x1 = x.
(44) 1x = 0, unless x is individual, when 1x = x.
(45) q1 = 1, where q is the converse of an unlimited relative.
(46) 1x = 1, provided x > 0. †1
(47) x,1 = x. (Boole)
(48) p1 = 0, where p is a limited relative.
(49) 1 x = 1.

These, again, give us the following:

(50) 0 inline image 1 = 1 (64) 01=0
(51) 0 inline image 1 = 1 (65) 1 1=1
(52) 00 = 0 (66) 1,1=1
(53) 0,0 = 0 (67) 11=1
(54) 00 = 1 (68) 11 = 1
(55) 10 = 0 (69) 1,1 = 1
(56) 01 = 0 (70) 11 = 1
(57) 0,1 = 0 (71) 11=1
(58) 01 = 0 (72) 11 = 1
(59) 10 = 1 (73) 1,1 = 1
(60) 01 = 0 (74) 11 = 1
(61) 10 = 0 (75) 11 = 0
(62) 0,1 = 0 (76) 1, = 1
(63) 10 = 1

From (64) we may infer that 0 is a limited relative, and from (60) that it is not the converse of an unlimited relative. From (70) we may infer that 1 is not a limited relative, and from (68) that it is the converse of an unlimited relative.

Formulæ Relating to the Numbers of Terms

87. We have already seen that

(77) If x ⤙ y, then [x] ⤙ [y].
(78) When x,y = 0, then [x inline image y] = [x] inline image [y],
(79) When [xy]:[nxy] = [x]:[nx], then [xy] = [x][y].
(80) When [x𝖓y] = [x][𝖓y][1], then [x y] = [x][y].

It will be observed that the conditions which the terms must conform to, in order that the arithmetical equations shall hold, increase in complexity as we pass from the more simple relations and processes to the more complex.

88. We have seen that

(81) [0] = 0.
(82) [1] = 1.

Most commonly the universe is unlimited, and then

(83) [1] = ∞; †1

and the general properties of 1 correspond with those of infinity. Thus,

x inline image 1 = 1 corresponds to x + ∞ = ∞,
q1 = 1 corresponds to q ∞ = ∞,
1x = 1 corresponds to ∞ x= ∞,
p1 = 0 corresponds to p ∞ = 0,
1x = 1 corresponds to x = ∞.

The formulæ involving commutative multiplication are derived from the equation 1, = 1. But if 1 be regarded as infinite, it is not an absolute infinite; for 10 = 0. On the other hand, 11 = 0.

It is evident, from the definition of the number of a term, that

(84) [x,] = [x]:[1].

We have, therefore, if the probability of an individual being x to any y is independent of what other y's it is x to, and if x is independent of y,

(85) [xy,] = [x,][y].

§5. General Method of Working with This Notation

89. Boole's logical algebra contains no operations except our invertible addition and commutative multiplication, together With the corresponding subtraction and division. He has, therefore, only to expand expressions involving division, by means of (30), so as to free himself from all non-determinative operations, in order to be able to use the ordinary methods of algebra, which are, moreover, greatly simplified by the fact that

x,x = x.

90. Mr. Jevons's modification †1 of Boole's algebra involves only non-invertible addition and commutative multiplication, without the corresponding inverse operations. He is enabled to replace subtraction by multiplication, owing to the principle of contradiction, and to replace division by addition, owing to the principle of excluded middle. For example, if x be unknown, and we have

x inline image m = a,

or what is denoted by x together with men make up animals, we can only conclude, with reference to x, that it denotes (among other things, perhaps) all animals not men; that is, that the x's not men are the same as the animals not men. Let m̅ denote non-men; then by multiplication we have

xm̅, inline image m,m̅ = x,m̅ = a,m̅,

because, by the principle of contradiction,

m,m̅ = 0.

Or, suppose, x being again unknown, we have given

a,x = m.

Then all that we can conclude is that the x's consist of all the m's and perhaps some or all of the non-a's, or that the x's and non-a's together make up the m's and non-a's together. If, then, ā denote non-a, add ā to both sides and we have

a,x inline image ā = m inline image ā.
Then by (28) (a inline image ā),(x inline image ā) = m inline image ā.

But by the principle of excluded middle,

a inline image ā = 1
and therefore x inline image ā = m inline image ā.

I am not aware that Mr. Jevons actually uses this latter process, but it is open to him to do so. In this way, Mr. Jevons's algebra becomes decidedly simpler even than Boole's.

It is obvious that any algebra for the logic of relatives must be far more complicated. In that which I propose, we labor under the disadvantages that the multiplication is not generally commutative, that the inverse operations are usually indeterminative, and that transcendental equations, and even equations like

abx = cdex + fx + x,

where the exponents are three or four deep, are exceedingly common. It is obvious, therefore, that this algebra is much less manageable than ordinary arithmetical algebra.

91. We may make considerable use of the general formulæ already given, especially of (1), (21), and (27), and also of the following, which are derived from them:

(86) If ab then there is such a term x that a inline image x = b.
(87) If ab then there is such a term x that b,x = a.
(88) If b,x = a then ab.
(89) If ab c inline image ac inline image b.
(90) If ab cacb.
(91) If ab acbc.
(92) If ab cbca †1
(93) If ab acbc.
(94) a,ba

There are, however, very many cases in which the formulæ thus far given are of little avail.

92. Demonstration of the sort called mathematical is founded on suppositions of particular cases. The geometrician draws a figure; the algebraist assumes a letter to signify a single quantity fulfilling the required conditions. But while the mathematician supposes an individual case, his hypothesis is yet perfectly general, because he considers no characters of the individual case but those which must belong to every such case. The advantage of his procedure lies in the fact that the logical laws of individual terms are simpler than those which relate to general terms, because individuals are either identical or mutually exclusive, and cannot intersect or be subordinated to one another as classes can. Mathematical demonstration is not, therefore, more restricted to matters of intuition than any other kind of reasoning. Indeed, logical algebra conclusively proves that mathematics extends over the whole realm of formal logic; and any theory of cognition which cannot be adjusted to this fact must be abandoned. We may reap all the advantages which the mathematician is supposed to derive from intuition by simply making general suppositions of individual cases.

93. In reference to the doctrine of individuals, †1 two distinctions should be borne in mind. The logical atom, or term not capable of logical division, must be one of which every predicate may be universally affirmed or denied. For, let A be such a term. Then, if it is neither true that all A is X nor that no A is X, it must be true that some A is X and some A is not X; and therefore A may be divided into A that is X and A that is not X, which is contrary to its nature as a logical atom. Such a term can be realized neither in thought nor in sense. Not in sense, because our organs of sense are special — the eye, for example, not immediately informing us of taste, so that an image on the retina is indeterminate in respect to sweetness and non-sweetness. When I see a thing, I do not see that it is not sweet, nor do I see that it is sweet; and therefore what I see is capable of logical division into the sweet and the not sweet. It is customary to assume that visual images are absolutely determinate in respect to color, but even this may be doubted. I know no facts which prove that there is never the least vagueness in the immediate sensation. In thought, an absolutely determinate term cannot be realized, because, not being given by sense, such a concept would have to be formed by synthesis, and there would be no end to the synthesis because there is no limit to the number of possible predicates. A logical atom, then, like a point in space, would involve for its precise determination an endless process. We can only say, in a general way, that a term, however determinate, may be made more determinate still, but not that it can be made absolutely determinate. Such a term as "the second Philip of Macedon" is still capable of logical division—into Philip drunk and Philip sober, for example; but we call it individual because that which is denoted by it is in only one place at one time. It is a term not absolutely indivisible, but indivisible as long as we neglect differences of time and the differences which accompany them. Such differences we habitually disregard in the logical division of substances. In the division of relations, etc., we do not, of course, disregard these differences, but we disregard some others. There is nothing to prevent almost any sort of difference from being conventionally neglected in some discourse, and if I be a term which in consequence of such neglect becomes indivisible in that discourse, we have in that discourse,

[I] = 1.

This distinction between the absolutely indivisible and that which is one in number from a particular point of view is shadowed forth in the two words individual {to atomon} and singular {to kath' hekaston}; but as those who have used the word individual have not been aware that absolute individuality is merely ideal, it has come to be used in a more general sense. †P1

94. The old logics distinguish between individuum signatum and individuum vagum. "Julius Cæsar" is an example of the former; "a certain man," of the latter. The individuum vagum, in the days when such conceptions were exactly investigated, occasioned great difficulty from its having a certain generality, being capable, apparently, of logical division. If we include under the individuum vagum such a term as "any individual man," these difficulties appear in a strong light, for what is true of any individual man is true of all men. Such a term is in one sense not an individual term; for it represents every man. But it represents each man as capable of being denoted by a term Which is individual; and so, though it is not itself an individual term, it stands for any one of a class of individual terms. If we call a thought about a thing in so far as it is denoted by a term, a second intention, we may say that such a term as "any individual man" is individual by second intention. The letters which the mathematician uses (whether in algebra or in geometry) are such individuals by second intention. Such individuals are one in number, for any individual man is one man; they may also be regarded as incapable of logical division, for any individual man, though he may either be a Frenchman or not, is yet altogether a Frenchman or altogether not, and not some one and some the other. Thus, all the formal logical laws relating to individuals will hold good of such individuals by second intention, and at the same time a universal proposition may at any moment be substituted for a proposition about such an individual, for nothing can be predicated of such an individual which cannot be predicated of the whole class.

95. There are in the logic of relatives three kinds of terms which involve general suppositions of individual cases. The first are individual terms, which denote only individuals †1; the second are those relatives whose correlatives are individual: I term these infinitesimal relatives †2; the third are individual infinitesimal relatives, and these I term elementary relatives. †3

Individual Terms

96. The fundamental formulæ relating to individuality are two. Individuals are denoted by capitals.

(95) If x > 0 x = X inline image X' inline image X'' inline image X''' inline image etc.
(96) yX = yX.

We have also the following which are easily deducible from these two:

(97) (y,z)X = (y X),(z X). (99) [X] = 𝟏.
(98) X,y0 = X,y X. (100) 1X = X.

We have already seen that

1x = 0, provided that [x] > 𝟏.

97. As an example of the use of the formulæ we have thus far obtained, let us investigate the logical relations between "benefactor of a lover of every servant of every woman," "that which stands to every servant of some woman in the relation of benefactor of a lover of him," "benefactor of every lover of some servant of a woman," "benefactor of every lover of every servant of every woman," etc.

In the first place, then, we have by (95)

sw = s(W' inline image W'' inline image W''' inline image etc.) = sW' inline image sW'' inline image sW''' inline image etc.
sw = sW' inline image W'' inline image W''' inline image etc. = sW' inline image sW'' inline image sW''' inline image etc.

From the last equation we have by (96)

sw = (sW'),(sW''),(sW'''), etc.

Now by (31) x' inline image x'' inline image etc. = x',x",x"', etc. inline image etc.,
or
(101) Π' ⤙ Σ',

where Π' and Σ' signify that the addition and multiplication with commas †1 are to be used. From this it follows that

(102) swsw. †2

If w vanishes, this equation fails, because in that case (95) does not hold.

From (102) we have
(103) (ls)w lsw. †3
Since a = a,b inline image etc.,
b = a,b inline image etc.,
we have la = l(a,b inline image etc.) = l(a,b) inline image l (etc.),
lb = l(a,b inline image etc.) = l(a,b) inline image l (etc.).

Multiplying these two equations commutatively we have

(la),(lb) = l(a,b) inline image etc.

or

(104) lΠ' ⤙ Π'l. †1
Now (ls)w = (s)W' inline image W'' inline image W''' inline image etc. = Π'(ls)W = Π'lsW,
lsw = lsW' inline image W'' inline image W''' inline image etc. = lΠ'sW' = lΠ'sW.
Hence,
(105) lsw ⤙ (ls)w,

or every lover of a servant of all women stands to every woman in the relation of lover of a servant of hers.

From (102) we have
(106) lswlsw. †2
By (95) and (96) we have
lsw = ls(W' inline image W'' inline image W''' etc.) = lsW' inline image lsW'' inline image lsW''' inline image etc.
= lsW' inline image lsW'' inline image lsW''' + etc.
Now sW = sW'inline image W'' inline image W''' inline image etc. = sW',sW'',sW''', etc.
So that by (94) swsW'sW'.
Hence by (92)
lsW' ⤙ lsw, lsW'' ⤙ lsw lsW''' ⤙ lsw.
Adding, lsW' inline image lsW'' inline image lsW''' ⤙ lsw;
or
(107) lsw ⤙ lsw.

That is, every lover of every servant of any particular woman is a lover of every servant of all women.

By (102) we have
(108) lswlsw. †3
Thus we have
lswlsw ⤙ lswlsw ⤙ (ls)wlsw. †P1

98. By similar reasoning we can easily make out the relations shown in the following table. It must be remembered
that the formulæ do not generally hold when exponents vanish. inline image †1

99. It appears to me that the advantage of the algebraic notation already begins to be perceptible, although its powers are thus far very imperfectly made out. At any rate, it seems to me that such a prima facie case is made out that the reader who still denies the utility of the algebra ought not to be too indolent to attempt to write down the above twenty-two terms in ordinary language With logical precision. Having done that, he has only to disarrange them and then restore the arrangement by ordinary logic, in order to test the algebra so far as it is yet developed.

Infinitesimal Relatives

100. We have by the binomial theorem by (49) and by (47),

(1 +x)n = 1 + Σpxn-p + xn.

Now, if we suppose the number of individuals to which any one thing is x to be reduced to a smaller and smaller number, we reach as our limit

x𝟐 = 0,

Σpxn-p = [n].𝟏n-†𝟏,x†𝟏 = xn,

(1 + x)n = 1 + x n.

101. If, on account of the vanishing of its powers, we call x an infinitesimal here and denote it by i, and if we put

x n = i n = y,

our equation becomes

(109) (1 +i) (y/i) = 1 + y.
Putting y = 1, and denoting (1 + i)(1/i) by inline image, we have
(110) inline image = (1 +i)1/i = 1 + 1.

102. In fact, this agrees With ordinary algebra better than it seems to do; for 1 is itself an infinitesimal, and inline image is inline image1. If the higher powers of 1 did not vanish, we should get the ordinary development of inline image.

103. Positive powers of inline image are absurdities in our notation. For negative powers we have

(111) inline image-x = 1 - x.

104. There are two ways of raising inline image-x to the yth power. In the first place, by the binomial theorem,

(1-x)y = 1-[y].1y-†𝟏,x†𝟏 + ([y].[y-𝟏]/𝟐).1y-‡𝟑,x‡𝟐—etc.;

and, in the second place, by (111) and (10).

inline image-xy = 1 - xy. †1

It thus appears that the sum of all the terms of the binomial development of (1-x)y, after the first, is -xy. †2 The truth of this may be shown by an example. Suppose the number of y's are four, viz. Y', Y'', Y''', and Y''''. Let us use x', x'', x''', and x'''' in such senses that

x Y' = x', x Y'' = x'', x Y''' = x''', x Y'''' = x''''.

Then the negatives of the different terms of the binomial developement are,

[y].1y-†𝟏,x†𝟏 = x' + x'' + x''' + x''''.
-(([y].[y-𝟏])/𝟐).1y-‡𝟐,x‡𝟐 = -x',x''-x',x'''-x',x''''-x'',x'''-x'',x''''-x''',x''''. †3
+(([y].[y-𝟏][y-𝟐])/𝟐.𝟑).1-||𝟑x||𝟑 = x',x'',x'''+ x',x'',x'''' + x',x''',x'''' + x'',x''',x''''.
xy = -x', x'', x''''. †4

Now, since this addition is invertible, in the first term, x' that is x'', is counted over twice, and so with every other pair. The second term subtracts each of these pairs, so that it is only counted once. But in the first term the x' that is x'' that is x''' is counted in three times only, while in the second term it is subtracted three times; namely, in (x',x''), in (x',x''') and in (x'',x'''). On the whole, therefore, a triplet would not be represented in the sum at all, were it not added by the third term. The whole quartette is included four times in the first term, is subtracted six times by the second term, and is added four times in the third term. The fourth term subtracts it once, and thus in the sum of these negative terms each combination occurs once, and once only; that is to say the sum is

x' inline image x'' inline image x''' inline image x'''' = x(Y' inline image Y'' inline image Y''' inline image Y'''') = xy.

105. If we write (a x)𝟑 for [x].[x-𝟏].[x-𝟐].1x-†𝟑,a†𝟑, that is for whatever is a to any three x's, regard being had for the order of the x's; and employ the modern numbers as exponents with this signification generally, then

1 - a x + (𝟏/𝟐!)(a x)2 - (𝟏/𝟑!)(a x)3 + etc.

is the development of (1 - a)x and consequently it reduces itself to 1 - a x. That is,

(112) x = x - (𝟏/𝟐!)x2 + (𝟏/𝟑!)x3 + (𝟏/𝟒!)x4 +etc.

106. 1 - x denotes everything except x, that is, whatever is other than every x; so that inline image- means "not." We shall take log x in such a sense that

inline imagelog x = x. †P1

107. I define the first difference of a function by the usual formula,

(113) Δφx = φ(x + Δx) - φx,

where Δx is an indefinite relative which never has a correlate in common with x. So that

(114) x,(Δx) = 0 x + Δx = x inline image Δx.

Higher differences may be defined by the formulæ

(115) Δ𝖓·x = 0 if 𝖓 > 1
Δ2x =ΔΔx = φ(x+𝟐.Δx)-𝟐.φ(x+Δx)+φx,
Δ3·φx = ΔΔ2.x = φ(x+𝟑.Δx)-𝟑.φ(x+𝟐.Δx)+𝟑.φ(xx)-φx.
(116) Δ𝖓·φx = φ(x+𝖓.Δx)-𝖓.φ(x+(𝖓-𝟏).Δx)
+ (𝖓.(𝖓-𝟏))/𝟐.φ(x+(𝖓-𝟐).Δx) - etc.

108. The exponents here affixed to Δ denote the number of times this operation is to be repeated, and thus have quite a different signification from that of the numerical coefficients in the binomial theorem. I have indicated the difference by putting a period after exponents significative of operational repetition. Thus, m2 may denote a mother of a certain pair, m2. a maternal grandmother.

109. Another circumstance to be observed is, that in taking the second difference of x, if we distinguish the two increments which x successively receives as Δ'x and Δ''x, then by (114)

(Δ'x),(Δ''x) = 0

If Δx is relative to so small a number of individuals that if the number were diminished by one Δ𝖓·φx would vanish, then I term these two corresponding differences differentials, and write them with 𝐝 instead of Δ.

110. The difference of the invertible sum of two functions is the sum of their differences; for by (113) and (18),

(117) Δ(φx + ψx) = φ(x + Δx) + ψ(x + Δx) - φx - ψx
= φ(x + Δx) - φx + ψ(x + Δx) - ψx = Δψx + Δψx.

If a is a constant, we have

(118) Δaφx = ax inline image Δφx) - aφx = aΔφx - (aΔφx),aφx,
Δ2.aφx = -Δaφx,aΔx, etc.
Δ(φx)a = (Δφx)a - ((Δφx)a),φxa,
Δ2.(φx)a = -Δ(φx)a, etc.
(119) φ(ax) = a,Δφx.

Let us differentiate the successive powers of x. We have in the first place,

Δ(x𝟐) = (x + Δx)𝟐 - x𝟐 = 𝟐.x𝟐-†𝟏,(Δx)†𝟏 + (Δx)𝟐.

Here, if we suppose Δx to be relative to only one individual, (Δx)𝟐 vanishes, and we have, with the aid of (115),

𝐝(x𝟐) = 𝟐.x𝟏,𝐝x .

Considering next the third power, we have, for the first differential,

Δ(x𝟑) = (x + Δx)𝟑 - x𝟑 = 𝟑.x𝟑-†𝟏,(Δx)†𝟏 + 𝟑.x𝟑-‡𝟐,(Δx)‡𝟐+(Δx)𝟑, 𝐝(x𝟑) = 𝟑.x𝟐,𝐝(x).

To obtain the second differential, we proceed as follows:

Δ𝟐.(x𝟑) = (x + 𝟑.Δx)𝟑 - 𝟐.(x + Δx)𝟑 + x𝟑
= x𝟑 + 𝟔.x𝟑-†𝟏,(Δx)†𝟏 + 𝟏𝟐.x𝟑-‡𝟐,(Δx)‡𝟐 + 𝟖.(Δx)𝟑
- 𝟐.x𝟑 - 𝟔.x𝟑-||𝟏,(Δx)||𝟏 - 𝟔.x𝟑-§𝟐,(Δx)§𝟐 - 2.(Δx)𝟑 + x𝟑
= 𝟔.x𝟑-‡𝟐,(Δx)‡𝟐 + 𝟔.(Δx)𝟑.

Here, if Δx is relative to less than two individuals, Δφx vanishes. Making it relative to two only, then, we have

𝐝𝟐.(x𝟑) = 𝟔.x𝟏,(𝐝x)𝟐.

These examples suffice to show what the differentials of x𝖓 will be. If for the number 𝖓 we substitute the logical term n, we have

Δ(xn) = (x + Δx)n - xn = [n].xn-†𝟏,(Δx)†𝟏 + etc.
𝐝(xn) = [n].xn-𝟏,(𝐝x).

We should thus readily find

(120) 𝐝𝖒·(xn) = [n].[n-𝟏].[n-𝟐]....[n-𝖒+𝟏].xn-†𝖒,(𝐝x)†𝖒.

Let us next differentiate lx. We have, in the first place,

Δlx = lx inline image Δx - lx = lx,lΔx - lx = lx,(lΔx - 1).

The value of lΔ'x - 1 is next to be found.

We have by (111) inline imagelΔx - 1 = lΔx.
Hence, lΔx - 1 = log lΔx.
But by (10) log lΔx = (log lx.

Substituting this value of l𝐝x - 1 in the equation lately found for 𝐝lx we have

(121) 𝐝lx = lx,(log,l) 𝐝x = lx,(l - 1) 𝐝x = -lx,(1 - l) 𝐝x.

111. In printing this paper, I here make an addition which supplies an omission in the account given above †1 of involution in this algebra. We have seen that every term which does not vanish is conceivable as logically divisible into individual terms. Thus we may write

s = S'inline image S'' inline image S''' inline image etc.

where not more than one individual is in any one of these relations to the same individual, although there is nothing to prevent the same person from being so related to many individuals. †1 Thus, "bishop of the see of" may be divided into first bishop, second bishop, etc., and only one person can be 𝖓th bishop of any one see, although the same person may (where translation is permitted) be 𝖓th bishop of several sees. Now let us denote the converse of x by 𝐊x; thus, if s is "servant of," 𝐊s is "master or mistress of." Then we have

𝐊s = 𝐊S' inline image 𝐊S'' inline image 𝐊S''' inline image etc.;

and here each of the terms of the second member evidently expresses such a relation that the same person cannot be so related to more than one, although more than one may be so related to the same. Thus, the converse of "bishop of the see of —" is "see one of whose bishops is —," the converse of "first bishop of —" is "see whose first bishop is —," etc. Now, the same see cannot be a see whose 𝖓th bishop is more than one individual, although several sees may be so related to the same individual. Such relatives I term infinitesimal on account of the vanishing of their higher powers. Every relative has a converse, and since this converse is conceivable as divisible into individual terms, the relative itself is conceivable as divisible into infinitesimal terms. To indicate this we may write

(122) If x > 0 x = X, inline image X,, inline image X,,, inline image etc.

112. As a term which vanishes is not an individual, nor is it composed of individuals, so it is neither an infinitesimal nor composed of infinitesimals.

As we write l S',l S'',l S''', etc. = ls,
so may we write
(123) L,s,L,,s,L,,,s, etc. = ls,

But as the first formula is affected by the circumstance that zero is not an individual, so that lsw does not vanish on account of no woman having the particular kind of servant denoted by S'', lsw denoting merely every lover of whatever servant there is of any woman; so the second formula is affected in a similar way, so that the vanishing of L[,]s does not make ls to vanish, but this is to be interpreted as denoting everything which is a lover, in whatever way it is a lover at all, of a servant. †1 Then just as we have by (112), that

(124) ls = 1 - (1 - l)s; †2
so we have
(125) ls = 1 - l(1 - s). †3

Mr. De Morgan denotes ls and ls by L S, and L,S respectively, †4 and he has traced out the manner of forming the converse and negative of such functions in detail. The following table contains most of his results in my notation. †5 For the converse of m, I write ɯ; and for that of n, u.

x 𝐊x
mn
mn - (1-m)(1-n)
mn = (1-m)(1-n)

uɯ = (1-u)(1-ɯ)
uɯ = (1-u)(1-ɯ)
inline image-x 𝐊inline image-x
(1-m)n = m(1-n)
(1-m)m
m(1-n)
ɯ(1-u) = (1-ɯ)u
ɯ(1-u)
(1-m)n

113. I shall term the operation by which w is changed to lw, backward involution. All the laws of this but one are the same as for ordinary involution, and the one exception is of that kind which is said to prove the rule. It is that whereas with ordinary involution we have,

(ls)w = l(sw);

in backward involution we have

(126) l(sw) = (ls)w;

that is, the things which are lovers to nothing but things that are servants to nothing but women are the things which are lovers of servants to nothing but women.

114. The other fundamental formulæ of backward involution are as follows:

(127) l inline image sw = lw,sw,

or, the things which are lovers or servants to nothing but women are the things which are lovers to nothing but women and servants to nothing but women.

(128) l(f,u) = lf,lu,

or, the things which are lovers to nothing but French violinists are the things that are lovers to nothing but Frenchmen and lovers to nothing but violinists. This is perhaps not quite axiomatic. It is proved as follows. By (125) and (30)

l(f,u) = inline image-l (1-f,u) = inline image-(l(1-f) inline image l(1-u))

By (125), (13), and (7),

lf,lu = inline image-l(1-f),inline image-l(1-u) = inline image-(l(1-f) inline image l(1-u)).

Finally, the binomial theorem holds with backward involution. For those persons who are lovers of nothing but Frenchmen and violinists consist first of those who are lovers of nothing but Frenchmen; second, of those who in some ways are lovers of nothing but Frenchmen and in all other ways of nothing but violinists, and finally of those who are lovers only of violinists. That is,

(129) l(u inline image f) = lu inline image Σpl-pu,pf inline image lf.

In order to retain the numerical coefficients, we must let {l} be the number of persons that one person is lover of. We can then write

l(u+f) = lu + {l}l-†𝟏u,†𝟏f+ (({l}·{l-𝟏})/𝟐)l-‡-𝟐u,‡𝟐f + etc.

115. We have also the following formula which combines the two involutions:

(130) l(sw) = (ls)w;

that is, the things which are lovers of nothing but what are servants of all women are the same as the things which are related to all women as lovers of nothing but their servants.

116. It is worth while to mention, in passing, a singular proposition derivable from (128). Since, by (124) and (125)

xy = (1-x)(1-y), †1

and since

1-(u inline image f) = inline image-(u inline image f) = inline image-u,inline image-f = (1-u),(1-f),

(128) gives us,

(1-l)(1-u),(1-f) = (1-l)(1-u) inline image Σp(1-(l-p))(1-u),
(1-p)(1-f) inline image (1-l)(1-f).

This is, of course, as true for u and f as for (1-u) and (1-f). Making those substitutions, and taking the negative of both sides, we have, by (124)

(131) l(u,f) = (lu),Π'p((l-p)u inline image pf),(lf),

or, the lovers of French violinists are those persons who, in reference to every mode of loving whatever, either in that way love some violinists or in some other way love some Frenchmen. This logical proposition is certainly not self-evident, and its practical importance is considerable. In a similar way, from (12) we obtain

(132) (e,c)f = Π'p(e(f-p) inline image cp),

that is, to say that a person is both emperor and conqueror of the same Frenchman is the same as to say that, taking any class of Frenchmen whatever, this person is either an emperor of some one of this class, or conqueror of some one among the remaining Frenchmen.

117. The properties of zero and unity, with reference to backward involution, are easily derived from (125). I give them here in comparison with the corresponding formulæ for forward involution.

(133) 0x = 1 x0= 1.
(134) q0 = 0 0r = 0,

where q is the converse of an unlimited relative, and r is greater than zero.

(135) 1x = x x1 = x.
(136) y1 = y 1z = z,

where y is infinitesimal, and z is individual. Otherwise, both vanish.

(137) 1s = 0 p1 = 0,

where s is less than unity and p is a limited relative.

(138) x1 = 1 1x = 1.

118. In other respects the formulæ for the two involutions are not so analogous as might be supposed; and this is owing to the dissimilarity between individuals and infinitesimals. We have, it is true, if X' is an infinitesimal and X' an individual,

(139) X,(y,z) = X,y,X,z like (y,z)X' = y X',z X';
(140) X,y0 = X,,X,y " X',y0 = X',y X';
(141) {X,} = 𝟏 " [X'] = 𝟏.

We also have

(142) X,yX,y.

But we have not X'y = X'y, and consequently we have not sw ⤙ sw, for this fails if there is anything which is not a servant at all, while the corresponding formula swsw only fails if there is not anything which is a woman. Now, it is much more often the case that there is something which is not x, than that there is not anything which is x. We have with the backward involution, as with the forward, †1 the formulæ

(143) If xy yzxz; †1
(144) If xy zxzy; †1

The former of these gives us

(145) lsw ⤙ (ls)w,

or, whatever is lover to nothing but what is servant to nothing but women †2 stands to nothing but a woman in the relation of lover of every servant of hers. The following formulæ can be proved without difficulty.

(146) lsw ⤙ lsw,

or, every lover of somebody who is servant to nothing but a woman stands to nothing but women in the relation of lover of nothing but a servant of them.

(147) lsw ⤙ l(sw),

or, whatever stands to a woman in the relation of lover of nothing but a servant of hers is a lover of nothing but servants of women.

The differentials of functions involving backward involution are

(148) 𝐝nx = {n}n-𝟏x,𝐝x.
(149) 𝐝xl = xl,𝐝x log.x.

In regard to powers of inline image we have

(150) xinline image = inline imagex.

Exponents with a dot may also be put upon either side of the letters which they affect.

119. The greater number of functions of x in this algebra may be put in the form

φx = Σp Σq pA pxq pBq.

For all such functions Taylor's and Maclaurin's theorems hold good in the form,

inline image

The symbol inline image is used to denote that a is to be substituted for b in what follows. For the sake of perspicuity, I will write Maclaurin's theorem at length.

inline image

The proof of these theorems is very simple. The (p+q)th differential of pxq is the only one which does not vanish when x vanishes. This differential then becomes [p+q]!.p(𝐝x)q. It is plain, therefore, that the theorems hold when the coefficients pAq and pBq are 1. But the general development, by Maclaurin's theorem, of aφx or (φx)a is in a form which (112) reduces to identity. It is very likely that the application of these theorems is not confined within the limits to which I have restricted it. We may write these theorems in the form

inline image

provided we assume that when the first differential is positive

inline image𝐝 = (𝟏/𝟎!)𝐝𝟎 + (𝟏/𝟏!)𝐝𝟏 + (𝟏/𝟐!)𝐝𝟐 + etc.,

but that when the first differential is negative this becomes by (111),

inline image𝐝 = 1 + 𝐝.

120. As another illustration of the use which may be made of differentiation in logic, let us consider the following problem. In a certain institution all the officers (x) and also all their common friends (f) are privileged persons (y). How shall the class of privileged persons be reduced to a minimum? Here we have

y = x + fx,
𝐝y = 𝐝x + 𝐝fx = 𝐝x - fx,(1-f)𝐝x.

When y is at a minimum it is not diminished either by an increase or diminution of x. That is,

[𝐝y] 0,

and when [x] is diminished by one,

[𝐝y] ⤙ 0 ,

When x is a minimum, then

[𝐝x-fx,(1-f)𝐝x] ⤚ 0 [𝐝x-fx-I,(1-f)𝐝x] ⤙ 0
(A) [𝐝x]-[fx,(1-f)𝐝x ⤚ 0 [𝐝x]-[fx-I,(1-f)𝐝x] ⤙ 0

Now we have by (30)

fx,(1-f)𝐝x = fx-(0;0),(1-f)𝐝x.

Hence,

[fx] ⤙ [𝐝x]+(0;0).[(1-f)𝐝x].

[fx-𝟏] [𝐝x]+(0;0).[(1-f)𝐝x].

But [0;0,] lies between the limits 0 and 1, and

(153) [𝐝x] = 𝟏.

We have, therefore,

[fx] ⤙ 1 + [(1-f)𝟏.] [fx-𝟏] ⤚ 𝟏.

This is the general solution of the problem. If the event of a person who may be an official in the institution being a friend of a second such person is independent of and equally probable with his being a friend of any third such person, and if we take p, or the whole class of such persons, for our universe, we have,

p = 1;

[fx,] = [fx]/[p] = ([f]/[p])[x],

[(1-f)𝐝x] = [1-f].[𝐝x] = ([p]-[f]).[𝐝x],

[fx,(1-f)𝐝x] = ([f])/[p])[x].([p]-[f]).[𝐝x]

Substituting these values in our equations marked (A) we get, by a little reduction,

[x] (log([p]-[f]))/(log[p]-log[f]),

[x] ⤙ (log([p]-[f]))/(log[p]-log[f]) + 𝟏.

The same solution would be reached through quite a different road by applying the calculus of finite differences in the usual way.

Elementary Relatives †1

121. By an elementary relative I mean one which signifies a relation which exists only between mutually exclusive pairs (or in the case of a conjugative term, triplets, or quartettes, etc.) of individuals, or else between pairs of classes in such a way that every individual of one class of the pair is in that relation to every individual of the other. If we suppose that in every school, every teacher teaches every pupil (a supposition which I shall tacitly make whenever in this paper I speak of a school), then pupil is an elementary relative. That every relative may be conceived of as a logical sum of elementary relatives is plain, from the fact that if a relation is sufficiently determined it can exist only between two individuals. Thus, a father is either father in the first ten years of the Christian era, or father in the second ten years, in the third ten years, in the first ten years, B. C., in the second ten years, or the third ten years, etc. Any one of these species of father is father for the first time or father for the second time, etc. Now such a relative as "father for the third time in the second decade of our era, of —" signifies a relation which can exist only between mutually exclusive pairs of individuals, and is therefore an elementary relative; and so the relative father may be resolved into a logical sum of elementary relatives.

122. The conception of a relative as resolvable into elementary relatives has the same sort of utility as the conception of a relative as resolvable into infinitesimals or of any term as resolvable into individuals.

123. Elementary simple relatives are connected together in systems of four. For if A:B be taken to denote the elementary relative which multiplied into B gives A, then this relation
existing as elementary, we have the four elementary relatives

A:A A:B B:A B:B.

An example of such a system is—colleague: teacher: pupil: schoolmate. In the same way, obviously, elementary conjugatives are in systems the number of members in which is (𝖓+1)𝖓+1 where 𝖓 is the number of correlates which the conjugative has. At present, I shall consider only the simple relatives.

124. The existence of an elementary relation supposes the existence of mutually exclusive pairs of classes. The first members of those pairs have something in common which discriminates them from the second members, and may therefore be united in one class, while the second members are united into a second class. Thus pupil is not an elementary relative unless there is an absolute distinction between those who teach and those who are taught. We have, therefore, two general absolute terms which are mutually exclusive, "body of teachers in a school," and "body of pupils in a school." These terms are general because it remains undetermined what school is referred to. I shall call the two mutually exclusive absolute terms which any system of elementary relatives supposes, the universal extremes of that system. There are certain characters in respect to the possession of which both members of any one of the pairs, between which there is a certain elementary relation, agree. Thus, the body of teachers and the body of pupils in any school agree in respect to the country and age in which they live, etc., etc. Such characters I term scalar characters for the system of elementary relatives to which they are so related; and the relatives written with a comma which signify the possession of such characters, I term scalars for the system. Thus, supposing French teachers have only French pupils and vice versa, the relative

f,

will be a scalar for the system "colleague: teacher: pupil: schoolmate." If r is an elementary relative for which s is a scalar,

(154) s,r = rs,.

125. Let c, t, p, s, denote the four elementary relatives of any system; such as colleague, teacher, pupil, schoolmate; and let a,, b,, c,, d,, be scalars for this system. Then any relative which is capable of expression in the form

a,c + b,t + c,p + d,s

I shall call a logical quaternion. Let such relatives be denoted by q, q', q'', etc. It is plain, then, from what has been said, that any relative may be regarded as resolvable into a logical sum of logical quaternions.

126. The multiplication of elementary relatives of the same system follows a very simple law. For if u and v be the two universal extremes of the system c, t, p, s, we may write

c = u:u t = u:v p = v:u s = v:v,

and then if w and w' are each either u or v, we have

(155) (w':w)inline image-w = 0.

This gives us the following multiplication-table, where the multiplier is to be entered at the side of the table and the multiplicand at the top, and the product is found in the middle:

(156)

c t p s
c c t 0 0
t 0 0 c t
p p s 0 0
s 0 0 p s

The sixteen propositions expressed by this table are in ordinary language as follows: †1

The colleagues of the colleagues of any person are that person's colleagues;

The colleagues of the teachers of any person are that person's teachers;

There are no colleagues of any person's pupils;

There are no colleagues of any person's schoolmates;

There are no teachers of any person's colleagues;

There are no teachers of any person's teachers;

The teachers of the pupils of any person are that person's colleagues;

The teachers of the schoolmates of any person are that person's teachers;

The pupils of the colleagues of any person are that person's pupils;

The pupils of the teachers of any person are that person's schoolmates;

There are no pupils of any person's pupils;

There are no pupils of any person's schoolmates;

There are no schoolmates of any person's colleagues;

There are no schoolmates of any person's teachers;

The schoolmates of the pupils of any person are that person's pupils;

The schoolmates of the schoolmates of any person are that person's schoolmates.

This simplicity and regularity in the multiplication of elementary relatives must clearly enhance the utility of the conception of a relative as resolvable into a sum of logical quaternions.

127. It may sometimes be convenient to consider relatives each one of which is of the form

a,i + b,j + c,k + d,l + etc.

where a,, b,, c,, d,, etc. are scalars, and i, j, k, l, etc. are each of the form

m,u + n,v + o,w + etc.

where m,, n,, o,, etc. are scalars, and u, v, w, etc. are elementary relatives. In all such cases (155) Will give a multiplication-table for i, j, k, l, etc. For example, if we have three classes of individuals, u[1], u[2], u[3], which are related to one another in pairs, we may put

u1:u1 = i u1:u2 = j u1:u3 = k
u2:u1 = l u2:u2 = m u2:u3 = n
u3:u1 = o u3:u2 = p u3:u3 = q

and by (155) we get the multiplication-table

i j k l m n o p q
i i j k 0 0 0 0 0 0
j 0 0 0 i j k 0 0 0
k 0 0 0 0 0 0 i j k
l l m n 0 0 0 0 0 0
m 0 0 0 l m n 0 0 0
n 0 0 0 0 0 0 l m m
o o p q 0 0 0 0 0 0
p 0 0 0 o p q 0 0 0
q 0 0 0 0 0 0 o p q

128. If we take

i = u1:u2 + u2:u3 + u3:u4,

j = u1:u3 + u2:u4,

k = 2.u1:u4,

we have

i j k
i j k 0
j k 0 0
k 0 0 0

129. If we take

i = u1:u2 + u2:u3 + u3:u4 +u5:u6 + u7:u8,

j = u1:u3 +u2:u4,

k = 2.u1:u4,

l = u6:u8 + 𝖆.u5:u7 + 2𝖇.u1:u9 + u9:u4 + 𝖈.u5:u6,

m = u5:u8,

we have

i j k l m
i j k 0 m 0
j k 0 0 0 0
k 0 0 0 0 0
l 𝖆.m 0 0 𝖇.k+
𝖈.m
0
m 0 0 0 0 0

130. These multiplication-tables have been copied from Professor Peirce's monograph on Linear Associative Algebras. †P1 I can assert, upon reasonable inductive evidence, that all such algebras can be interpreted on the principles of the present notation in the same way as those given above. In other words, all such algebras are complications and modifications of the algebra of (156). It is very likely that this is true of all algebras whatever. The algebra of (156), which is of such a fundamental character in reference to pure algebra and our logical notation, has been shown by Professor Peirce †1 to be the algebra of Hamilton's quaternions. †2 In fact, if we put

inline image

where a, b, c, are scalars, then 1, i', j', k' are the four fundamental factors of quaternions, the multiplication-table of which is as follows:

1 i' j' k'
1 1 i' j' k'
i' i' -1 k' -j'
j' j' -k' -1 i'
k' k' j' -i' -1

131. It is no part of my present purpose to consider the bearing upon the philosophy of space of this occurrence, in pure logic, of the algebra which expresses all the properties of space; but it is proper to point out that one method of working with this notation would be to transform the given logical expressions into the form of Hamilton's quaternions (after representing them as separated into elementary relatives), and then to make use of geometrical reasoning. The following formulæ will assist this procesS. Take the quaternion relative

q = xi + yj + zk + w l,

where x, y, z, and w are scalars. The conditions of q being a scalar, vector, etc. (that is, being denoted by an algebraic expression which denotes a scalar, a vector, etc., in geometry), are

(157) Form of a scalar: x(i + l).
(158) Form of a vector: xi+y i+zk-x l.
(159) Form of a versor:
x/y((x/z)-1)i + y/x((x/z)-1)j + z/y((z/x)-1)k + y/z((z/x)-1)l.
(160) Form of zero: xi + x y j + (z/y)k + z l.
(161) Scalar of q: Sq = ½(x + w)(i + 1).
(162) Vector of q: Vq = ½(x-w)i + yj + zk + ½(w-x)l.
(163) Tensor of q: inline image
(164) Conjugate of q: Kq = w i - yj - zk + x l.

132. In order to exhibit the logical interpretations of these functions, let us consider a universe of married monogamists, in which husband and wife always have country, race, wealth, and virtue, in common. Let i denote "man that is —," j "husband of —," k "wife of —," and l "woman that is —"; x "negro that is —," y "rich person that is —," z "American that is —," and w "thief that is —." Then, q being defined as above, the q's of any class will consist of so many individuals of that class as are negro-men or women-thieves together with all persons who are rich husbands or American wives of persons of this class. Then, 2Sq denotes, by (160), †1 all the negroes and besides all the thieves, while Sq is the indefinite term which denotes half the negroes and thieves. Now, those persons who are self-q's of any class (that is, the q's of themselves among that class) are xi + w l; add to these their spouses and we have 2Sq. In general, let us term (j + k) the "correspondent of —." Then, the double scalar of any quaternion relative, q, is that relative which denotes all self-q's, and, besides, "all correspondents of self-q's of —." (Tq)2 denotes all persons belonging to pairs of corresponding self-q's minus all persons belonging to pairs of corresponding q's of each other.

133. As a very simple example of the application of geometry to the logic of relatives, we may take the following. Euclid's axiom concerning parallels corresponds to the quaternion principle that the square of a vector is a scalar. From this it follows, since by (157) yz + zk [?] is a vector, that the rich husbands and American wives of the rich husbands and American wives of any class of persons are wholly contained under that class, and can be described without any discrimination of sex. In point of fact, by (156), the rich husbands and American wiveS of the rich husbands and American wives of any class of persons, are the rich Americans of that class.

Lobatchewsky †1 has shown that Euclid's axiom concerning parallels may be supposed to be false without invalidating the propositions of spherical trigonometry. In order, then, that corresponding propositions should hold good in logic, we need not resort to elementary relatives, but need only take S and V in such senses that every relative of the class considered should be capable of being regarded as a sum of a scalar and a vector, and that a scalar multiplied by a scalar should be a scalar, while the product of a scalar and a vector is a vector. Now, to fulfill these conditions we have only to take Sq as "self-q of," and Vq as "alio-q of" (q of another, that other being —), and q may be any relative whatever. For, "lover," for example, is divisible into self-lover and alio-lover; a self-lover of a self-benefactor of personS of any class is contained under that class, and neither the self-lover of an alio-benefactor of any persons nor the alio-lover of the self-benefactor of any persons are among those persons. Suppose, then, we take the formula of spherical trigonometry,

cos a = cos b cos c + cos A sin b sin c.

In quaternion form, this is,

(165) S(p q) = (Sp)(Sq)+ S((Vp) (Vq)).

Let p be "lover," and q be "benefactor." Then this reads, lovers of their own benefactors consist of self-lovers of self-benefactors together with alio-lovers of alio-benefactors of themselves. So the formula

sin b cos p b'= -sin a cos c cos p a' -sin c cos a cos p c'+ sin a sin c sin b cos p b,

where A', B', C', are the positive poles of the sides a, b, c, is in quaternions

(166) V(p q) = (Vp)(Sq) + (Sp)(Vq) + V((Vp)(Vq)),

and the logical interpretation of this is: lovers of benefactors of others consist of alio-lovers of self-benefactors, together with self-lovers of alio-benefactors, together with alio-lovers of alio-benefactors of others. It is a little striking that just as in the non-Euclidean or imaginary geometry of Lobatchewsky the axiom concerning parallels holds good only with the ultimate elements of space, so its logical equivalent holds good only for elementary relatives.

134. It follows from what has been said that for every proposition in geometry there is a proposition in the pure logic of relatives. But the method of working with logical algebra which is founded on this principle seems to be of little use. On the other hand, the fact promises to throw some light upon the philosophy of space. †P1

§6. Properties of Particular Relative Terms Classification of Simple Relatives †1

135. Any particular property which any class of relative terms may have may be stated in the form of an equation, and affords us another premiss for the solution of problems in which such terms occur. A good classification of relatives is, therefore, a great aid in the use of this notation, as the notation is also an aid in forming such a classification.

136. The first division of relatives is, of course, into simple relatives and conjugatives. The most fundamental divisions of simple relatives are based on the distinction between elementary relatives of the form (A:A), and those of the form (A:B). These are divisions in regard to the amount of opposition between relative and correlative.

a. Simple relatives are in this way primarily divisible into relatives all of whose elements are of the form (A:A) and those which contain elements of the form (A:B). The former express a mere agreement among things, the latter set one thing over against another, and in that sense express an opposition ({antikeisthai}); I shall therefore term the former concurrents, †P1 and the latter opponents. The distinction appears in this notation as between relatives with a comma, such as (w,), and relatives without a comma, such as (w); and is evidently of the highest importance. The character which is signified by a concurrent relative is an absolute character, that signified by an opponent is a relative character, that is, one which cannot be prescinded from reference to a correlate.

b. The second division of simple relatives with reference to the amount of opposition between relative and correlative is into those whose elements may be arranged in collections of squares, each square like this,

A:A A:B A:C
B:A B:B B:C
C:A C:B C:C

and those whose elements cannot be so arranged. †1 The former (examples of which are, "equal to —," "similar to —") may be called copulatives, †P2 the latter non-copulatives. A copulative multiplied into itself gives itself. Professor Peirce calls letters having this property, idempotents. †2 The present distinction is of course very important in pure algebra. All concurrents are copulatives.

c. Third, relatives are divisible into those which for every element of the form (A:B) have another of the form (B:A), and those which want this symmetry. This is the old division into equiparants †P1 and disquiparants, †P2 or in Professor De Morgan's language, convertible and inconvertible relatives. †1 Equiparants are their own correlatives. All copulatives are equiparant.

d. Fourth, simple relatives are divisible into those which contain elements of the form (A:A) and those which do not. The former express relations such as a thing may have to itself, the latter (as cousin of —, hater of —) relations which nothing can have to itself. The former may be termed self-relatives, †P3 the latter alio-relatives. All copulatives are self-relatives.

e. The fifth division is into relatives some power (i.e. repeated product) of which contains †P4 elements of the form (A:A), and those of which this is not true. †2 The former I term cyclic, the latter non-cyclic †P5 relatives. As an example of the former, take

(A:B) inline image (B:A) inline image (C:D) inline image (D:E) inline image (E:C).

The product of this into itself is

(A:A) inline image (B:B) inline image (C:E) inline image (D:C) inline image (E:D).

The third power is

(A:B) inline image (B:A) inline image (C:C) inline image (D:D) inline image (E:E).

The fourth power is

(A:A) inline image (B:B) inline image (C:D) inline image (D:E) inline image (E:C).

The fifth power is

(A:B) inline image (B:A) inline image (C:E) inline image (D:C) inline image (E:D).

The sixth power is

(A:A) inline image (B:B) inline image (C:C) inline image (D:D) inline image (E:E).

where all the terms are of the form (A:A). Such relatives, as cousin of —, are cyclic. All equiparants are cyclic.

f. The sixth division is into relatives no power of which is zero, and relatives some power of which is zero. The former may be termed inexhaustible, the latter exhaustible. An example of the former is "spouse of —," of the latter, "husband of —." All cyclics are inexhaustible.

g. Seventh, simple relatives may be divided into those whose products into themselves are not zero, and those whose products into themselves are zero. The former may be termed repeating, the latter, non-repeating relatives. All inexhaustible relatives are repeating.

h. Repeating relatives may be divided (after De Morgan) into those whose products into themselves are contained under themselves, and those of which this is not true. The former are well named by De Morgan †1 transitive, the latter intransitive. All transitives are inexhaustible; all copulatives are transitive; and all transitive equiparants are copulative. The class of transitive equiparants has a character, that of being self-relatives, not involved in the definitions of the terms. †2

i. Transitives are further divisible into those whose products by themselves are equal to themselves, and those whose products by themselves are less than themselves; the former may be termed continuous, †P1 the latter discontinuous. An example of the second is found in the pure mathematics of a continuum, where if a is greater than b it is greater than something greater than b; and as long as a and b are not of the same magnitude, an intervening magnitude always exists. All concurrents are continuous.

j. Intransitives may be divided into those the number of the powers (repeated products) of which not contained in the first is infinite, and those some power of which is contained in the first. The former may be called infinites, the latter finites. Infinite inexhaustibles are cyclic.
In addition to these, the old divisions of relations into relations of reason and real relations, of the latter into aptitudinal and actual, and of the last into extrinsic and intrinsic, are often useful. †P1

"NOT"

137. We have already seen that "not," or "other than," is denoted by inline image-1. It is often more convenient to write it, n. The fundamental property of this relative has been given above (111). It is that,

inline image-x = 1 - x.

Two other properties are expressed by the principles of contradiction and excluded middle. They are,

x,inline image-x = 0; †1 x inline image inline image-x = 1. †1

The following deduced properties are of frequent application:

(167) inline image-(x,y) = inline image-x inline image inline image-y; †2
(168) inline image-xy = inline image-xy

The former of these is the counterpart of the general formula, zx inline image y = zx,zy †2, The latter enables us always to bring the exponent of the exponent of inline image- down to the line, and make it a factor. By the former principle, objects not French violinists consist of objects not Frenchmen, together with objects not violinists; by the latter, individuals not servants of all women are the same as non-servants of some women.

Another singular property of inline image- is that,

If [x] > 1 inline image-1x = 1.

"Case of the Existence of —," and "Case of the Non-Existence of —."

138. That which first led me to seek for the present extension of Boole's logical notation was the consideration that as he left his algebra, neither hypothetical propositions nor particular propositions could be properly expressed. It is true that Boole was able to express hypothetical propositions in a way which answered some purposes perfectly. He could, for example, express the proposition, "Either the sun will shine, or the enterprise will be postponed," by letting x denote "the truth of the proposition that the sun will shine," and y "the truth of the proposition that the enterprise will be postponed"; and writing,

x inline image y = 1,

or, with the invertible addition,

x + (1 - x),y = 1.

But if he had given four letters denoting the four terms, "sun," "what is about to shine," "the enterprise," and "what is about to be postponed," he could make no use of these to express his disjunctive proposition, but would be obliged to assume others. The imperfection of the algebra here was obvious. As for particular propositions, Boole could not accurately express them at all. He did undertake to express them and wrote

Some Y's are X's: v,y = v,x;
Some Y's are not X's: v,y = v,(1-x).

The letter v is here used, says Boole, for an "indefinite class symbol." †1 This betrays a radical misapprehension of the nature of a particular proposition. To say that some Y's are X's, is not the same as saying that a logical species of Y's are X's. For the logical species need not be the name of anything existing. It is only a certain description of things fully expressed by a mere definition, and it is a question of fact whether such a thing really exist or not. St. Anselm wished to infer existence from a definition, but that argument has long been exploded. If, then, v is a mere logical species in general, there is not necessarily any such thing, and the equation means nothing. If it is to be a logical species, then, it is necessary to suppose in addition that it exists, and further that some v is y. In short, it is necessary to assume concerning it the truth of a proposition, which, being itself particular, presents the original difficulty in regard to its symbolical expression. Moreover, from

v,y = v,(1-x)

we can, according to algebraic principles, deduce successively

v,y = v - v,x

v,x = v - v,y = v,(1-y).

Now if the first equation means that some Y's are not X's, the last ought to mean that some X's are not Y's; for the algebraic forms are the same, and the question is, whether the algebraic forms are adequate to the expression of particulars. It would appear, therefore, that the inference from Some Y's are not X's to Some X's are not Y's, is good; but it is not so, in fact.

139. What is wanted, in order to express hypotheticals and particulars analytically, is a relative term which shall denote "case of the existence of —," or "what exists only if there is any —"; or else "case of the non-existence of —," or "what exists only if there is not —." When Boole's algebra is extended to relative terms, it is easy to see what these particular relatives must be. For suppose that having expressed the propositions "it thunders," and "it lightens," we wish to express the fact that "if it lightens, it thunders." Let

A = 0 and B = 0,

be equations meaning respectively, it lightens and it thunders. Then, if φx vanishes when x does not and vice versa, whatever x may be, the formula

φA ⤙ φB

expresses that if it lightens it thunders; for if it lightens, A vanishes; hence φA does not vanish, hence φB does not vanish, hence B vanishes, hence it thunders. It makes no difference what the function φ is, provided only it fulfills the condition mentioned. Now, 0x is such a function, vanishing when x does not, and not vanishing when x does. Zero, therefore, may be interpreted as denoting "that which exists if, and only if, there is not —." Then the equation

00 = 1

means, everything which exists, exists only if there is not anything which does not exist. So,

0x = 0

means that there is nothing which exists if, and only if, some x does not exist. The reason of this is that some x means some existing x.

"It lightens" and "it thunders" might have been expressed by equations in the forms

A = 1, B = 1.

In that case, in order to express that if it lightens it thunders, in the form

φA ⤙ φB,

it would only be necessary to find a function, φx, which should vanish unless x were 1, and should not vanish if x were 1. Such a function is 1x. We must therefore interpret 1 as "that which exists if, and only if, there is —," 1x as "that which exists if, and only if, there is nothing but x," and 1x as "that which exists if, and only if, there is some x." Then the equation

1x = 1,

means everything exists if, and only if, whatever x there is exists.

140. Every hypothetical proposition may be put into four equivalent forms, as follows:

If X, then Y.

If not Y, then not X.

Either not X or Y.

Not both X and not Y.

If the propositions X and Y are A = 1 and B = 1, these four forms are naturally expressed by

1A ⤙ 1B,

1(1-A) ⤙ 1(1-B), †1

1(1-A) inline image B = 1,

1A, 1(1-B) = 0.

For 1x we may always substitute 0(1-x).

141. Particular propositions are expressed by the consideration that they are contradictory of universal propositions. Thus, as h,(1-b) = 0 means every horse is black, so 0h,(1-b) = 0 means that some horse is not black; and as h,b = 0 means that no horse is black, so 0h,b = 0 means that some horse is black. We may also write the particular affirmative 1(h,b) = 1, and the particular negative 1(h,nb) = 1.

142. Given the premisses, every horse is black, and every horse is an animal; required the conclusion. We have given

h ⤙ b;

h ⤙ a.

Commutatively multiplying, we get

h ⤙ a,b.

Then, by (92) or by (90),

0a,b ⤙ 0h, or 1h ⤙ 1(a,b).

Hence, by (40) or by (46),

If h > 0 0a,b = 0, or 1(a,b) = 1;

or if there are any horses, some animals are black. I think it would be difficult to reach this conclusion, by Boole's method unmodified.

143. Particular propositions may also be expressed by means of the signs of inequality. Thus, some animals are horses, may be written

a,h > 0;

and the conclusion required in the above problem might have been obtained in thiS form, very easily, from the product of the premisses, by (1) and (21).

We shall presently see †1 that conditional and disjunctive propositions may also be expressed in a different way.

Conjugative Terms

144. The treatment of conjugative terms presents considerable difficulty, and would no doubt be greatly facilitated by algebraic devices. I have, however, studied this part of my notation but little.

A relative term cannot possibly be reduced to any combination of absolute terms, nor can a conjugative term be reduced to any combination of simple relatives; but a conjugative having more than two correlates can always be reduced to a combination of conjugatives of two correlates. Thus for "winner over of —, from —, to —," we may always substitute 𝐮, or "gainer of the advantage — to —," where the first correlate is itself to be another conjugative 𝐯, or "the advantage of winning over of — from —." Then we may write,

𝐰 = 𝐮 𝐯.

It is evident that in this way all conjugatives may be expressed as production of conjugatives of two correlates.

145. The interpretation of such combinations as 𝐛am, etc., is not very easy. When the conjugative and its first correlative can be taken together apart from the second correlative, as in (𝐛a)m and (𝐛a)m and (𝐛a)m and (𝐛a)m, there is no perplexity, because in such cases (𝐛a) or (𝐛a) is a simple relative. We have, therefore, only to call the betrayer to an enemy an inimical betrayer, when we have

(𝐛a)m = inimical betrayer of a man = betrayer of a man to an enemy of him,

(𝐛a)m = inimical betrayer of every man = betrayer of every man to an enemy of him.

And we have only to call the betrayer to every enemy an unbounded betrayer, in order to get

(𝐛a)m = unbounded betrayer of a man = betrayer of a man to every enemy of him,

(𝐛a)m = unbounded betrayer of every man = betrayer of every man to every enemy of him.

The two terms 𝐛am and 𝐛am are not quite so easily interpreted. Imagine a separated into infinitesimal relatives, A,,A,,,A,,,, etc., each of which is relative to but one individual which is m. Then, because all powers of A,,A,,,A,,,, etc., higher than the first, vanish, and because the number of such terms must be m, we have,

am = (A, inline image A,, inline image A,,, inline image etc.)m = (A,m),(A,,m),(A,,,m), etc.

or if M', M'', M''', etc., are the individual m's,

am = (A,M'),(A,,M''),(A,,,M'''), etc.

It is evident from this that 𝐛am is a betrayer to an A, of M', to an A,, of M'', to an A,,, of M''', etc., in short of all men to some enemy of them all. In order to interpret 𝐛am we have only to take the negative of it. This, by (124), is (1-𝐛)am, or a non-betrayer of all men to some enemy of them. Hence, 𝐛am, or that which is not this, is a betrayer of some man to each enemy of all men. To interpret 𝐛(am) we may put it in the form (1-𝐛)(1-a)m. This is "non-betrayer of a man to all non-enemies of all men." Now, a non-betrayer of some X to every Y, is the same as a betrayer of all X's to nothing but what is not Y; and the negative-of "non-enemy of all men," is "enemy of a man." Thus, 𝐛(am) is, "betrayer of all men to nothing but an enemy of a man." To interpret 𝐛am we may put it in the form (1-𝐛)(1-a)m, which is, "non-betrayer of a man to every non-enemy of him." This is a logical sum of terms, each of which is "non-betrayer of an individual man M to every non-enemy of M." Each of these terms is the same as "betrayer of M to nothing but an enemy of M." The sum of them, therefore, which is 𝐛am is "betrayer of some man to nothing but an enemy of him." In the same way it is obvious that 𝐛am is "betrayer of nothing but a man to nothing but an enemy of him." We have 𝐛am = 𝐛(1-a)(1-m) or "betrayer of all non-men to a non-enemy of all non-men." This is the same as "that which stands to something which is an enemy of nothing but a man in the relation of betrayer of nothing but men to what is not it." The interpretation of 𝐛am is obviously "betrayer of nothing but a man to an enemy of him." It is equally plain that 𝐛am is "betrayer of no man to anything but an enemy of him," and that 𝐛am is "betrayer of nothing but a man to every enemy of him." By putting 𝐛am in the form 𝐛(1-a)(1-m) we find that it denotes "betrayer of something besides a man to all things which are enemies of nothing but men." When an absolute term is put in place of a, the interpretations are obtained in the same way, with greater facility.

146. The sign of an operation is plainly a conjugative term. Thus, our commutative multiplication might be denoted by the conjugative

1,.

For we have

l,sw = 1,l,sw.

As conjugatives can all be reduced to conjugatives of two correlates, they might be expressed by an operative sign (for which a Hebrew letter might be used) put between the symbols for the two correlates. There would often be an advantage in doing this, owing to the intricacy of the usual notation for conjugatives. If these operational signs happened to agree in their properties with any of the signs of algebra, modifications of the algebraic signs might be used in place of Hebrew letters. For instance, if inline image were such that

inline imagexinline imageyz = inline image[13]inline imageyz, †1

then, if we were to substitute for inline image the operational sign ר we have

xר(yרz) = (xרy)רz,

which is the expression of the associative principle. So, if

inline imagexy = inline imageyx

we may write,

xרy = yרx

which is the commutative principle. If both these equations held for any conjugative, we might conveniently express it by a modified sign +. For example, let us consider the conjugative "what is denoted by a term which either denotes — or else —." For this, the above principles obviously hold, and we may naturally denote it by inline image. Then, if p denotes Protestantism, r Romanism, and f what is false,

p inline image r ⤙ f

means either all Protestantism or all Romanism is false. In this way it is plain that all hypothetical propositions may be expressed. Moreover, if we suppose any term as "man" (m) to be separated into its individuals, M', M'', M''', etc., then,

M' inline image M'' inline image M''' inline image etc.,

means "some man." This may very naturally be written

'm'

and this gives us an improved way of writing a particular proposition; for

'x' ⤙ y

seems a simpler way of writing "Some X is Y" than

0x,y = 0.

Converse

147. If we separate lover into its elementary relatives, take the reciprocal of each of these, that is, change it from

A:B to B:A,

and sum these reciprocals, we obtain the relative loved by. There is no such operation as this in ordinary arithmetic, but if we suppose a science of discrete quantity in quaternion form (a science of equal intervals in space), the sum of the reciprocals of the units of such a quaternion will be the conjugate-quaternion. For this reason, I express the conjugative term "what is related in the way that to — is —, to the latter" by 𝐊. The fundamental equations upon which the properties of this term depend are

(169) K K = 1.
(170) If x < yz then z ⤙ (𝐊y)x,
or 1(x,y z) = 1(z,𝐊yx)
We have, also,  
(171) 𝐊Σ = Σ𝐊,
(172) 𝐊Π = Π𝐊,
where Π denotes the product in the reverse order. Other equations will be found in Mr. De Morgan's table, given above. †1

Conclusion

148. If the question is asked, What are the axiomatic principles of this branch of logic, not deducible from others? I reply that whatever rank is assigned to the laws of contradiction and excluded middle belongs equally to the interpretations of all the general equations given under the head of "Application of the Algebraic signs to Logic," together with those relating to backward involution, and the principles expressed by equations (95), (96), (122), (142), (156), (25), (26), (14), (15).

149. But these axioms are mere substitutes for definitions of the universal logical relations, and so far as these can be defined, all axioms may be dispensed with. The fundamental principles of formal logic are not properly axioms, but definitions and divisions; and the only facts which it contains relate to the identity of the conceptions resulting from those processes with certain familiar ones.



Paper 4: On the Application of Logical Analysis to Multiple Algebra †1

150. The letters of an algebra express the relation of the product to the multiplicand. Thus, i A expresses the quantity which is related to A in the manner denoted by i. This being the conception of these algebras, for each of them we may imagine another "absolute" algebra, as we may call it, which shall contain letters which can only be products and multiplicands, not multipliers. Let the general expression of the absolute algebra be a I + b J + c K + d L + etc. Multiply this by any letter i of the relative algebra, and denote the product by

(A1a + A2b + A3c + etc.)I.
+ (B1a + B2b + B3c + etc.)J.
+ etc.

Now we may obviously enlarge the given relative algebra, so that

i = A1i11 + A2i12 + A3i13 + etc.
+ B1i21 + B2i22 + B3i23 + etc.
+ etc.

where i11i12 etc., are such that the product of either of them into any letter of the absolute algebra shall equal some letter of that algebra. That there is no self-contradiction involved in this supposition seems axiomatic. †2

151. In this way each letter of the given algebra is resolved into a sum of terms of the form a A:B, a being a scalar, and A:B such that

(A:B)(B:C) = A:C.

(A:B)(C:D) = 0.

The actual resolution is usually performed with ease, but in some cases a good deal of ingenuity is required. I have not found the process facilitated by any general rules. I have actually resolved all the Double, Triple, and Quadruple algebras, and all the Quintuple ones, that appeared to present any difficulty. I give a few examples.

bi5. †1

i j k l m
i j 0 l 0 0
j 0 0 0 0 0
k j + al 0 0 0 bj + cl
l 0 0 0 0 0
m a'j + b'l 0 c'j + d'l 0 l

i = cd'A:B+b'B:C+b'D:E.

j = b'cd'A:C.

k = cd'A:B+a c d'D:B+b'c2d'D:F+cd'E:C+b b'cd'A:F.

l = b'cd'D:C.

m = a'cd'A:B+b'c'A:E+b'cd'D:B+b'd'D:E+b'cd'D:F+F:C.

bd5. †2

i j k l m
i j 0 l 0 0
j 0 0 0 0 0
k j + rl 0 i + m 0 - j - rl
l 0 0 j 0 0
m (r2 - 1)j 0 -l 0 -r2j

i = A:D+D:F+B:E+C:F.

j = A:F.

k = r A:B+r B:C+D:E-(1/r)D:F+E:F.

l = A:E-(1/r)A:F+B:F.

m = r2A:C-A:D-B:E-C:F.

bh6. †1

i j k l m n
i i j k l m
j j k
k k
l l ak k
m k
n n

i = A:A+B:B+C:C+D:D.

j = A:B+B:C.

k = A:C.

l = a A:B+A:D+D:C.

m = A:E.

n = E:C.

br5. †2

i j k l m
i j
j
k l m
l
m

i = A:B+B:C

j = A:C.

k = D:B+D:E+E:F.

l = D:C.

m = D:F.



Paper 5: Note on Grassman's Calculus of Extension †1

152. The last Mathematische Annalen contains a paper by H. Grassmann, on the application of his calculus of extension to mechanics. †2

He adopts the quaternion addition of vectors. But he has two multiplications, internal and external, just as the principles of logic require.

The internal product of two vectors, v1 and v2, is simply what is written in quaternions as—S.v1v2. He writes it [v1|v2]. So that

[v1|v2] = [v2|v1],

v2 = (Tv)2.

The external product of two vectors is the parallelogram they form, account being taken of its plane and the direction of running round it, which is equivalent to its aspect. We therefore have:

inline image

where I is a new unit. This reminds me strongly of what is written in quaternions as — V(v1v2). But it is not the same thing in fact, because [v1v2]v3 is a solid, and therefore a new kind of quantity. In truth, Grassmann has got hold (though he did not say so) of an eight-fold algebra, which may be written in my system as follows:

Three Rectangular Versors †1

i = M:A - B:Z + C:Y + X:N

j = M:B - C:X + A:Z + Y:N

k = M:C - A:Y + B:X + Z:N

Three Rectangular Planes

I = M:X + A:N

J = M:Y + B:N

K = M:Z + C:N

One Solid

V = M:N

Unity

1 = M:M + A:A + B:B + C:C + N:N + X:X + Y:Y + Z:Z

This unity might be omitted.

153. The recognition †2 of the two multiplications is exceedingly interesting. The system seems to me more suitable to three dimensional space, and also more natural than that of quaternions. The simplification of mechanical formulæ is striking, but not more than quaternions would effect, that I see.

By means of eight rotations through two-thirds of a circumference, around four symmetrically placed axes, together with unity, all distortions of a particle would be represented linearly. I have therefore thought of the nine-fold algebra thus resulting.



Paper 6: On the Algebra of Logic †1

Part I. †2 Syllogistic †P1

§1. Derivation of Logic

154. In order to gain a dear understanding of the origin of the various signs used in logical algebra and the reasons of the fundamental formulæ, we ought to begin by considering how logic itself arises.

155. Thinking, as cerebration, is no doubt subject to the general laws of nervous action.

156. When a group of nerves are stimulated, the ganglions with which the group is most intimately connected on the whole are thrown into an active state, which in turn usually occasions movements of the body. The stimulation continuing, the irritation spreads from ganglion to ganglion (usually increasing meantime). Soon, too, the parts first excited begin to show fatigue; and thus for a double reason the bodily activity is of a changing kind. When the stimulus is withdrawn, the excitement quickly subsides.

It results from these facts that when a nerve is affected, the reflex action, if it is not at first of the sort to remove the irritation, will change its character again and again until the irritation is removed; and then the action will cease.

157. Now, all vital processes tend to become easier on repetition. Along whatever path a nervous discharge has once taken place, in that path a new discharge is the more likely to take place.

Accordingly, when an irritation of the nerves is repeated, all the various actions which have taken place on previous similar occasions are the more likely to take place now, and those are most likely to take place which have most frequently taken place on those previous occasions. Now, the various. actions which did not remove the irritation may have previously sometimes been performed and sometimes not; but the action which removes the irritation must have always been performed, because the action must have every time continued until it was performed. Hence, a strong habit of responding to the given irritation in this particular way must quickly be established.

158. A habit so acquired may be transmitted by inheritance.

One of the most important of our habits is that one by virtue of which certain classes of stimuli throw us at first, at least, into a purely cerebral activity.

159. Very often it is not an outward sensation but only a fancy which starts the train of thought. In other words, the irritation instead of being peripheral is visceral. In such a case the activity has for the most part the same character; an inward action removes the inward excitation. A fancied conjuncture leads us to fancy an appropriate line of action. It is found that such events, though no external action takes. place, strongly contribute to the formation of habits of really acting in the fancied way when the fancied occasion really arises. †1

160. A cerebral habit of the highest kind, which will determine what we do in fancy as well as what we do in action, is called a belief. The representation to ourselves that we have a specified habit of this kind is called a judgment. A belief-habit in its development begins by being vague, special, and meagre; it becomes more precise, general, and full, without limit. The process of this development, so far as it takes place in the imagination, is called thought. A judgment is formed; and under the influence of a belief-habit this gives rise to a new judgment, indicating an addition to belief. Such a process is called an inference; the antecedent judgment is called the premiss; the consequent judgment, the conclusion; the habit of thought, which determined the passage from the one to the other (when formulated as a proposition), the leading principle. †P1

161. At the same time that this process of inference, or the spontaneous development of belief, is continually going on within us, fresh peripheral excitations are also continually creating new belief-habits. Thus, belief is partly determined by old beliefs and partly by new experience. Is there any law about the mode of the peripheral excitations? The logician maintains that there is, namely, that they are all adapted to an end, that of carrying belief, in the long run, toward certain predestinate conclusions which are the same for all men. This is the faith of the logician. This is the matter of fact, upon which all maxims of reasoning repose. In virtue of this fact, what is to be believed at last is independent of what has been believed hitherto, and therefore has the character of reality. Hence, if a given habit, considered as determining an inference, is of such a sort as to tend toward the final result, it is correct; otherwise not. Thus, inferences become divisible into the valid and the invalid; and thus logic takes its reason of existence.

§2. Syllogism and Dialogism †1

162. The general type of inference is

P

∴ C,

where ∴ is the sign of illation.

163. The passage from the premiss (or set of premisses) P to the conclusion C takes place according to a habit or rule active within us. All the inferences which that habit would determine when once the proper premisses were admitted, form a class. The habit is logically good provided it would never (or in the case of a probable inference, seldom) lead from a true premiss to a false conclusion; otherwise it is logically bad. That is, every possible case of the operation of a good habit would either be one in which the premiss was false or one in which the conclusion would be true; whereas, if a habit of inference is bad, there is a possible case in which the premiss would be true, while the conclusion was false. When we speak of a possible case, we conceive that from the general description of cases we have struck out all those kinds which we know how to describe in general terms but which we know never will occur; those that then remain, embracing all whose nonoccurrence we are not certain of, together with all those whose non-occurrence we cannot explain on any general principle, are called possible.

164. A habit of inference may be formulated in a proposition which shall state that every proposition c, related in a given general way to any true proposition p, is true. Such a proposition is called the leading principle of the class of inferences whose validity it implies. When the inference is first drawn, the leading principle is not present to the mind, but the habit it formulates is active in such a way that, upon contemplating the believed premiss, by a sort of perception the conclusion is judged to be true. †P1 Afterwards, when the inference is subjected to logical criticism, we make a new inference, of which one premiss is that leading principle of the former inference, according to which propositions related to one another in a certain way are fit to be premiss and conclusion of a valid inference, while another premiss is a fact of observation, namely, that the given relation does subsist between the premiss and conclusion of the inference under criticism; whence it is concluded that the inference was valid.

165. Logic supposes inferences not only to be drawn, but also to be subjected to criticism; and therefore we not only require the form P ∴ C to express an argument, but also a form, Pi ⤙ Ci, to express the truth of its leading principle. Here Pi denotes any one of the class of premisses, and Ci the corresponding conclusion. The symbol ⤙ is the copula, and signifies primarily that every state of things in which a proposition of the class Pi is true is a state of things in which the corresponding propositions of the class Ci are true. But logic also supposes some inferences to be invalid, and must have a form for denying the leading premiss [?principle]. This we shall write Pi ~⤙ Ci, a dash over any symbol signifying in our notation the negative of that symbol. †P1 Elec. ed.: A tilde (~) preceding the symbol signifies the negative in the electronic edition.

Thus, the form Pi ⤙ Ci implies

either, 1, that it is impossible that a premiss of the class Pi should be true,

or, 2, that every state of things in which Pi is true is a state of things in which the corresponding Ci is true.

The form Pi ~⤙ Ci implies

both, 1, that a premiss of the class Pi is possible,

and, 2, that among the possible cases of the truth of a Pi there is one in which the corresponding Ci is not true.

This acceptation of the copula differs from that of other systems of syllogistic in a manner which will be explained below in treating of the negative.

166. In the form of inference P ∴ C the leading principle is not expressed; and the inference might be justified on several separate principles. One of these, however, Pi ⤙ Ci, is the formulation of the habit which, in point of fact, has governed the inferences. This principle contains all that is necessary besides the premiss P to justify the conclusion. (It will generally assert more than is necessary.) We may, therefore, construct a new argument which shall have for its premisses the two propositions P and Pi ⤙ Ci taken together, and for its conclusion, C. This argument, no doubt, has, like every other, its leading principle, because the inference is governed by some habit; but yet the substance of the leading principle must already be contained implicitly in the premisses, because the proposition Pi ⤙ Ci contains by hypothesis all that is requisite to justify the inference of C from P. Such a leading principle, which contains no fact not implied or observable in the premisses, is termed a logical principle, and the argument it governs is termed a complete, in contradistinction to an incomplete, argument, or enthymeme.

The above will be made clear by an example. Let us begin with the enthymeme,

Enoch was a man,

∴ Enoch died.

The leading principle of this is, "All men die." Stating it, we get the complete argument,

All men die,

Enoch was a man;

∴ Enoch was to die.

The leading principle of this is nota notae est nota rei ipsius. Stating this as a premiss, we have the argument,

Nota notae est nota rei ipsius,

Mortality is a mark of humanity, which is a mark of Enoch;

∴ Mortality is a mark of Enoch.

But this very same principle of the nota notae is again active in the drawing of this last inference, so that the last state of the argument is no more complete than the last but one.

167. There is another way of rendering an argument complete, namely, instead of adding the leading principle Pi ⤙ Ci conjunctively to the premiss P, to form a new argument, we might add its denial disjunctively to the conclusion; thus,

P

∴ Either C or Pi ~⤙ Ci.

168. A logical principle is said to be an empty or merely formal proposition, because it can add nothing to the premisses of the argument it governs, although it is relevant; so that it implies no fact except such as is presupposed in all discourse, as we have seen in §1 that certain facts are implied. We may here distinguish between logical and extralogical validity; the former being that of a complete, the latter that of an incomplete argument. The term logical leading principle we may take to mean the principle which must be supposed true in order to sustain the logical validity of any argument. Such a principle states that among all the states of things which can be supposed without conflict with logical principles, those in which the premiss of the argument would be true would also be cases of the truth of the conclusion. Nothing more than this would be relevant to the logical leading principle, which is, therefore, perfectly determinate and not vague, as we have seen an extralogical leading principle to be.

169. A complete argument, with only one premiss, is called an immediate inference. Example: All crows are black birds; therefore, all crows are birds. If from the premiss of such an argument everything redundant is omitted, the state of things expressed in the premiss is the same as the state of things expressed in the conclusion, and only the form of expression is changed. Now, the logician does not undertake to enumerate all the ways of expressing facts: he supposes the facts to be already expressed in certain standard or canonical forms. But the equivalence between different ones of his own standard forms is of the highest importance to him, and thus certain immediate inferences play the great part in formal logic. Some of these will not be reciprocal inferences or logical equations, but the most important of them will have that character.

170. If one fact has such a relation to a different one that, if the former be true, the latter is necessarily or probably true, this relation constitutes a determinate fact; and therefore, since the leading principle of a complete argument involves no matter of fact (beyond those employed in all discourse), it follows that every complete and material (in opposition to a merely formal) argument must have at least two premisses.

171. From the doctrine of the leading principle it appears that if we have a valid and complete argument from more than one premiss, we may suppress all premisses but one and still have a valid but incomplete argument. This argument is justified by the suppressed premisses; hence, from these premisses alone we may infer that the conclusion would follow from the remaining premisses. In this way, then, the original argument

P Q R S T

∴C

is broken up into two, namely, 1st,

P Q R S

∴ T ⤙ C

and, 2d,

T ⤙ C

T

∴C.

By repeating this process, any argument may be broken up into arguments of two premisses each. A complete argument having two premisses is called a syllogism. †P1

172. An argument may also be broken up in a different way by substituting for the second constituent above, the form

T ⤙ C

∴ Either C or not T.

In this way, any argument may be resolved into arguments, each of which has one premiss and two alternative conclusions. Such an argument, when complete, may be called a dialogism.

§3. Forms of Propositions

173. In place of the two expressions A ⤙ B and B ⤙ A taken together we may write A = B; †P2 in place of the two expressions A ⤙ B and B ~⤙ A taken together we may write A < B or B > A; and in place of the two expressions A ~⤙ B and B ~⤙ A taken together [disjunctively] we may write A ≍ B. †1

174. De Morgan, in the remarkable memoir with which he opened his discussion of the syllogism (1846, p. 380, †2) has pointed out that we often carry on reasoning under an implied restriction as to what we shall consider as possible, which restriction, applying to the whole of what is said, need not be expressed. The total of all that we consider possible is called the universe of discourse, and may be very limited. One mode of limiting our universe is by considering only what actually occurs, so that everything which does not occur is regarded as impossible.

175. The forms A ⤙ B, or A implies B, and A ~⤙ B, or A does not imply B †3, embrace both hypothetical and categorical propositions. Thus, to say that all men are mortal is the same as to say that if any man possesses any character whatever then a mortal possesses that character. To say, 'if A, then B ' is obviously the same as to say that from A, B follows, logically or extralogically. By thus identifying the relation expressed by the copula with that of illation, we identify the proposition with the inference, and the term with the proposition. This identification, by means of which all that is found true of term, proposition, or inference is at once known to be true of all three, is a most important engine of reasoning, which we have gained by beginning with a consideration of the genesis of logic. †P1

176. Of the two forms A ⤙ B and A ~⤙ B, no doubt the former is the more primitive, in the sense that it is involved in the idea of reasoning, while the latter is only required in the criticism of reasoning. The two kinds of proposition are essentially different, and every attempt to reduce the latter to a special case of the former must fail. Boole †1 attempts to express 'some men are not mortal,' in the form 'whatever men have a certain unknown character v are not mortal.' But the propositions are not identical, for the latter does not imply that some men have that character v; and, accordingly, from Boole's proposition we may legitimately infer that 'whatever mortals have the unknown character v are not men'; †2 yet we cannot reason from 'some men are not mortal' to 'some mortals are not men.' †P2 On the other hand, we can rise to a more general form under which A ⤙ B and A ~⤙ B are both included. For this purpose we write A ~⤙ B in the form Ă ⤙ B̅, †3 where Ă is some-A and B̅ is not-B. This more general form is equivocal in so far as it is left undetermined whether the proposition would be true if the subject were impossible. When the subject is general this is the case, but when the subject is particular (i.e., is subject to the modification some) it is not. †1 The general form supposes merely inclusion of the subject under the predicate. The short curved mark over the letter in the subject shows that some part of the term denoted by that letter is the subject, and that that is asserted to be in possible existence.

177. The modification of the subject by the curved mark and of the predicate by the straight mark gives the old set of propositional forms, viz.:

A. ab Every a is b. Universal affirmative.
E. a No a is b. Universal negative.
I. ăb Some a is b. Particular affirmative.
O. ă Some a is not b. Particular negative.

178. There is, however, a difference between the senses in which these propositions are here taken and those which are traditional; namely, it is usually understood that affirmative propositions imply the existence of their subjects, while negative ones do not. Accordingly, it is said that there is an immediate inference from A to I and from E to O. But in the sense assumed in this paper, universal propositions do not, while particular propositions do, imply the existence of their subjects. The following figure illustrates the precise sense here assigned to the four forms A, E, I, O.

inline image

179. In the quadrant marked 1 there are lines which are all vertical; in the quadrant marked 2 some lines are vertical and some not; in quadrant 3 there are lines none of which are vertical; and in quadrant 4 there are no lines. Now, taking line as subject and vertical as predicate,

A is true of quadrants 1 and 4 and false of 2 and 3.

E is true of quadrants 3 and 4 and false of 1 and 2.

I is true of quadrants 1 and 2 and false of 3 and 4.

O is true of quadrants 2 and 3 and false of 1 and 4.

Hence, A and O precisely deny each other, and so do E and I. But any other pair of propositions may be either both true or both false or one true while the other is false. †1

180. De Morgan ("On the Syllogism," No. I., 1846, p. 381) has enlarged the system of propositional forms by applying the sign of negation which first appears in A ~⤙ B to the subject and predicate. He thus gets

A ⤙ B. Every A is B. †2 A is species of B. †3
A ~⤙ B. Some A is not B. A is exient of B.
A ⤙ B̅. No A is B. A is external of B.
A ~⤙ B̅. Some A is B. A is partient of B.
Ā ⤙ B. Everything is either A or B. A is complement of B.
Ā ~⤙ B. There is something besides A and B A is coinadequate of B
Ā ⤙ B̅. A includes all B. A is genus of B.
Ā ~⤙ B̅. A does not include all B. A is deficient of B.

De Morgan's table of the relations of these propositions must be modified to conform to the meanings here attached to ⤙ and to ~⤙.

181. We might confine ourselves to the two propositional forms S ⤙ P and S ~⤙ P. If we once go beyond this and adopt the form S ⤙ P̅, we must, for the sake of completeness, adopt the whole of De Morgan's system. But this system, as we shall see in the next section, is itself incomplete, and requires to complete it the admission of particularity in the predicate. This has already been attempted by Hamilton, with an incompetence which ought to be extraordinary. †4 I shall allude to this matter further on, but I shall not attempt to say how many forms of propositions there would be in the completed system. †P1

§4. The Algebra of the Copula

182. From the identity of the relation expressed by the copula with that of illation, springs an algebra. In the first place, this gives us

(1) xx

the principle of identity, which is thus seen to express that what we have hitherto believed we continue to believe, in the absence of any reason to the contrary. In the next place, this identification shows that the two inferences

(2) x
y and x
∴ z yz

are of the same validity. Hence we have

(3) {x ⤙ (yz)} = {y ⤙ (xz)}. †P2

183. From (1) we have

(xy) ⤙ (xy), whence by (2)

(4)
xy x
y

is a valid inference.

184. By (4), if x and xy are true y is true; and if y and yz are true z is true. Hence, the inference is valid

x xy yz

z.

By the principle of (2) this is the same as to say that

(5) xy yz
xz

is a valid inference. This is the canonical form of the syllogism Barbara. The statement of its validity has been called the dictum de omni, the nota notae, etc.; but it is best regarded, after De Morgan, †P1 as a statement that the relation signified by the copula is a transitive one. †P2 It may also be considered as implying that in place of the subject of a proposition of the form A ⤙ B, any subject of that subject may be substituted, and that in place of its predicate any predicate of that predicate may be substituted. †P3 The same principle may be algebraically conceived as a rule for the elimination of y from the two propositions xy and yz. †P4

185. It is needless to remark that any letters may be substituted for x, y, z; and that, therefore, , , , some or all, may be substituted. Nevertheless, after these purely extrinsic changes have been made, the argument is no longer called Barbara, but is said to be some other universal mood of the first figure. There are evidently eight such moods.

186. From (5) we have, by (2), these two forms of valid immediate inference:

(6) S ⤙ P
∴ (x ⤙ S) ⤙ (x ⤙ P)

and

(7) S ⤙ P
∴ (P ⤙ x) ⤙ (S ⤙ x).

The latter may be termed the inference of contraposition. †1

187. From the transitiveness of the copula, the following inference is valid:

(S ⤙ M) ⤙ (S ⤙ P)

(S ⤙ P) ⤙ x

∴ (S ⤙ M) ⤙ x.

But, by (6), from (M ⤙ P) we can infer the first premiss immediately; hence the inference is valid

(8)

M ⤙ P

(S ⤙ P) ⤙ x

∴ (S ⤙ M) ⤙ x.

This may be called the minor indirect syllogism. The following is an example:

All men are mortal,

If Enoch and Elijah were mortal, the Bible errs;

∴ If Enoch and Elijah were men, the Bible errs.

188. Again we may start with this syllogism in Barbara

(M ⤙ P) ⤙ (S ⤙ P),

(S ⤙ P) ⤙ x;

∴ (M ⤙ P) ⤙ x.

But by the principle of contraposition (7), the first premiss immediately follows from (S ⤙ M), so that we have the inference valid

(9)

S ⤙ M,

(S ⤙ P) ⤙ x;

∴ (M ⤙ P) ⤙ x.

This may be called the major indirect syllogism.

Example: All patriarchs are men,

If all patriarchs are mortal, the Bible errs;

∴ If all men are mortal, the Bible errs.

189. In the same way it might be shown that (6) justifies the syllogism

(10)

M ⤙ P,

x ⤙ (S ⤙ M);

x ⤙ (S ⤙ P).

And (7) justifies the inference

(11)

S ⤙ M,

x ⤙ (M ⤙ P);

x ⤙ (S ⤙ P).

But these are only slight modifications of Barbara.

190. In the form (10), x may denote a limited universe comprehending some cases of S. Then we have the syllogism

(12)

M ⤙ P,

S ~⤙ M̅;

∴ S ~⤙ P̅.

This is called Darii. A line might, of course, be drawn over the S. So, in the form (11), x may denote a limited universe comprehending some P. Then we have the syllogism

(13)

S ⤙ M,

M̅ ~⤙ P;

∴ S̅ ~⤙ P.

Here a line might be drawn over the P. But the forms (12) and (13) are deduced from (10) and (11) only by principles of interpretation which require demonstration.

191. On the other hand, if in the minor indirect syllogism (8), we put "what does not occur" for x, we have by definition

{(S ⤙ P) ⤙ x} = (S ~⤙ P)

and we then have

(14)

M ⤙ P,

S ~⤙ P;

∴ S ~⤙ M,

which is the syllogism Baroko. If a line is drawn over P, the syllogism is called Festino; and by other negations eight essentially identical forms are obtained, which are called minor-particular moods of the second figure. †P1 In the same way the major indirect syllogism (9) affords the form

(15)

S ⤙ M,

S ~⤙ P;

∴ M ~⤙ P.

This form is called Bocardo. If P is negatived, it is called Disamis. Other negations give the eight major-particular moods of the third figure.

192. We have seen that S ~⤙ P is of the form (S ⤙ P) ⤙ x. Put A for S ⤙ P, and we find that Ā is of the form A ⤙ x. Then the principle of contraposition (7) gives the immediate inference

(16)

S ⤙ P

∴ P̅ ⤙ S̅.

Applying this to the universal moods of the first figure justifies six moods. These are two in the second figure,

x zy x (Camestres)
zy ;

two in the third figure,

yx z z
yx ;

and two others which are said to be in the fourth figure,

xy yz
x z .

But the negative has two other properties not yet taken into account. These are

(17)

x

or x is not not-X, which is called the principle of contradiction; and

(18)

x̄̄x

or what is not not-X is x, which is called the principle of excluded middle. †1

193. By (17) and (16) we have the immediate inference

(19)

S ⤙ P̅

∴ P ⤙ S̅

which is called the conversion of E. By (18) and (16) we have

(20)

S̅ ⤙ P

∴ P̅ ⤙ S.

By (17), (18), and (16), we have

(21)

S̅ ⤙ P̅

∴ P ⤙ S.

Each of the inferences (19), (20), (21), justifies six universal syllogisms; namely, two in each of the figures, second, third, and fourth. The result is that each of these figures has eight universal moods; two depending only on the principle that Ā is of the form A ⤙ x, †P1 two depending also on the principle of contradiction, two on the principle of excluded middle, and two on all three principles conjoined.

194. The same formulæ (16), (19), (20), (21), applied to the minor-particular moods of the second figure, will give eight minor-particular moods of the first figure; and applied to the major-particular moods of the third figure, will give eight major-particular moods of the first figure. †P2

195. The principle of contradiction in the form (19) may be further transformed thus:

(22)

If (P ∴ C̅) is valid, then (C ∴ P̅) is valid. †1

Applying this to the minor-particular moods of the first figure, will give eight minor-particular moods of the third figure; and applying it to the major-particular moods of the first figure will give eight major-particular moods of the second figure.

It is very noticeable that the corresponding formula,

(23)

If (P̅ ∴ C) is valid, then (C̅ ∴ P) is valid, †1

has no application in the existing syllogistic, because there are no syllogisms having a particular premiss and universal conclusion. In the same way, in the Aristotelian system an affirmative conclusion cannot be drawn from negative premisses, the reason being that negation is only applied to the predicate. So in De Morgan's system the subject only is made particular, not the predicate.

196. In order to develop a system of propositions in which the predicate shall be modified in the same way in which the subject is modified in particular propositions, we should consider that to say S ⤙ P is the same as to say (S ~⤙ x) ⤙ (P ~⤙ x), whatever x may be. That

(S ⤙ P) ⤙ {(S ~⤙ x) ⤙ (P ~⤙ x)}

follows at once from Bokardo (15) by means of (2). Moreover, since Ā may be put in the form A ⤙ x, it follows that ~Ā may be put in the form A ~⤙ x, so that by the principles of contradiction and excluded middle, A may be put in the form A ~⤙ x. On the other hand, to say S ~⤙ P̅ is the same as to say (S ⤙ ) ⤙ (P ~⤙ x), whatever x may be; for

(S ~⤙ P̅) ⤙ {(S ⤙ ~x) ⤙ (P ~⤙ x)}

is the principle of Ferison, a valid syllogism of the third figure; and if for x we put S̅, we have

(S ⤙ ~S̅) ⤙ (P ~⤙ S̅),

which is the same as to say that P ~⤙ S̅ is true if the principle of contradiction is true. So that it follows that P ~⤙ S̅ is S ~⤙ P̅ from the principle of contradiction. Comparing

S ⤙ P or (S ~⤙ x) ⤙ (P ~⤙ x)

with

S ~⤙ P̅ or (S ⤙ ) ⤙ (P ~⤙ x),

we see that they differ by a modification of the subject. Denoting this by a short curve over the subject, we may write S̆ ⤙ P for S ~⤙ P̅. We see then that while for A we may write A ~⤙ x, where x is anything whatever, so for Ă we may write A ⤙ . If we attach a similar modification to the predicate also, we have

S̆ ⤙ P̆ or (S ⤙ ) ⤙ (P ⤙ ),

which is the same as to say that you can find an S which is any P you please. We thus have

(24)

(S ⤙ P) ⤙ (P̆⤙ S̆),

a formula of contraposition, similar to (16).

It is obvious that

(25)

(S̆ ⤙ P) ⤙ (P̆ ⤙ S);

for, negating both propositions, this becomes, by (16),

(P ⤙ S̅) ⤙ (S ⤙ P̅),

which is (19). The inference justified by (25) is called the conversion of I. From (25) we infer

(26)

$x, †1

which may be called the principle of particularity. This is obviously true, because the modification of particularity only consists in changing (A ~⤙ x) to (A ⤙ ~x), which is the same as negating the copula and predicate, and a repetition of this will evidently give the first expression again. For the same reason we have

(27)

x ⤙ $, †1

which may be called the principle of individuality. This gives

(28)

(S ⤙ P̆) ⤙ (P ⤙ S̆),

and (26) and (27) together give

(S̆ ⤙ P̆) ⤙ (P ⤙ S).

It is doubtful whether the proposition S ⤙ P̆ ought to be interpreted as signifying that S and P are one sole individual, or that there is something besides S and P. I here leave this branch of the subject in an unfinished state.

197. Corresponding to the formulæ which we have obtained by the principle (2) are an equal number obtained by the following principle:

(2') The inference

x

∴ Either y or z

has the same validity as

x ~⤙ y

z

[Add the formula (3') {(x ~⤙ y) ~⤙ z} = {(x ~⤙ z) ~⤙ y}.—1880.]

From (1) we have

(x ~⤙ y) ⤙ (x ~⤙ y),

whence, by (2),

(4')

x

∴ Either (xy) or y. †1

This gives

x

∴ Either x ~⤙ y or y ~⤙ z or z.

Then, by (2),

(5')

x ~⤙ z

x ~⤙ y or y ~⤙ z,

which is the canonical form of dialogism. The minor indirect dialogism is

(8')

x ~⤙ (M ~⤙ P)

∴ Either x ~⤙ (S ~⤙ P) or S ~⤙ M.

The major indirect dialogism is

x ~⤙ (S ~⤙ M)

∴ Either x ~⤙ (S ~⤙ P) or M ~⤙ P.

We have also

(12')

(S ~⤙ P) ~⤙ x

∴ Either (S ~⤙ M) or (M ~⤙ P) ~⤙ x

and

(13') (S ~⤙ P) ~⤙ x

∴ Either (M ~⤙ P) or (S ~⤙ M) ~⤙ x.

We have A of the form x ~⤙ Ā. And we have the inferences

S ~⤙ P S ~⤙ P̅ S̅ ~⤙ P S̅ ~⤙ P̅
∴P̅ ~⤙ S̅. ∴P ~⤙ S̅. ∴P̅ ~⤙ S. ∴P ~⤙ S.

PART II. The Logic of Non-Relative Terms

§1. The Internal Multiplication and the Addition of Logic

198. We have seen that the inference

x and y

z

is of the same validity with the inference

x

∴ Either or z

and the inference

x

∴ Either y or z,

with the inference

x and

z.

In like manner,

xy

is equivalent to

(The possible) ⤙ Either or y,

and to

x which is ⤙ (The impossible).

To express this algebraically, we need, in the first place, symbols for the two terms of second intention, the possible and the impossible. Let ∞ and 0 be the terms; then we have the definitions

(1)

x ⤙ ∞ 0 ⤙ x

whatever x may be. †P1

199. We need also two operations which may be called non-relative addition and multiplication. They are defined as follows: †P2

(2) If ax and bx,
then a + bx
If xa and xb
then xa × b;
and conversely and conversely,
(3) if a + bx,
then ax and bx.
if xa × b,
then xa and xb.

200. From these definitions we at once deduce the following formulæ:

(4) A. †P1 aa + b a × ba (Peirce, 1870) †P2
ba + b a × bb.

These are proved by substituting a + b and a X b for x in (3).

(5) B. x = x + x x × x = x (Jevons, 1864).

By substituting x for a and b in (2), we get

x + xx xx X x;

and, by (4),

(6) xx + x x × xx.
C. a + b = b + a a × b = b × a (Boole, Jevons).

These formulæ are examples of the commutative principle. From (4) and (2),

b + aa + b a × bb × a

and interchanging a and b we get the reciprocal inclusion implied in (6).

(7) D. (a+b)+c = a+(b+c) a× (b×c) = (a×bc (Boole, Jevons).

These are cases of the associative principle. By (4), cb+c and b×cc; also b+ca+(b+c) and a×(b×c) ⤙ b×c; so that ca+(b+c) and a×(b×c) ⤙ c. In the same way, ba+(b+c) and a×(b×c) ⤙ b, and, by (4), aa+(b+c) and a×(b×c) ⤙ a. Hence, by (2), a+ba+(b+c) and a×(b×c) ⤙ a×b. And, again by (2), (a+b)+ca+(b+c) and a×(b×c) ⤙ (a×bc. In a similar way we should prove the converse propositions to these and so establish (7).

(8)

E. (a+bc = (a×c)+(b×c) (a×b)+c = (a+c)×(b+c). †P3

These are cases of the distributive principle. They are easily proved by (4) and (2), but the proof is too tedious to give. †1

(9)

F. (a+b)+c = (a+c)+(b+c) (a×bc = (a×c)×(b×c).

These are other cases of the distributive principle. They are proved by (5), (6) and (7). These formulæ, which have hitherto escaped notice, are not without interest.

(10)

G. a+(a×b) = a a×(a+b) = a (Grassmann, Schröder).
By (4), aa+(a×b) a×(a+b) ⤙ a.

Again, by (4), (a×b) ⤙ a and aa+b; hence, by (2)

a+(a×b) ⤙ a aa×(a+b).

(11)

H. (a+ba) = (baXb).

This proposition is a transformation of Schröder's two propositions 21 (p. 25), one of which was given by Grassmann. By (3)

(a+ba) ⤙ (ba) (baXb) ⤙ (ba).
Hence, since bb, aa

we have, by (2),

(12)

(a + ba) ⤙ (ba × b) (ba × b) ⤙ (a + ba)

I.
(ab) × (xy) ⤙ (a + xb + y)
(ab) × (xy) ⤙ (a × xby)

(Peirce, 1970).

Readily proved from (2) and (4).

(13)

J. (ab+x)×(a×xb) = (ab).

This is a generalization of a theorem by Grassmann. In stating it, he erroneously unites the first two propositions by + instead of ×. By (12), (5), and (8),

(ab+x) ⤙ {a ⤙ (a×b)+(a×x)}

(a×xb) ⤙ {(a+b)×(x+b) ⤙ b}.

But by (4)

aa+b a×bb.

Hence, by (2), it is doubly proved that

(ab+x)×(a×xb) ⤙ (ab).

The demonstration of the converse is obvious.

We have immediately, from (2) and (3),

(14)

K. (a+bc) = (ac)×(bc)
(ca×b) = (ca)×(cb) (14)

(15)

L. (ca+b) = Σ{(pa)×(qb)} where p+q = c
(a×bc) = Σ{(ap)×(bq)} where c = p×q.

The propositions (15) are new. By (12)

{(p ⤙ a)×(q ⤙ b)} ⤙ (c ⤙ a+b) where p+q = c

{(ap)×(bq)} ⤙ (a×bc) where c = p×q.

and, since these are true for any set of values of p and q, we have by (2)

Σ{(pa)×(qb)} ⤙ (ca+b), where p+q = c.

Σ{(ap)×(bq)} ⤙ (a×bc), where c = p×q.

By (4) and (8), we have

(ca+b) ⤙ {(a×c)+(b×c) = c}

(a×bc) ⤙ {(c+a)×(c+b) = c}.

Hence, putting

a×c = p b×c = q, where p+q = c
a+c = p b+c = q, where p×q = c,

we have

(ca+b) ⤙ (pa)×(qb), where p+q = c

(a×bc) ⤙ (ap)×(bq), where c = p×q,

whence, by (4)

(ca+b) ⤙ Σ{(pa)×(qb)} where p+q = c

(a×bc) ⤙ Σ{(ap)×(bq)} where c = p×q.

A formula analogous to (15) will be found below (35).

201. From (1) and (2) and (4) we have

(16)

x+0 = x x = x×∞.

From (1) and (4),

(17)

x+∞ = ∞ 0 = x×0.

The definition of the negative has as we have seen three clauses: first, that ā is of the form ax; second, a ⤙ ~ā; third, ~~a ⤙ a.

From the first we have that if

c a

b

is valid, then

c

ā

is valid. Or

(18)

(c×ab) ⤙ (c×ā). †1

Also, that if

b

∴ Either c or a

is valid, then

ā

∴ Either c or

is valid; or

(19)

(bc+a) ⤙ (āc+).

Combining (18) and (19), we have

(20)

(a×bc+d) = (a×c+).

By the principles of contradiction and excluded middle, this gives

(21)

(a×c+) ⤙ (a×bc+d).

Thus the formula

(22)

(a×bc+d) = (a×c+)

embodies the essence of the negative.

202. If in (22) we put, first, a = d b = c = 0, and then a = d = ∞ b = c, we have from the formula of identity

(23)

a×ā = 0 a+ā = ∞.

We have

(24)

p = (p×x)+(p×) p = (p+x)×(p+),

by the distributive principle and (23). If we write

i = p+(a×) j = p+(b×x) k = p×(c+x) l = p×(d+),

we equally have

(25)

p = (i×x)+(j×) p = (l+x)×(k+).

Now p may be a function of x, and such values may perhaps be assigned to a, b, c, d, that i, j, k, l, shall be free from x. It is obvious that if the function results from any complication of the operations + and ×, this is the case. Supposing, then, i, j, k, l, to be constant, we have, putting successively, ∞, and 0, for x.

φ∞ = i = i = k

φ0 = j = l

so that

(26)

φx = (φ∞×x)+(φ0×) φx = (φ0+x)×(φ∞+).

The first of these formulæ was given by Boole for his addition. I showed (1867) †1 that both hold for the modified addition. These formulæ are real analogues of mathematical developments; but practically they are not convenient. Their connection suggests the general formula

(27)

(a+x)×(b+) = (a×)+(b×x)

a formula of frequent utility.

The distributive principle and (3) applied to (26) give

(28)

φ0 × φ∞ ⤙ φx φx ⤙ φ∞ + φ0.

Hence

(29)

x=0) ⤙ (φ0×φ∞=0) (φx=∞)⤙(φ0+φ∞=∞).

Boole gave the former, and I (1867) †2 the latter. These formulæ are not convenient for elimination.

203. The following formulæ (probably given by De Morgan) are of great importance:

(30)

~(a×b) = ā + ~(a+b) = ā × . †3

By (23)

(a×b)x~(a×b) ⤙ 0 ∞ ⤙ (a+b)+~(a+b),

whence by (22) and the associative principle

b×~(a×b) ⤙ ā āb+~(a+b)

~(a×b) ⤙ ā+ ā×a+b.

By (4) and (22)

ā ⤙ ~(a×b) ~(a+b) ⤙ ā

⤙ ~(a×b) ~(a+b) ⤙ ,

whence by (2)

ā+ ⤙ ~(a+b) †4 ~(a+b) ⤙ ā×.

The application of (22) gives from (11)

(31)

(b ~⤙ a×b) = (a+b ~⤙ a);

from (12)

(32)

(a+x ~⤙ b+y) ⤙ (a ~⤙ b)+(x ~⤙ y)

(a×x ~⤙ b×y) ⤙ (a ~⤙ b)+(x ~⤙ y);

from (13)

(33)

(a ~⤙ b) = (a ~⤙ b+x)+(aXx ~⤙ b);

from (14)

(34)

(a+b ~⤙ c) = (a ~⤙ c)+(b ~⤙ c); (c ~⤙ aXb) = (c ~⤙ a)+(c ~⤙ b);

from (15)

(35)

(c ~⤙ a+b) = Π{(p ~⤙ a)+(q ~⤙ b)} where p+q = c

(aXb ~⤙ c) = Π{(a ~⤙ p)+(b ~⤙ q)} where pXq=c;

from (22)

(36)

(aXb ~⤙ c+d) = (aX ~⤙ c+). (36)

§2. The Resolution of Problems in Non-Relative Logic

204. Four different algebraic methods of solving problems in the logic of non-relative terms have already been proposed by Boole, Jevons, Schröder, and McColl. I propose here a fifth method which perhaps is simpler and certainly is more natural than any of the others. It involves the following processes:

205. First Process. Express all the premisses with the copulas ⤙ and ~⤙, remembering that A = B is the same as A ⤙ B and B ⤙ A.

206. Second Process. Separate every predicate into as many factors and every subject into as many aggregant terms as is possible without increasing the number of different letters used in any subject or predicate.

207. An expression might be separated into such factors or aggregants (let us term them prime factors and ultimate aggregants) by one or other of these formulæ:

φx = (φ∞ × x)+(φ0 × )

φx = (φ∞ + )×(φ0 + x)

But the easiest method is this. To separate an expression into its inline image take any inline image of all the different letters of the expression, each taken either positively or negatively (that is, with a dash over it). By means of the fundamental formulæ

X × Y ⤙ Y ⤙ Y + Z

examine whether the inline image taken is a inline image of every inline image of the given expression. If so, it is a inline image of that expression; otherwise not. Proceed in thus way until as many inline image have been found as the expression possesses. This number is found in the case of a inline image of letters, as follows. Let m be the number of different letters in the expression (a letter and its negative not being considered different); let n be the total number of letters whether the same or different, and let p be the number of inline image. Then the number of inline image is

2m + n - m p - p.

208. For example, let it be required to separate x+(y×z) into its prime factors. Here m = 3, n = 3, p = 2. Hence the number of factors is three. Trying x+y+z, we have

xx+y+z y×zx+y+z,

so that this is a factor. Trying x+y+, we have

xx+y+ y×zx+y+,

so that this is also a factor. It is, also, obvious that x++z is the third factor. Accordingly,

x+(y×z) = (x+y+z)×(x+y+)×(x++z).

Again, let us develop the expression

(ā+b+c)×(a++)×(a+b+c)

Here m = 3, n = 9, p = 3; so that the number of ultimate aggregants is five. Of the eight possible products of three letters, then, only three are excluded, namely: (a××), (ā×b×c) and (ā××). We have, then,

(ā+b+c)×(a++)×(a+b+c) =
(a×b×c)+(a×b×)+(a××c)+(ā×b×)+(ā××c).

209. Third Process. Separate all complex propositions into simple ones by means of the following formulæ from the definitions of + and ×:

(X+Y ⤙ Z) †1 = (X ⤙ Z) × (Y ⤙ Z)

(X ⤙ Y × Z) †2 = (X ⤙ Y) × (X ⤙ Z)

(X+Y ~⤙ Z) †3 = (X ~⤙ Z)+(Y ~⤙ Z)

(X ~⤙ Y × Z) †4 = (X ~⤙ Y)+(X ~⤙ Z).

In practice, the first three operations will generally be performed off-hand in writing down the premisses.

210. Fourth Process. If we have given two propositions, one of one of the forms

ab+x a×b,

and the other of one of the forms

cd+ c×xd,

we may, by the transitiveness of the copula, eliminate x, and so obtain

a×cb+d.

211. Fifth Process. We may transpose any term from subject to predicate or the reverse, by changing it from positive to negative or the reverse, and at the same time its mode of connection from addition to multiplication or the reverse. Thus,

(x×yz) = (x+z).

We may, in this way, obtain all the subjects and predicates of any letter; or we may bring all the letters into the subject, leaving the predicate 0, or all into the predicate, leaving the subject ∞.

212. Sixth Process. Any number of propositions having a common inline image are, taken together, equivalent to their inline image.

213. As an example of this method, we may consider a well-known problem given by Boole. †1 The data are

×v×(y×+×w)

×x×w ⤙ (y×z)+(×)

(x×y)+(v×x×) = (z×)+(×w).

The quæsita are: first, to find those predicates of x which involve only y, z, and w; second, to find any relations which may be implied between y, z, w; third, to find the predicates of y; fourth, to find any relation which may be implied between x, z, and w. By the first three processes, mentally performed, we resolve the premisses as follows: the first into

×v

×y+w

×+;

the second into

×x×wy+

×x×w+z;

the third into

x×yz+w

x×y+

v×x×z+w

v×x×+

z×x

×wv+y

z×x

×wv+y.

We must first eliminate v, about which we want to know nothing. We have, on the one hand, the propositions

v×x×z+w

v×x×+;

and, on the other, the propositions

×v

×x×wy+

×x×w+z

z×v+y

×wv+y.

The conclusions from these propositions are obtained by taking one from each set, multiplying their subjects, adding their predicates, and omitting v. The result will be a merely empty proposition if the same letter in the same quality as to being positive or negative be found in the subject and in the predicate, or if it be found twice with opposite qualities either in the subject or in the predicate. Thus, it will be useless to combine the proposition v×x×z+w with any which contains , y, z, or w, in the subject. But all of the second set do this, so that nothing can be concluded from this proposition. So it will be useless to combine v×x×+ with any which contains , y, , in the subject, or z in the predicate. This excludes every proposition of the second set except ×x×wy+, which, combined with the proposition under discussion, gives

x×wy++

or x×wy+,

which is therefore to be used in place of all the premisses containing v.

One of the other propositions, namely, ×+ is obviously contained in another, namely: ×wx. Rejecting it, our premisses are reduced to six, namely:

×y+w

x×yz+w

x×y+

z×x

×wx

x×wy+

The second, third, and sixth of these give the predicates of x Their product is

x ⤙ (+z+w)×(++)×(y++)

or

xy×z× + y××w + ×z× + ××w + ××

or

xz× + ×w + ××.

To find whether any relation between y, z, and w can be obtained by the elimination of x, we find the subjects of x by combining the first, fourth, and fifth premisses. Thus we find

×× + z× + ×wx.

It is obvious that the conclusion from the last two propositions is a merely identical proposition, and therefore no independent relation is implied between y, z, and w.

To find the predicates of y we combine the second and third propositions. This gives

y ⤙ (+z+w)×(++)

or yx×z× + x××w + .

Two relations between x, z, and w are given in the premisses, namely: z×x and ×wx. To find whether any other is implied, we eliminate y between the above proposition and the first and sixth premisses. This gives

×x×z× + w +

x×wx×z× + + .

The first conclusion is empty. The second is equivalent to x×w, which is a third relation between x, z, and w.

Everything implied in the premisses in regard to the relations of x, y, z, w may be summed up in the proposition

∞ ⤙ x+z×w + y××.

Part III. The Logic of Relatives

§1. Individual and Simple Terms

214. Just as we had to begin the study of Logical Addition and Multiplication by considering ∞ and 0, terms which might have been introduced under the Algebra of the Copula, being defined in terms of the copula only, without the use of + or ×, but which had not been there introduced, because they had no application there, so we have to begin the study of relatives by considering the doctrine of individuals and simples, — a doctrine which makes use only of the conceptions of nonrelative logic, but which is wholly without use in that part of the subject, while it is the very foundation of the conception of a relative, and the basis of the method of working with the algebra of relatives.

215. The germ of the correct theory of individuals and simples is to be found in Kant's Critic of the Pure Reason, "Appendix to the Transcendental Dialectic," where he lays it down as a regulative principle, that, if

ab b ~⤙ a,

then it is always possible to find such a term x, that

ax xb

x ~⤙ a b ~⤙ x.

Kant's distinction of regulative and constitutive principles is unsound, but this law of continuity, †1 as he calls it, must be accepted as a fact. The proof of it, which I have given elsewhere, †2 depends on the continuity of space, time, and the intensities of the qualities which enter into the definition of any term. If, for instance, we say that Europe, Asia, Africa and North America are continents, but not all the continents, there remains over only South America. But we may distinguish between South America as it now exists and South America in former geological times; we may, therefore, take x as including Europe, Asia, Africa, North America, and South America as it exists now, and every x is a continent, but not every continent is x.

216. Just as in mathematics we speak of infinitesimals and infinites, which are fictitious limits of continuous quantity, and every statement involving these expressions has its interpretation in the doctrine of limits, †3 so in logic we may define an individual, A, as such a term that

A ~⤙ 0,

but such that if

x < A

then

x ⤙ 0.

And in the same way, we may define the simple, α, as such a term that

∞ ~⤙ α,

but such that if

α < x

then

∞ ⤙ x.

The individual and the simple, as here defined, are ideal limits, and every statement about either is to be interpreted by the doctrine of limits.

217. Every term may be conceived as a limitless logical sum of individuals, †1 or as a limitless logical product of simples; thus,

α = A1+A2+A3+A4+A5+ etc.

ᾱ = Ā1×Ā2×Ā3×Ā4×Ā5× etc.

It will be seen that a simple is the negative of an individual.

§2. Relatives

218. A relative is a term whose definition describes what sort of a system of objects that is whose first member (which is termed the relate) is denoted by the term; and names for the other members of the system (which are termed the correlates) are usually appended to limit the denotation still further. In these systems the order of the members is essential; so that (A, B, C) and (A, C, B) are different systems. As an example of a relative, take 'buyer of — for — from'; we may append to this three correlates, thus, ' buyer of every horse of a certain description in the market for a good price from its owner.'

219. A relative of only one correlate, so that the system it supposes is a pair, may be called a dual relative; a relative of more than one correlate may be called plural. A nonrelative term may be called a term of singular reference.

220. Every relative, like every term of singular reference, is general; its definition describes a system in general terms; and, as general, it may be conceived either as a logical sum of individual relatives, or as a logical product of simple relatives. †P1 An individual relative refers to a system all the members of which are individual. The expressions

(A:B) (A:B:C)

may denote individual relatives. Taking dual individual relatives, for instance, we may arrange them all in an infinite block, thus,

A:A A:B A:C A:D A:E etc.
B:A B:B B:C B:D B:E etc.
C:A C:B C:C CD C:E etc.
D:A D:B D:C D:D D:E etc.
E:A E:B E:C E:D E:E etc.
etc. etc. etc. etc. etc.

In the same way, triple individual relatives may be arranged in a cube, and so forth. The logical sum of all the relatives in this infinite block will be the relative universe, ∞, where

x ⤙ ∞,

whatever dual relative x may be. It is needless to distinguish the dual universe, the triple universe, etc., because, by adding a perfectly indefinite additional member to the system, a dual relative may be converted into a triple relative, etc. Thus, for lover of a woman, we may write lover of a woman coexisting with anything. In the same way, a term of single reference is equivalent to a relative with an indefinite correlate; thus, woman is equivalent to woman coexisting with anything. Thus, we shall have

A = A:A+A:B+A:C+A:D+A:E+etc.

A:B = A:B:A+A:B:B+A:B:C+A:B:D+etc.

221. From the definition of a simple term given in the last section, it follows that every simple relative is the negative of an individual term. But while in non-relative logic negation only divides the universe into two parts, in relative logic the same operation divides the universe into 2n parts, where n is the number of objects in the system which the relative supposes; thus,

∞ = A+Ā

∞ = A:B+Ā:B+A:B̅+Ā:B̅

∞ = (A:B:C)+(Ā:B:C)+(A:B̅:C)+(A:B:C̅)

+ (Ā:B̅:C̅)+(A:B̅:C̅)+(Ā:B:C̅)+(Ā:B̅:C).

Here, we have

A = A:B+A:B̅; Ā = Ā:B+Ā:B̅;

A:B = A:B:C+A:B:C̅; A:B̅ = A:B̅:C+A:B̅:C̅;

Ā:B = Ā:B:C+Ā:B:C̅; Ā:B̅ = Ā:B̅:C+Ā:B̅:C̅.

It will be seen that a term which is individual when considered as n-fold is not so when considered as more than n-fold; but an n-fold term when made (m+n)-fold, is individual as to n members of the system, and indefinite as to m members.

222. Instead of considering the system of a relative as consisting of non-relative individuals, we may conceive of it as consisting of relative individuals. Thus, since

A = A:A+A:B+A:C+A:D+etc.,

we have

A:B = (A:A):B+(A:B):B+(A:C):B+(A:D):B+etc.

But

B = B:A+B:B+B:C+B:D+etc.;

so that

A:B = A:(B:A)+A:(B:B)+A:(B:C)+A:(B:D)+etc. †1

§3. Relatives Conneted by Transposition of Relate and Correlate

223. Connected with every dual relative, as

l = Σ(A:B) = Π(α:β), †2

is another which is called its converse,

k-l = Σ(B:A) = Π(β:α),

in which the relate and correlate are transposed. The converse, k, is itself a relative, being

k = Σ[(A:B):(B:A)];

that is, it is the first of any pair which embraces two individual dual relatives, each of which is the converse of the other. The converse of the converse is the relation itself, thus

k-k-l = l,

or say kk = 1.

We have also

k- = ~(k-l)

kΣ = Σk

kΠ = Πk.

224. In the case of triple relatives there are five transpositions possible. Thus, if

b = Σ[(A:B):C]

we may write

Ib = Σ[(B:A):C]

Jb = Σ[(A:C):B]

Kb = Σ[(C:B):A]

Lb = Σ[(C:A):B]

Mb = Σ[(B:C):A].

Here we have

LM †1 = ML = 1

II = JJ = KK = 1

IJ = JK = KI = L

JI = KJ = IK = M

IL = MI = J = KM = LK

JL = MJ = K = IM = LI

KL = MK = I = JM = LJ.

If we write a:b to express the operation of putting A in place of B in the original relative

b = Σ[(A:B):C]

then we have

I = a:b+b:a+c:c †2

J = a:a+b:c+c:b

K = a:c+b:b+c:a

L = a:b+b:c+c:a

M = a:c+b:a+c:b

1 = a:a+b:b+c:c.

Then we have

I+J+K = 1+L+M,

which does not imply

(I+J+K)l = (1+L+M)l.

In a similar way the n-fold relative will have (n! - 1) transposition-functions.

§4. Classification of relatives

225. Individual relatives are of one or other of the two forms

A:A A:B,

and simple relatives are negatives of one or other of these two forms.

226. The forms of general relatives are of infinite variety, but the following may be particularly noticed.

Relatives may be divided into those all whose individual aggregants are of the form A:A and those which contain individuals of the form A:B. The former may be called concurrents, †1 the latter opponents. Concurrents express a mere agreement among objects. Such, for instance, is the relative 'man that is —,' and a similar relative may be formed from any term of singular reference. We may denote such a relative by the symbol for the term of singular reference with a comma after it; thus (m,) will denote 'man that is —' if (m) denotes 'man.' In the same way a comma affixed to an n-fold relative will convert it into an (n+1)-fold relative. Thus, (l) being 'lover of —,' (l,) will be 'lover that is — of —.'

The negative of a concurrent relative will be one each of whose simple components is of the form ~(A:A), and the negative of an opponent relative will be one which has components of the form ~(A:B).

We may also divide relatives into those which contain individual aggregants of the form A: A and those which contain only aggregants of the form A: B. The former may be called self-relatives, †2 the latter alio-relatives. We also have negatives of self-relatives and negatives of alio-relatives.

227. These different classes have the following relations. Every negative of a concurrent and every alio-relative is both an opponent and the negative of a self-relative. Every concurrent and every negative of an alio-relative is both a self-relative and the negative of an opponent. There is only one relative which is both a concurrent and the negative of an alio-relative; this is 'identical with —.' There is only one relative which is at once an alio-relative and the negative of a concurrent; this is the negative of the last, namely, ' other than —.' †P1 The following pairs of classes are mutually exclusive, and divide all relatives between them:

Alio-relatives and self-relatives,

Concurrents and opponents,

Negatives of alio-relatives and negatives of self-relatives,

Negatives of concurrents and negatives of opponents.

No relative can be at once either an alio-relative or the negative of a concurrent, and at the same time either a concurrent or the negative of an alio-relative.

228. We may append to the symbol of any relative a semicolon to convert it into an alio-relative of a higher order. Thus (l;) will denote a 'lover of — that is not —.'

229. This completes the classification of dual relatives founded on the difference of the fundamental forms A:A and A:B. Similar considerations applied to triple relatives would give rise to a highly complicated development, inasmuch as here we have no less than five fundamental forms of individuals, namely,

(A:A):A (A:A):B (A:B):A (B:A):A (A:B):C.

The number of individual forms for the (n+2)-fold relative is

2+(2n-1) . 3 + 1/2! {(3n -1) - 2(2n - 1)} . 4 + 1/3! {(4n - 1)
- 3(3n - 1) + 3(2n - 1)} . 5 + 1/4! {(5n - 1) - 4(4n - 1)
+ 6(3n - 1) -4(2n - 1)} . 6 + 1/5!{(6n - 1) - 5(5n - 1)
+10(4n - 1) - 10(3n - 1) + 5(2n - 1)} . 7 + etc.

If this number be called fn, we have

Δnf0 = f(n-1)

f0 = 1.

The form of calculation is

1
2 1
3 3 2
15 10 7 5
52 37 27 20 15
203 151 114 87 67 52
where the diagonal line is copied number by number from the vertical line, as fast as the latter is computed. †P1

230. Relatives may also be classified according to the general amount of filling up of the above-mentioned block, cube, etc., they present. In the first place, we have such relatives in whose block, cube, etc., every line in a certain direction in which there is a single individual is completely filled up. Such relatives may be called complete in regard to the relate, or first, second, third, etc., correlate. The dual relatives which are equivalent to terms of singular reference are complete as to their correlate.

231. A relative may be incomplete with reference to a certain correlate or to its relate, and yet every individual of the universe may in some way enter into that correlate or relate. Such a relative may be called unlimited in reference to the correlate or relate in question. Thus, the relative

A:A+A:B+C:C+C:D+E:E+E:F+G:G+G:H + etc.

is unlimited as to its correlate. The negative of an unlimited relative will be unlimited unless the relative has as an integrant a relative which is complete with regard to every other relate and correlate than that with reference to which the given relative is unlimited.

232. A totally unlimited relative is one which is unlimited in reference to the relate and all the correlates. A totally unlimited relative in which each letter enters only once into the relate and once into the correlate is termed a substitution.

233. Certain classes of relatives are characterized by the occurrence or non-occurrence of certain individual aggregants related in a definite way to others which occur. A set of individual dual relatives each of which has for its relate the correlate of the last, the last of all being considered as preceding the first of all, may be called a cycle. If there are n individuals in the cycle it may be called a cycle of the nth order. For instance,

A:B B:C C:D D:E E:A

may be called the cycle of the fifth order. Now, if a certain relative be such that of any cycle of the nth order of which it contains any m terms, it also contains the remaining (n - m) terms, it may be called a cyclic relative of the nth order and mth degree. If, on the other hand, of any cycle of the nth order of which it contains m terms the remaining (n - m) are wanting, the relative may be called an anticyclic relative of the nth order and mth degree.

234. A cyclic relative of the first order and first degree contains all individual components of the form A:A. A cyclic relative of the second order and first degree is called an equiparant †1 in opposition to a disquiparant.

235. A highly important class of relatives is that of transitives; that is to say, those which for every two individual terms of the forms (A:B) and (B:C) also possess a term of the form (A:C). †2

§5. The Composition of Relatives

236. Suppose two relatives either individual or simple, and having the relate or correlate of the first identical with the relate or correlate of the second or of its negative. This pair of relatives will then be of one or other of sixteen forms, viz.:

(A:B)(B:C) ~(A:B)(B:C) (A:B)~(B:C) ~(A:B)~(B:C)
(A:B)(C:B) ~(A:B)(C:B) (A:B)~(C:B) ~(A:B)~(C:B)
(B:A)(B:C) ~(B:A)(B:C) (B:A)~(B:C) ~(B:A)~(B:C)
(B:A)(C:B) ~(B:A)(C:B) (B:A)~(C:B) ~(B:A)~(C:B)

Now we may conceive an operation upon any one of these sixteen pairs of relatives of such a nature that it will produce one or other of the four forms (A:C),~(A:C),(C:A),~(C:A). Thus, we shall have sixty-four operations in all.

237. We may symbolize them as follows: Let

A:B(|||)B:C = A:C;

that is, let (|||) signify such an operation that this formula necessarily holds. The three lines in the sign of this operation are to refer respectively to the first relative operated upon, the second relative operated upon, and to the result. When either of these lines is replaced by a hyphen (-), let the operation signified be such that the negative of the corresponding relative must be substituted in the above formula. Thus,

~(A:B)(-||)B:C = A:C.

In the same way, let a semicircle ($) signify that the converse of the corresponding relative is to be taken. The hyphen and the semicircle may be used together. If, then, we write the symbol of a relative with a semicircle or curve over it to denote the converse of that relative, we shall have, for example,

$(A:B)($||)B:C = A:C.

238. Then any combination of the relatives a and e, in this order, is equivalent to others formed from this by making any of the following changes:

First. Putting a straight or curved mark over a and changing the first mark of the sign of operation in the corresponding way; that is,

for ă, from | to $ or from - to $- or conversely,

for ā, from | to - or from $ to $- or conversely,

for ā̆, from | to $- or from - to $ or conversely.

Second. Making similar simultaneous modifications of e and of the second mark.

Third. Changing the third mark from | to - or from $ to $- or conversely, and simultaneously writing the mark of negation over the whole expression.

Fourth. Changing the third mark from | to $ or from - to $- or conversely, and interchanging a and e and also the first and second marks.

239. We have thus far defined the effect of the sixty-four operations when certain members of the individual relatives operated upon are identical. When these members are not identical, we may suppose either that the operation ||| produces either the first or second relative or 0. We cannot suppose that it produces ∞ for a reason which will appear further on. Let us elect the formula

A:B(|||)C:D = 0.

The other excluded operations will be included in a certain manner, as we shall see below. From this formula, by means of the rules of equivalence, it will follow that all operations in whose symbol there is no hyphen in the third place will also give 0 in like circumstances, while all others will give 0̅ or ∞.

240. We have thus far only defined the effect of the sixty-four operations upon individual or simple terms. To extend the definitions to other cases, let us suppose first that the rules of equivalence are generally valid, and second, that

If ab and cd then a(|||)cb(|||)d

or

(ab)×(cd) ⤙ {(a(|||)cb(|||)d}.

Then, this rule will hold good in all operations in whose symbols the first and second places agree with the third in respect to having or not having hyphens. For operations, in whose symbols the inline image mark disagrees with the third in this respect we must write inline image instead of inline image in this rule. Thus, the sixty-four operations are divisible into four classes according to which one of the four rules so produced they follow.

241. It now appears that only the hyphens and not the curved marks are of significance in reference to the rule which an operation follows. Let us accordingly reject all operations whose symbols contain curved marks, and there remain only eight. For these eight the following formulæ hold:

A:B(|||)B:C = A:C A:B(||-)B:C = ~(A:C)
~(A:B)(-||)B:C = A:C ~(A:B)(-|-)B:C = ~(A:C)
A:B(|-|)~(B:C) = A:C A:B(|—)~(B:C) = ~(A:C)
~(A:B)(—|)~(B:C) = A:C ~(A:B(—-)~(B:C) = ~(A:C)
A:B(|||)C:D = 0 A:B(||-)C:D = ∞
~(A:B)(-||)C:D = 0 ~(A:B)(-|-)C:D = ∞
A:B(|-|)~(C:D) = 0 A:B(|—)~(C:D) = ∞
~(A:B)(—|)~(C:D) = 0 ~(A:B)(—-)~(C:D) = ∞
(ab)X(cd) ⤙ {a (|||) cb (|||) d}
(ab)X(cd) ⤙ {a (—-) cb (—-) d}
(ba)X(cd) ⤙ {a (-||) cb (-||) d}
(ba)X(cd) ⤙ {a (|—) cb (|—) d}
(ab)X(dc) ⤙ {a (|-|) cb (|-|) d}
(ab)X(dc) ⤙ {a (-|-) cb (-|-) d}
(ba)X(dc) ⤙ {a (—|) cb (—|) d}
(ba)X(dc) ⤙ {a (||-) cb (||-) d}

242. As it is inconvenient to consider so many as eight distinct operations, we may reject one-half of these so as to retain one under each of the four rules. We may reject all those whose symbols contain an odd number of hyphens (as being negative). We then retain four, to which we may assign the following names and symbols:

a (|||) e = ae Relative or external multiplication.

a (|—) e = ae Regressive involution.

a (-|-) e = ae Progressive involution.

a (—|) e = ae Transaddition. †P1

243. We have then the following table of equivalents, negatives, and converses: †P1

x $x̄ ae = ā ○ ē
ae = āē
ae = āē
a ○ e = āē
āe = aē
āe = a ○ ē
= ā ○ e
aē = āe
ĕă = ē̆ ○ ā̆
ĕă = ē̆ā̆
ĕă = ē̆ā̆
ĕ ○ ă = ē̆ ā̆
ē̆ā̆ = ē̆ā̆
ĕā̆ = ē̆ ○ ă
ē̆ă = ĕ ○ ă
ĕā̆ = ē̆ă

244. If l denote 'lover' and s 'servant,' then

ls denotes whatever is lover of a servant of —,
ls whatever is lover of every servant of—,
ls whatever is in every way (in which it loves at all) lover of a servant,
ls whatever is not a non-lover only of a servant of —
or whatever is not a lover of everything but servants of —
or whatever is some way a non-lover of some thing besides servants of—.

§6. Methods in the Algebra of Relatives

245. The universal method in this algebra is the method of limits. For certain letters are to be substituted an infinite sum of individuals or product of simples; whereupon certain transformations become possible which could not otherwise be effected.

246. The following theorems are indispensable for the application of this method:

First

lA:B = l(A:B) + kB̅.

Since B̅ is equivalent to the relative term which comprises all individual relatives whose relates are not B, so kB̅ may be conveniently used, as it is here, to express the aggregate of all individual relatives whose correlate is B̅. To prove this proposition, we observe that

lA:B = ~((A:B)).

Now (A:B) contains only individual relatives whose correlate is B, and of these it contains those which are not included in l(A:B). Hence the negative of (A:B) contains all individual relatives whose correlates are not B, together with all contained in l(A:B). Q. E. D.

Second

A:Bl = (A:B)l+Ā.

Here Ā is used to denote the aggregate of all individual relatives whose relates are not A. This proposition is proved like the last.

Third

~(A:Bl) = (A:B)+Ā.

This is evident from the second proposition, because

~(A:Bl) = (A:B)

Fourth

l~(A:B) = (A:B)+kB̅.

Another method of working with the algebra is by means of negations. This becomes quite indispensable when the operations are defined by negations, as in this paper.

247. To illustrate the use of these methods, let us investigate the relations of lb and lb to lb when l and b are totally unlimited relatives.

Write

l = Σi(Li:Mi) b = Σj(Bj:Cj).

Then, by the rules of the last section,

lbL:Mb lblB:C;

whence, by the second and third propositions above,

lb ⤙ (Li:Mi)b+L̄i, lbl(Bj:Cj)+kj.

But by the first rule of the last section

(Li:Mi)blb l(Bj:Cj) ⤙ lb;

hence,

lblb + L̄i lblb + kj.

There will be propositions like these for all the different values of i and j. Multiplying together all those of the several sets, we have

lblb + Πii lblb + Πjkj.

But

Πii = ~(ΣiLi) Πjkj = ~(ΣjkBj)

and since the relatives are unlimited,

ΣiLi = ∞ ΣjkBj = ∞

~(ΣiLi) = 0 ~(ΣjkBj) = 0;

hence

lblb lblb.

In the same way it is easy to show that, if the negatives of l and b are totally unlimited,

lblb lblb.

§7. The General Formulæ for Relatives

248. The principal formulæ of this algebra may be divided into distribution formulæ and association formulæ. The distribution formulæ are those which give the equivalent of a relative compounded with a sum or product of two relatives in such terms as to separate the latter two relatives. The association formulæ are those which give the equivalent of a relative A compounded with a compound of B and C in terms of a compound of A and B compounded with C.

249. I. DISTRIBUTION FORMULÆ

1. AFFIRMATIVE

i. Simple Formulæ

(a+b)c = ac+bc a(b+c) = ab+ac
(a×b)c = ac×bc ab+c = ab×ac
a+bc = ac×bc a(b×c) = ab×ac
(a×b)○c = (ac)+(bc) a○(b×c) = (ab)+(ac)

ii. Developments

(a×b)c = Πp{a(c×p)+b(c×)}
(a+b)c = Σp{ac×p×bc×}
(a×b)c = Σp{a(c+pb(c+)}
(a+b)○c = Πp{a○(c+p)+b○(c+)}
a(b×c) = Πp{(a×p)b+(a×)c}
ab×c = Σp{(a+p)b×(a+)c}
a(b+c) = Σp{a×pb×a×c}
a○(b+c) = Πp{(a+p)○b+(a+)○c}

2. NEGATIVE

i. Simple Formulæ

~((a+b)c) = ~(ac)×~(bc) ~(a(b+c)) = ~(ab)×~(ac)
~((a×b)c) = ~(ac)+~(bc) ~(ab+c) = ~(ab)+~(ac)
~(a+bc) = ~(ac)+~(bc) ~(a(b×c)) = ~(ab)+~(ac)
~((a×b)○c) = ~(ac)×~(bc) ~(a○(b×c)) = ~(ab)×~(ac)

ii. Developments

~((a×b)c) = Σp{~(a(c×p)) × ~(b(c×))}
~((a+b)c) = Πp{~(ac×p) + ~(bc×)}
~((a×b)c) = Πp{~(a(c+p)) + ~(b(c+))}
~((a+b)○c) = Σp{~(a○(c+p)) × ~(b○(c+))}
~(a(b×c)) = Σp{~((a×p)b) × ~((a×)c)}
~(ab×c) = Πp{~((a+p)b) + ~((a+)c)}
~(a(b+c)) = Πp{~(a×pb) + ~(a×c)}
~(a○(b+c)) = Σp{~((a+p)○b) X ~((a+)○c)}

250.II. ASSOCIATION FORMULÆ

1. AFFIRMATIVE i. Simple Formulæ

~(a~(bc)) = a(bc) = (ab)c = ~(~(ab)c)
~(a○~(bc)) = a(bc) = (ab)c = ~(~(ab)○c)
~(a○~(bc)) = a(bc) = (ab)c = ~(~(ab)c)
~(a~(bc)) = a(bc) = (ab)○c = ~(~(ab)c)
~(a~(bc)) = a(bc) = (ab)c = ~(~(ab)c)
~(a~(bc)) = a○(bc) = (ab)○c = ~(~(ab)c)
~(a~(bc)) = a(bc) = (ab)c = ~(~(ab)○c)
~(a~(bc)) = a○(bc) = (ab)c = ~(~(ab)c)

ii. Developments

(A and E are individual aggregants, and α and ε simple components of a and e. The summations and products are relative to all such aggregants and components. The formulæ are of four classes; and for any relative c either all formulæ of Class 1 or all of Class 2, and also either all of Class 3 or all of Class 4 hold good.)

CLASS 1.

~(a~(bc)) = a(bc) = Π{(Ab)c} = Π{~(~(Ab)c)}
~(a~(bc)) = a○(bc) = Σ{(ab)c} = Σ{~(~(ab)c)}
~(a○~(bc)) = a(bc) = Π{(ab)c} = Π{~(~(ab)c)}
~(a~(bc)) = a(bc) = Σ{(Ab)c} = Σ{~(~(Ab)c)}

CLASS 2.

~(~(cd)e) = (cd)e = Π{c○(dE)} = Π{~(c~(dE))}
~(~(cd)e) = (cd)○e = Σ{c(de)} = Σ{~(c○~(dε))}
~(~(cd)○e) = (cd)e = Π{c○(d○ε)} = Π{~(c~(d○ε))}
~(~(cd)e) = (cd)e = Σ{c(dE)} = Σ{~(c○~(dE))}

CLASS 3.

~(a~(bc)) = a○(bc) = Σ{((ab)c} = Σ{~(~(αb)c)}
~(a~(bc)) = a(bc) = Π{(Ab)○c} = Π{~(~(Ab)c)}
~(a~(bc)) = a(bc) = Σ{(Ab)c} = Σ{~(~(Ab)○c)}
~(a○~(bc)) = a(bc) = Π{(ab)○c} = Π{~(~(ab)c)}

CLASS 4.

~(~(cd)e) = (cd)○e = Σ{c(d○ε)} = Σ{~(c~(d○ε))}
~(~(cd)e) = (cd)e = Π{c(dE)} = Π{~(c~(dE))}
~(~(cd)e) = (cd)e = Σ{c(dE)} = Σ{~(c~(dE))}
~(~(cd)○e) = (cd)e = Π{c(dε)} = Π{~(c~(dε))}

The negative formulæ are derived from the affirmative by simply drawing or erasing lines over the whole of each member of every equation.

251. In order to see the general rules which these formulæ follow, we must imagine the operations symbolized by three marks, as in the commencement of this part [237]. We may term the operation uniting the two letters within the parenthesis the interior operation, and that which unites the whole parenthesis to the third letter the exterior operation. By junction-marks will be meant, in case the parenthesis inline image the third letter the inline image mark of the symbol of the interior operation and the inline image mark of the symbol of the exterior operation. Using these terms, we may say that the exterior junction-mark and the third mark of the interior operation may always be changed together. When they are the same there is a simple association formula. This formula consists in the possibility of simultaneously interchanging the junction-marks, the third marks, and the exteriority or interiority of the two operations. When the exterior junction-mark and the third mark of the interior operation are unlike, there is a developmental association formula. The general term of this formula is obtained by making the same interchanges as in the simple formulæ, and then changing a to A when after these interchanges ab or ab occurs in parenthesis, changing a to α when ab or ab occurs in parenthesis, changing e to E when de or de occurs in parenthesis, and changing e to ε when de or de occurs in parenthesis. When the third mark in the symbol of the exterior operation is affirmative the development is a summation; when this mark is negative there is a continued product.

In the first half of the formulæ, the second mark in the sign of the interior operation is a line in Class 1 and a hyphen in Class 3. In the second half, the first mark in the sign of the interior operation is a hyphen in Class 2 and a line in Class 4.



Paper 7: On the Logic of Number †1

§1. Definition of QuantityE

252. Nobody can doubt the elementary propositions concerning number: those that are not at first sight manifestly true are rendered so by the usual demonstrations. But although we see they are true, we do not so easily see precisely why they are true; so that a renowned English logician †2 has entertained a doubt as to whether they were true in all parts of the universe. The object of this paper is to show that they are strictly syllogistic consequences from a few primary propositions. The question of the logical origin of these latter, which I here regard as definitions, would require a separate discussion. In my proofs I am obliged to make use of the logic of relatives, in which the forms of inference are not, in a narrow sense, reducible to ordinary syllogism. They are, however, of that same nature, being merely syllogisms in which the objects spoken of are pairs or triplets. Their validity depends upon no conditions other than those of the validity of simple syllogism, unless it be that they suppose the existence of singulars, while syllogism does not.

The selection of propositions which I have proved will, I trust, be sufficient to show that all others might be proved with like methods.

253. Let r be any relative term, so that one thing may be said to be r of another, and the latter r'd by the former. If in a certain system of objects, whatever is r of an r of anything is itself r of that thing, then r is said to be a transitive relative in that system. (Such relatives as "lover of everything loved by —" are transitive relatives.) In a system in which r is transitive, let the q's of anything include that thing itself, and also every r of it which is not r'd by it. Then q may be called a fundamental relative of quantity; its properties being, first, that it is transitive; second, that everything in the system is q of itself, and, third, that nothing is both q of and q'd by anything except itself. The objects of a system having a fundamental relation of quantity are called quantities, and the system is called a system of quantity.

254. A system in which quantities may be q's of or q'd by the same quantity without being either q's of or q'd by each other is called multiple; †P1 a system in which of every two quantities one is a q of the other is termed simple.

§2. Simple Quantity

255. In a simple system every quantity is either "as great as" or "as small as" every other; whatever is as great as something as great as a third is itself as great as that third, and no quantity is at once as great as and as small as anything except itself.

256. A system of simple quantity is either continuous, discrete, or mixed. A continuous system is one in which every quantity greater than another is also greater than some intermediate quantity greater than that other. †1 A discrete system is one in which every quantity greater than another is next greater than some quantity (that is, greater than without being greater than something greater than). A mixed system is one in which some quantities greater than others are next greater than some quantities, while some are continuously greater than some quantities.

§3. Discrete Quantity

257. A simple system of discrete quantity is either limited, semi-limited, or unlimited. A limited system is one which has an absolute maximum and an absolute minimum quantity; a semi-limited system has one (generally considered a minimum) without the other; an unlimited has neither.

258. A simple, discrete, system, unlimited in the direction of increase or decrement, is in that direction either infinite or super-infinite. An infinite system is one in which any quantity greater than x can be reached from x by successive steps to the next greater (or less) quantity than the one already arrived at. In other words, an infinite, discrete, simple, system is one in which, if the quantity next greater than an attained quantity is itself attained, then any quantity greater than an attained quantity is attained; and by the class of attained quantities is meant any class whatever which satisfies these conditions. So that we may say that an infinite class is one in which if it is true that every quantity next greater than a quantity of a given class itself belongs to that class, then it is true that every quantity greater than a quantity of that class belongs to that class. Let the class of numbers in question be the numbers of which a certain proposition holds true. Then, an infinite system may be defined as one in which from the fact that a certain proposition, if true of any number, is true of the next greater, it may be inferred that that proposition if true of any number is true of every greater. †1

259. Of a super-infinite system this proposition, in its numerous forms, is untrue.

§4. Semi-Infinite Quantity

260. We now proceed to study the fundamental propositions of semi-infinite, †2 discrete, and simple quantity, which is ordinary number. †3

Definitions

261. The minimum number is called one.

262. By x+y is meant, in case x = 1, the number next greater than y; and in other cases, the number next greater than x'+y, where x' is the number next smaller than x.

263. By x×y is meant, in case x = 1, the number y, and in other cases y+x'y, where x' is the number next smaller than x.

264. It may be remarked that the symbols + and × are triple relatives, their two correlates being placed one before and the other after the symbols themselves.

Theorems

265. The proof in each case will consist in showing, first, that the proposition is true of the number one, and second, that if true of the number n it is true of the number 1+n, next larger than n. The different transformations of each expression will be ranged under one another in one column, with the indications of the principles of transformation in another column.

266. (1) To prove the associative principle of addition, that

(x+y)+z = x+(y+z)

whatever numbers x, y, and z, may be. First it is true for x = 1; for

(1+y)+z

= 1+(y+z) by the definition of addition, second clause. Second, if true for x = n, it is true for x = 1+n; that is, if (n+y)+z = n+(y+z) then ((1+n)+y)+z = (1+n)+(y+z).

For

((1+n)+y)+z
= (1+(n+y))+z by the definition of addition:
= 1 +((n+y)+z) by the definition of addition:
= 1 +(n+(y+z)) by hypothesis:
= (1+n)+(y+z) by the definition of addition,

267. (2) To prove the commutative principle of addition that

x+y = y+x

whatever numbers x and y may be. First, it is true for x = 1 and y = 1, being in that case an explicit identity. Second if true for x = n and y = 1, it is true for x = 1+n and y = 1. That is, if n+1 = 1+n, then (1+n)+1 = 1+(1+n). For (1+n)+1

= 1+(n+1) by the associative principle:
= 1+(1+n) by hypothesis.

We have thus proved that whatever number x may be x+1 = 1+x, or that x+y = y+x for y = 1. It is now to be shown that if this be true for y = n, it is true for y = 1+n; that is, that if x+n = n+x, then x+(1+n) = (1+n)+x. Now

x+(1+n)
= (x+1)+n by the associative principle:
= (1+x)+n as just seen:
= 1+(x+n) by the definition of addition:
= 1+(n+x) by hypothesis:
= (1+n)+x by the definition of addition.

Thus the proof is complete.

268. (3) To prove the distributive principle, first clause. The distributive principle consists of two propositions:

First, (x+y)z = xz+yz
Second, x(y+z) = xy+xz.

We now undertake to prove the first of these. First it is true for x = 1. For

(1+y)z
= z+yz by the definition of multiplication:
= 1.z+yz by the definition of multiplication.

Second, if true for x = n it is true for x = 1+n; that is, if (n+y)z = n z+yz then ((1+n)+y)z = (1+n)z+yz. For

((1+n)+y)z
= (1+(n+y))z by the definition of addition:
= z+(n+y)z by the definition of multiplication:
= z+(n z+yz) by hypothesis:
= (z+n z)+yz by the associative principle of addition:
= (1+n)z+yz by the definition of multiplication.

269. (4) To prove the second proposition of the distributive principle, that

x(y+z) = xy+xz.

First, it is true for x = 1; for

1(y+z)
= y+z by the definition of multiplication;
= 1y+1z by the definition of multiplication.

Second, if true for x = n, it is true for x = 1+n; that is, if n(y+z) = n y+n z, then (1+n)(y+z) = (1+n)y+(1+n)z. For

(1+n)(y+z)
= (y+z)+n(y+z) by the definition of multiplication:
= (y+z)+(n y+n z) by hypothesis:
= (y+n y)+(z+n z) by the principles of addition:
= (1+n)y+(1+n)z by the definition of multiplication.

270. (5) To prove the associative principle of multiplication; that is, that

(xy)z = x(yz)

whatever numbers x, y, and z, may be. First, it is true for x = 1, for

(1y)z
= yz by the definition of multiplication:
= 1.yz by the definition of multiplication.

Second, if true for x = n, it is true for x = 1+n; that is, if (n y)z = n(yz) then ((1+n)y)z = (1+n)(yz). For

((1+n)y)z
= (y+n y)z by the definition of multiplication:
= yz+(n y)z by the distributive principle:
= yz+n(yz) by hypothesis:
= (1+n)(yz) by the definition of multiplication.

271. (6) To prove the commutative principle of multiplication; that

xy = yx,

whatever numbers x and y may be. In the first place we prove that it is true for y = 1. For this purpose, we first show that it is true for y = 1 x = 1; and then that if true for y = 1, x = n it is true for y = 1, x = 1+n. For y = 1 and x = 1, it is an explicit identity. We have now to show that if n1 = 1n then (1+n)1 = 1(1+n). Now

(1+n)1
= 1+n1 by the definition of multiplication:
= 1+1n by hypothesis:
= 1+n by the definition of multiplication:
= 1(1+n) by the definition of multiplication.

Having thus shown the commutative principle to be true for y = 1, we proceed to prove that if it is true for y = n, it is true for y = 1+n; that is, if x n = n x, then x(1+n) = (1+n)x. For

(1+n)x
= x+n x by the definition of multiplication:
= x+x n by hypothesis:
= 1x+x n by the definition of multiplication:
= x1+x n as already seen:
= x(1+n) by the distributive principle.

§5. Discrete Simple Quantity Infinite in Both Directions

272. A system of number infinite in both directions has no minimum, but a certain quantity is called one and the numbers as great as this constitute a partial system of semi-infinite number, of which this one is a minimum. The definitions of addition and multiplication require no change except that the one therein is to be understood in the new sense.

273. To extend the proofs of the principles of addition and multiplication to unlimited number, it is necessary to show that if true for any number (1+n) they are also true for the next smaller number n. For this purpose we can use the same transformations as in the second clauses of the former proof; only we shall have to make use of the following lemma.

274. If x+y = x+z then y = z whatever numbers x, y, and z, may be. First this is true in case x = 1 for then y and z are both next smaller than the same number. Therefore, neither is smaller than the other, otherwise it would not be next smaller to 1+y = 1+z. But in a simple system, of any two different numbers one is smaller. Hence, y and z are equal. Second, if the proposition is true for x = n, it is true for x = 1+n. For if (1+n)+y = (1+n)+z, then by the definition of addition 1+(n+y) = 1+(n+z); whence it would follow that n+y = n+z, and, by hypothesis, that y = z. Third, if the proposition is true for x = 1+n it is true for x = n. For if n+y = n+z, then 1+n+y = 1+n+z because the system is simple. The proposition has thus been proved to be true of 1 of every greater and of every smaller number and therefore to be universally true.

275. An inspection of the above proofs of the principles of addition and multiplication for semi-infinite number will show that they are readily extended to doubly infinite number by means of the proposition just proved.

276. The number next smaller than one is called naught, 0. This definition in symbolic form is 1+0 = 1. To prove that x+0 = x, let x' be the number next smaller than x. Then

x+0
= (1+x')+0 by the definition of x':
= (1+0)+x' by the principles of addition:
= 1+x' by the definition of naught:
= x by the definition of x'.

277. To prove that x0 = 0. First, in case x = 1 the proposition holds by the definition of multiplication. Next, if true for x = n it is true for x = 1+n. For

(1+n)0
= 1.0+n.0 by the distributive principle:
= 1.0+0 by hypothesis:
= 1.0 by the last theorem:
= as above.

Third, the proposition, if true for x = 1+n, is true for x = n. For, changing the order of the transformations,

1.0+0 = 1.0 = 0 = (1+n)0 = 1.0+n.0.

Then by the above lemma, n.0 = 0 so that the proposition is proved.

278. A number which added to another gives naught is called the negative of the latter. To prove that every number greater than naught has a negative. First, the number next smaller than naught is the negative of one; for, by the definition of addition, one plus this number is naught. Second, if any number n has a negative then the number next greater than n has for its negative the number next smaller than the negative of n. For let m be the number next smaller than the negative of n. Then n+(1+m) = 0.

But

n+(1+m)
= (n+1)+m by the associative principle of addition.
= (1+n)+m by the commutative principle of addition.

So that (1+n)+m = 0. Q.E.D. Hence every number greater than 0 has a negative, and naught is its own negative.

To prove that (-x)y = -(xy). We have

0 = x+(-x) by the definition of the negative:
0 = 0y = (x+(-x))y by the last proposition but one:
0 = xy+(-x)y by the distributive principle:
-(xy) = (-x)y by the definition of the negative.

279. The negative of the negative of a number is that number. For x+(-x) = 0. Whence by the definition of the negative x = -(-x).

§6. Limited Discrete Simple Quantity

280. Let such a relative term, c, that whatever is a c of anything is the only c of that thing, and is a c of that thing only, be called a relative of simple correspondence. †1 In the notation of the logic of relatives

c ⤙ 1, c̆c ⤙ 1. †2

281. If every object, s, of a class is in any such relation being c'd by †3 a number of a semi-infinite discrete simple system, and if further every number smaller than a number c of †4 an s is itself c of an s, then the numbers c of the s's are said to count them, †5 and the system of correspondence is called a count. In logical notation, putting g for as 'great as,' and n for a positive integral number,

sc̆n ğcscs

If in any count there is a maximum counting number the count is said to be finite, and that number is called the number of the count. Let [s] denote the number of a count of the s's, then

[s] ⤙ cs ḡcs ⤙ ~[s] .

282. The relative "identical with" satisfies the definition of a relative of simple correspondence, and the definition of a count is satisfied by putting "identical with" for c, and "positive integral number as small as x" for s. In this mode of counting, the number of numbers as small as x is x.

283. Suppose that in any count the number of numbers as small as the minimum number, one, is found to be n. Then, by the definition of a count, every number as small as n counts a number as small as one. But by the definition of one there is only one number as small as one. Hence, by the definition of single correspondence, no other number than one counts one. Hence, by the definition of one, no other number than one counts any number as small as one. Hence, by the definition of the count, one is, in every count, the number of numbers as small as one.

284. If the number of numbers as small as x is in some count y, then the number of numbers as small as y is in some count x. For if the definition of a simple correspondence is satisfied by the relative c, it is equally satisfied by the relative c'd by.

285. Since the number of numbers as small as x is in some count y, we have, c being some relative of simple correspondence,

First. Every number as small as x is c'd by a number.

Second. Every number as small as a number that is c of a number as small as x is itself c of a number as small as x.

Third. The number y is c of a number as small as x.

Fourth. Whatever is not as great as a number that is c of a number as small as x is not y.

286. Now let c1 be the converse of c. Then the converse of c1 is c; whence, since c satisfies the definition of a relative of simple correspondence, so also does c1. By the third proposition above, every number as small as y is as small as a number that is c of a number as small as x. Whence, by the second proposition every number as small as y is c of a number as small as x; and it follows that every number as small as y is c1'd by a number. It follows further that every number c1 of a number as small as y is c1 of something c1'd by (that is, c1 being a relative of simple correspondence, is identical with) some number as small as x. Also, "as small as" being a transitive relative, every number as small as a number c of a number as small as y is as small as x. Now by the fourth proposition y is as great as any number that is c of a number as small as x, so that what is not as small as y is not c of a number as small as x; whence whatever number is c'd by a number not as small as y is not a number as small as x. But by the second proposition, every number as small as x not c'd by a number not as small as y is c'd by a number as small as y. Hence, every number as small as x is c'd by a number as small as y. Hence, every number as small as a number c1 of a number as small as y is c1 of a number as small as y. Moreover, since we have shown that every number as small as x is c1 of a number as small as y, the same is true of x itself. Moreover, since we have seen that whatever is c1 of a number as small as y is as small as x, it follows that whatever is not as great as a number c1 of a number as small as y is not as great as a number as small as x; i.e. ("as great as" being a transitive relative) is not as great as x, and consequently is not x. †P1 We have now shown:

First, that every number as small as y is c1'd by a number;

Second, that every number as small as a number that is c1 of a number as small as y is itself c1 of a number as small as y;

Third, that the number x is c1 of a number as small as y; and

Fourth, that whatever is not as great as a number that is c1 of a number as small as y is not x.

These four propositions taken together satisfy the definition of the number of numbers as small as y counting up to x.

Hence, since the number of numbers as small as one cannot in any count be greater than one, it follows that the number of numbers as small as any number greater than one cannot in any count be one.

287. Suppose that there is a count in which the number of numbers as small as 1+m is found to be 1+n, since we have just seen that it cannot be 1. In this count, let m' be the number which is c of 1+n, and let n' be the number which is c'd by 1+m. Let us now consider a relative, e, which differs from c only in excluding the relation of m' to 1+n as well as the relation of 1+m to n' and in including the relation of m' to n'. Then e will be a relative of single correspondence; for c is so, and no exclusion of relations from a single correspondence affects this character, while the inclusion of the relation of m' to n' leaves m' the only e of n' and an e of n' only. Moreover, every number as small as m is e of a number, since every number except 1+m that is c of anything is e of something, and every number except 1+m that is as small as 1+m is as small as m. Also, every number as small as a number e'd by a number is itself e'd by a number; for every number c'd is e'd except 1+m, and this is greater than any number e'd. It follows that e is the basis of a mode of counting by which the numbers as small as m count up to n. Thus we have shown that if in any way 1+m counts up to 1+n, then in some way m counts up to n. But we have already seen that for x = 1 the number of numbers as small as x can in no way count up to other than x. Whence it follows that the same is true whatever the value of x. †P1

288. If every S is a P, and if the P's are a finite lot counting up to a number as small as the number of S's, then every P is an S. For if, in counting the P's, we begin with the S's (which are a part of them), and having counted all the S's arrive at the number n, there will remain over no P's not S's. For if there were any, the number of P's would count up to more than n. From this we deduce the validity of the following mode of inference:

Every Texan kills a Texan,

Nobody is killed by but one person,

Hence, every Texan is killed by a Texan,

supposing Texans to be a finite lot. For, by the first premiss, every Texan killed by a Texan is a Texan killer of a Texan, By the second premiss, the Texans killed by Texans are as many as the Texan killers of Texans. Whence we conclude that every Texan killer of a Texan is a Texan killed by a Texan, or, by the first premiss, every Texan is killed by a Texan. This mode of reasoning †1 is frequent in the theory of numbers.



Paper 8: Associative Algebras †1E

§1. On the Relative Forms of the Algebras

289. Given an associative algebra whose letters are i, j, k, l, etc., and whose multiplication table is

i2 = a11i + b11j + c11k + etc. †P1

ij = a12i + b12j + c12k + etc.

ji = a21i + b21j + c21k + etc.,

etc., etc.

I proceed to explain what I call the relative form of this algebra.

290. Let us assume a number of new units, A, I, J, K, L, etc., one more in number than the letters of the algebra, and every one except the first, A, corresponding to a particular letter of the algebra. These new units are susceptible of being multiplied by numerical coefficients and of being added together; †P2 but they cannot be multiplied together, and hence are called non-relative units.

291. Next, let us assume a number of operations each denoted by bracketing together two non-relative units separated by a colon. These operations, equal in number to the square of the number of non-relative units, may be arranged as follows:

(A:A) (A:I) (A:J) (A:K), etc.
(I:A) (I:I) (I:J) (I:K), etc.
(J:A) (J:I) (J:J) (J:K), etc.

292. Any one of these operations performed upon a polynomial in non-relative units, of which one term is a numerical multiple of the letter following the colon, gives the same multiple of the letter preceding the colon. Thus, (I:J)(a I+b J+c K) = b I. †P1 These operations are also taken to be susceptible of associative combination. Hence (I:J)(J:K) = (I:K); for (J:K)K = J and (I:J)J = I, so that (I:J)(J:K)K = I. And (I:J)(K:L) = 0; for (K:L)L = K and (I:J)K = (I:J)(0.J+K) = 0.I = 0. We further assume the application of the distributive principle to these operations; so that, for example, {(I:J)+(K:J)+(K:L)} (a J+b L) = a J+(a+b)K.

293. Finally, let us assume a number of complex operations denoted by i', j', k', l', etc., corresponding to the letters of the algebra and determined by its multiplication table in the following manner:

i' = (I:A) +a11(I:I)+b11(J:I)+c11(K:I)+etc.
+a12(I:J)+b12(J:J)+c12(K:J)+etc.
+a13(I:K)+b13(J:K)+c13(K:K)+etc.+etc.
j' = (J:A) +a21(I:I)+b21(J:I)+c21(K:I)+etc.
+a22(I:J)+b22(J:J)+c22(K:J)+etc.
+a23(I:K)+b23(J:K)+c23(K:K)+etc.+etc.
k' = etc.

294. Any two operations are equal which, being performed on the same operand, invariably give the same result. The ultimate operands in this case are the non-relative units. But any operations compounded by addition or multiplication of the operations i', j', k', etc., if they give the same result when performed upon A, will give the same result when performed upon any one of the non-relative units. For suppose i'j'A = k'l'A. We have

i'j'A = i'J = a12I+b12J+c12K+etc.

k'l'A = k'L = a34I+b34J+c34K+etc.

so that a12 = a34, b12 = b34, c12 = c34, etc., and in our original algebra ij = kl. Hence, multiplying both sides of the equation into any letter, say m, i j m = k l m. But

ijm = i(a25i+b25j+c25k+etc.)

= (a11a25+a12b25+a13c25+etc.)i + (b11a25 + b12b25 + b13c25 + etc.)j + etc.

But we have equally

i'j'm'A = (a11a25+a12b25+a13c25+etc.)I + (b11a25+b12b25+b13c25+etc.)J+etc.

So that i'j'm'A = k'l'm'A. Hence, i'j'M = k'l'M. It follows, then, that if i'j'A = k'l'A, then i'j' into any non-relative unit equals k'l' into the same unit, so that i'j' = k'l'. We thus see that whatever equality subsists between compounds of the accented letters i', j', k', etc., subsists between the same compounds of the corresponding unaccented letters i, j, k, so that the multiplication tables of the two algebras are the same. †P1 Thus, what has been proved is that any associative algebra can be put into relative form, i. e. (see my brochure entitled A Brief Description of the Algebra of Relatives) †1 that every such algebra may be represented by a matrix.

Take, for example, the algebra (b d5). †2 It takes the relative form

i = (I:A)+(J:I)+(L:K), j = (J:A),

k = (K:A)+(J:I)+𝖗(L:I)+(I:K)+(M:K)+𝖗(J:L)-(J:M)-𝖗(L:M),

l = (L:A)+(J:K), m = (M:A)+(𝖗2 - 1)(J:I)-(L:K)-𝖗2(J:M).

This is the same as to say that the general expression xi+yj+zk+u l+v m of this algebra has the same laws of multiplication as the matrix

0, 0, 0, 0, 0, 0,
x, 0, 0, z, 0, 0,
y, +x+z(𝖗2-1)v 0, u, 𝖗z, -z-𝖗2v,
z, 0, 0, 0, 0, 0,
u, 𝖗z, 0, x-v, 0, -𝖗z,
v, 0, 0, z, 0, 0,

295. Of course, every algebra may be put into relative form in an infinity of ways; and simpler ways than that which the rule affords can often be found. Thus, for the above algebra, the form given in the foot-note is simpler, and so is the following:

i = (B:A)+(C:B)+(F:D)+(C:E), j = (C:A),

k = (D:A)+(E:D)+(C:B)+𝖗(F:B)+𝖗(C:F),

l = (F:A)+(C:D), m = (E:A)+(𝖗2-1)(C:B)-(B:A)-(F:D)-(C:E).

These different forms will suggest transformations of the algebra. Thus, the relative form in the foot-note to (b d5) suggests putting

i1 = i+m, j1=𝖗2j, k1 = k+𝖗-1i+𝖗-1m, l1 = 𝖗l+j, m1 = -m,

when we get the following multiplication table, where {r} is put for 𝖗-1:

i j k l m
i 0 0 0 0 j
j 0 0 0 0 0
k 0 0 i j l
l 0 0 {r}j 0 0
m {r}2j 0 {r}l 0 j

296. Ordinary algebra with imaginaries, considered as a double algebra, is, in relative form,

1 = (X:X)+(Y:Y), inline image = (X:Y)-(Y:X).

This shows how the operation inline image turns a vector through a right angle in the plane of X, Y. Quaternions in relative form is

1 = (W:W)+(X:X)+(Y:Y)+(Z:Z),

i = (X:W)-(W:X)+(Z:Y)-(Y:Z),

j = (Y:W)-(Z:X)-(W:Y)+(X:Z),

k = (Z:W)+(Y:X)-(X:Y)-(W:Z).

We see that we have here a reference to a space of four dimensions corresponding to X, Y, Z, W.

§2. On the Algebras In Which Division is Unambiguous

297. (1) In the Linear Associative Algebra, the coefficients are permitted to be imaginary. In this note they are restricted to being real. It is assumed that we have to deal with an algebra such that from A B = A C we can infer that A = 0 or B = C. It is required to find what forms such an algebra may take.

298. (2) If A B = 0, then either A = 0 or B = 0. For if not, A C = A(B+C), although A does not vanish and C is unequal to B+C.

299. (3) The reasoning of §40 [of the Linear Associative Algebra] holds, although the coefficients are restricted to being real. It is true, then, that since there is no expression (in the algebra under consideration) whose square vanishes, there must be an expression, i, such that i2 = i.

300. (4) By §41, it appears that for every expression in the algebra we have

i A = A i = A.

301. (5) By the reasoning of §53, it appears that for every expression A there is an equation of the form

Σm(amAm)+b i = 0.

But i is virtually arithmetical unity, since i A = A i = A; and this equation may be treated by the ordinary theory of equations. Suppose it has a real root, α; then it will be divisible by (A-α), and calling the quotient B we shall have

(Ai)B = 0.

But Ai is not zero, for A was supposed dissimilar to i. Hence a product of finites vanishes, which is impossible. Hence the equation cannot have a real root. But the whole equation can be resolved into quadratic factors, and some one of these must vanish. Let the irresoluble vanishing factor be

(A-s)2+t2 = 0.

Then

((A-s)/t)2 = -1,

or, every expression, upon subtraction of a real number (i.e. a real multiple of i), can be converted, in one way only, into a quantity whose square is a negative number. We may express this by saying that every quantity consists of a scalar and a vector part. A quantity whose square is a negative number we here call a vector.

302. (6) Our next step is to show that the vector part of the product of two vectors is linearly independent of these vectors and of unity. That is, i and j being any two vectors, †P1 if

ij = s+v

where s is a scalar and v a vector, we cannot determine three real scalars a, b, c, such that

v = a+b i+c j.

This is proved, if we prove that no scalar subtracted from ij leaves a remainder b i+c j. If this be true when i and j are any unit vectors whatever, it is true when these are multiplied by real scalars, and so is true of every pair of vectors. We will, then, suppose i and j to be unit vectors. Now,

ij2 = -i.

If therefore we had

ij = a+b i+c j,

we should have

-i = ij2 = aj+b i j-c = ab-c+b2i+(a+bc)j;

whence, i and j being dissimilar,

-i = b2i, b2 = -1,

and b could not be real.

303. (7) Our next step is to show that, i and j being any two vectors, and

ij = s+v,

s being a scalar and v a vector, we have

ji = r(s-v),

where r is a real scalar. It will be obviously sufficient to prove this for the case in which i and j are unit vectors. Assuming them such, let us write

ji = s'+v', v v' = s'' +v'',

where s' and s'' are scalars, while v' and v'' are vectors. Then

ij.ji = (s+v)(s'+v') = s s'+s v'+s'v+v''+s''.

But we have

ij.ji = ij2i = -i2 = 1.

Hence,

v'' = 1-ss'-s''-sv'-s'v.

But v'' is the vector of v v', so that by the last paragraph such an equation cannot subsist unless v'' vanishes. Thus we get

0 = 1-ss'-s''-sv'-s'v,

or

sv' = 1-ss'-s''-s'v.

But a quantity can only be separated in one way into a scalar and a vector part; so that

s v' = -s'v.

That is,

ji = (s'/s)(s-v). Q. E. D.

304. (8) Our next step is to prove that s = s'; so that if ij = s+v then ji = s-v. It is obviously sufficient to prove this when i and j are unit vectors. Now from any quantity a scalar may be subtracted so as to leave a remainder whose square is a scalar. We do not yet know whether the sum of two vectors is a vector or not (though we do know that it is not a scalar). Let us then take such a sum as ai+b j and suppose x to be the scalar which subtracted from it makes the square of the remainder a scalar. Then, C being a scalar,

(-x+ai+b j)2 = C.

But developing the square we have

(-x+ai+bj)2=x2-a2-b2+abs+abs'-2axi+2bxj+ab(1-(s'/s))v = C;

i.e.

ab(1-(s'/s))v=C-x2+a2+b2-abs-abs'+2axi+2bxj.

But v being the vector of ij, by the last paragraph but one the equation must vanish. Either then v = 0 or 1-(s'/s) = 0. But if v = 0, ij = s, and multiplying into j,

-i = s j,

which is absurd, i and j being dissimilar. Hence 1-(s'/s) = 0 and

ji = s-v. Q.E.D.

305. (9) The number of independent vectors in the algebra cannot be two. For the vector of ij is independent of i and j. There may be no vector, and in that case we have the ordinary algebra of reals; or there may be only one vector, and in that case we have the ordinary algebra of imaginaries.

Let i and j be two independent vectors such that

ij = s+v.

Let us substitute for j

j1 = s i+j.

Then we have

ij1 = v, j1i = -v,
j1v = j1ij1 = -j12i = i, vj1 = ij12 = -i,
i v = i2j1 = -j1, v i = ij1i = -j1i2 = j1

Thus we have the algebra of real quaternions. Suppose we have a fourth unit vector, k, linearly independent of all the others, and let us write

j1k = s'+v',

ki = s''+v''.

Let us substitute for k

k1 = s''i+s'j1+k,

and we get

j1k1 = -s''v+v', k1j1= s''v-v',
k1i = -s'v +v'', ik1 = s'v-v''.

Let us further suppose

(ij1)k1 = s''+v'''.

Then, because ij1 is a vector,

k1(ij1) = s'''-v'''.

But

k1j1 = -j1k1, k1i = -ik1,

because both products are vectors.

Hence

i.j1k1 = -i.k1j1 = -ik1.j1 = k1i.j1 = k1.ij1.

Hence

s'''+v''' = s'''-v'''

or v''' = 0, and the product of the two unit vectors is a scalar. These vectors cannot, then, be independent, or k cannot be independent of ij = v. Thus it is proved that a fourth independent vector is impossible, and that ordinary real algebra, ordinary algebra with imaginaries, and real quaternions are the only associative algebras in which division by finites always yields an unambiguous quotient.



Paper 9: Brief Description of the Algebra of Relatives †1

306. Let A, B, C, etc., denote objects of any kind. These letters may be conceived to be finite in number or innumerable. The sum of them, each affected by a numerical coefficient (which may equal 0), is called an absolute term. Let x be such a term; then we write

x = (x)aA + (x)bB+ (x)cC+etc. = Σi(x)iI.

Here (x), etc., are numbers, which may be permitted to be imaginary or restricted to being real or positive, or to being roots of any given equation, algebraic or transcendental. †P1 By φx, any mathematical function of the absolute term x, we mean such an absolute term that

x)i = φ(x)i.

That is, each numerical coefficient of φx is the function, φ, of the corresponding coefficient of x. In particular,

(x+y)i = (x)i+(y)i,

(x×y)i = (x)i×(y)i.

Otherwise written,

x+y = {(x)a+(y)a}A + {(x)b+(y)b}B + etc.

x×y = {(x)a×(y)a}A + {(x)b×(y)b}B + etc.

307. Two peculiar absolute terms are suggested by the logic of the subject. I call them terms of second intention. The first is zero, 0, and is defined by the equation

(0)i = 0

or

0 = 0.A+0.B+0.C+etc.

The other is ens (or non-relative unity), 0̅, and is defined by the equation

(0̅)i = 1,

or

0̅ = A+B+C+etc.

308. The symbol (A:B) is called an individual dual relative. It signifies simply a pair of individual objects, (A:B) and (B:A) being different. An aggregate of such symbols, each affected by a numerical coefficient, is called a general dual relative. The totality of pairs of letters arrange themselves with obvious naturalness in the block,

A:A A:B A:C etc.
B:A B:B B:C etc.
C:A C:B C:C etc.
etc. etc. etc. etc.

309. If l denotes any general dual relative, then the coefficient of the pair I:J in l is written (l)ij These coefficients are thus each referred to a place in the above block, and may themselves be arranged in the block

(l)aa (l)ab (l)ac etc.
(l)ba (l)b b (l)bc etc.
(l)ca (l)cb (l)cc etc.
etc. etc. etc. etc.

310. Every relative term, x, is separable into a part called 'self-x,' Sx, such that

Sx = Σi(x)ii(I:I)

and the remaining part, called 'alio-x,' Vx; comprising all the terms in x not in the principal diagonal of the block; so that we write

x = Sx+Vx. †1

311. Each absolute term is considered to be equivalent to a certain relative term; namely,

A = (A:A)+(A:B)+(A:C) + etc.

or, if x be an absolute term,

(x)ij = (x)i.

The self-part of the relative equivalent to an absolute term is denoted by writing a comma after the term. Accordingly,

(x,)ii = (x)i, (x,)ij = 0

312. Besides 0 and 0̅, two other dual relative terms have been called terms of second intention. These are simply S0̅ and V0̅. The relative S0̅ or (0̅,) is also written 1, and is called unity, or 'identical with.' It is defined by the equations

(1)ii = 1, (1)ij = 0.

That is,

1 = (A:A)+(B:B)+(C:C) + etc.

The relative V0̅ is written 1̅ or 𝖓, and is called 'not,' or 'the negative of.' It is defined by the equations

(1̅)ij = 0, †1 (1̅)ij = 1.

313. By an absolute function of a relative term is meant that function taken according to the rule for taking the function of an absolute term. That is,

x)ij = φ(x)ij.

In particular,

(x+y)ij = (x)ij+(y)ij

(x×y)ij = (x)ij×(y)ij

314. Of the various external or relative combinations that have been employed the following may be particularly specified. †2 (1), External multiplication, defined by the equation

(xy)ijn(x)in(y)nj

(2), External progressive involution, defined by the equation

(xy)ijn=(x)in(y)nj

(3), External regressive involution, defined by the equation

(xy)ijn(y)nj(x)in.

In general, using Miss Ladd's notation †P1 for the different orders of multiplication,

inline image

Other modes of external combination have been used, but they are believed to have only a special utility. Division does not generally yield an unambiguous quotient. Indeed, I have shown that it does so only in the cases of ordinary real algebra, of imaginary algebra, and of real quaternions. †1

315. Besides the mathematical functions of relatives, there are various modes in which one relative may logically depend upon another. Thus, S x and V x may be said to be logical functions of x. The most important of such operations is that of taking the converse of a relative. The converse of x, written or Kx, is defined by the equation

()ij = (x)ji.

316. The algebraical laws of all these combinations are obtained with great facility by a method of which the following are examples:

Example 1.
{(xy)z}ij n(xy)in(z)njnΣm(x)im(y)mn(z)nj
{x(yz)}ij m(x)im(yz)mjmΣn(x)im(y)mn(z)nj
∴ (xy)z =x(yz).
Example 2.
{(x+y)z}ij n(x+y)in(z)njn{(x)in+(y)in}(z)nj
n(x)in(z)njn(y)in(z)nj=(xz)ij+(yz)ij
∴ (x+y)z =xz+yz.

The following are some of the elementary formulæ so obtained. Non-relative multiplication is indicated by a comma, relative multiplication by writing the factors one after the other, without the intervention of any sign.

(x+y)+z=x+(y+z), x+y=y+x,
(x,y),z=x,(y,z), x,y=y,x,
(x+y),z=(x,z)+(y,z),
(xy)z=x(yz),
(x+y)z=xz+yz, x(y+z)=xy+xz,
(xy)z=x(yz), x(yz)=(xy)z, x(yz)=(xy)z,
(x,y)z=(xz),(yz), x(y,z)=(xy),(xz),
xy+z=(xy),(xz), x+yz=(xz),(yz)
kkx=x
k(x+y)=kx+ky, k(x,y)=kx,ky
k(xy)=(ky)(kx) k(xy)=(ky)(kx)
0+x=0, 0,x=0x=x0=0, x0=0x=0̅,
0̅×x=x,0̅x=x0̅=0̅,
1x=x1=x1=1x=x,
(x1̅)ij = (1̅x)ij=
0, if xij
1, if xij=0

317. Just as the different pairs of letters, A, B, C, etc., have been conceived to be arranged in a square block, so the different triplets of them may be conceived to be arranged in a cube, and the algebraical sum of all such triplets, each affected with a numerical coefficient, may be called a triple relative.

Every dual relative may be regarded as equivalent to a triple relative, just as every absolute term is equivalent to a dual relative.

Every triple relative may be regarded as a sum of five parts, each being a linear expression in terms of one of the five forms,

(A:A):A (A:B):A (A:A):B (B:A):A (A:B):C

The sign of a dual relative followed by a comma denotes that part of the equivalent triple relative which consists of terms in one of the forms

(A:A):(A:A) (A:B):(A:B).

The multiplication of triple relatives is not perfectly associative and the multiplication of two triple relatives yields a quadruple relative.

The modes of combination of a triple relative followed by two dual relatives are the same as the modes of combination of three dual relatives. This ceases to be true for quadruple and higher relatives.

Corresponding to the operation of taking the converse of a dual relative, there are five operations upon triple relatives. They are defined as follows:

(Ix)ijk=(x)jik,(Jx)ijk=(x)ikj,(Kx)ijk=(x)kji,(Lx)ijk=(x)jki,(Mx)ijk=(x)kij.

Every quadruple or higher relative may be conceived as a product of triple relatives.

318. Thus, the essential characteristics of this algebra are (1) that it is a multiple algebra depending upon the addition of square blocks or cubes of numbers, (2) that in the external multiplication the rows of the block of the first factor are respectively multiplied by the columns of the block of the second factor, and (3) that the multiplication so resulting is, for the two-dimensional form of the algebra, always associative. I have proved in a paper presented to the American Academy of Arts and Sciences, May 11, 1875, †1 that this algebra necessarily embraces every associative algebra.

319. I have here described the algebra apart from the logical interpretation with which it has been clothed. In this interpretation a letter is regarded as a name applicable to one or more objects. By a name is usually meant something representative of an object to a mind. But I generalize this conception and regard a name as merely something in a conjoint relation to a second and a third, that is as a triple relative. †2 A sum of different individual names is a name for each of the things named severally by the aggregant letters. A name multiplied by a positive integral coefficient is the aggregate of so many different senses in which that name may be taken. The individual relative A:B is the name of A considered as the first member of the pair A:B. The signification of the external multiplication is then determined by its algebraical definition.

320. Professor Sylvester, in his "New Universal Multiple Algebra," †3 appears to have come, by a line of approach totally different from mine, upon a system which coincides, in some of its main features, with the Algebra of Relatives, as described in my four papers upon the subject, †P1 and in my lectures on logic. I am unable to judge, from my unprofessional acquaintance with pure mathematics, how much of novelty there may be in my conceptions; but as the researches of the illustrious geometer who has now taken up the subject must draw increased attention to this kind of algebra, I take occasion to redescribe the outlines of my own system, and at the same time to declare my modest conviction that the logical interpretation of it, far from being in any degree special, will be found a powerful instrument for the discovery and demonstration of new algebraical theorems.

321. Postscript. — I have this day had the delight of reading for the first time Professor Cayley's Memoir on Matrices, in the Philosophical Transactions for 1858. The algebra he there describes seems to me substantially identical with my long subsequent algebra for dual relatives. Many of his results are limited to the very exceptional cases in which division is a determinative process.

322. My own studies in the subject have been logical not mathematical, being directed toward the essential elements of the algebra, not towards the solution of problems.



Paper 10: On the Relative Forms of Quaternions †1

323. If X, Y, Z denote the three rectangular components of a vector, and W denote numerical unity (or a fourth rectangular component, involving space of four dimensions), and (Y:Z) denote the operation of converting the Y component of a vector into its Z component, then

1 = (W:W)+(X:X)+(Y:Y)+(Z:Z)

i = (X:W)-(W:X)-(Y:Z)+(Z:Y)

j = (Y:W)-(W:Y)-(Z:X)+(X:Z)

k = (Z:W)-(W:Z)-(X:Y)+(Y:X).

In the language of logic (Y:Z) is a relative term whose relate is a Y component, and whose correlate is a Z component. The law of multiplication is plainly (Y:Z)(Z:X) = (Y:X), (Y:Z)(X:W) = 0, and the application of these rules to the above values of 1, i, j, k gives the quaternion relations

i2 = j2 = k2 = -1, ijk = -1, etc.

The symbol a(Y:Z) denotes the changing of Y to Z and the multiplication of the result by a. If the relatives be arranged in the block

W:W W:X W:Y W:Z
X:W X:X X:Y X:Z
Y:W Y:X Y:Y Y:Z
Z:W Z:X Z:Y Z:Z,

then the quaternion w+xi+yj+zk is represented by the matrix of numbers

w -x -y -z
x w -z y
y z w -x
z -y x w.

The multiplication of such matrices follows the same laws as the multiplication of quaternions. The determinant of the matrix = the fourth power of the tensor of the quaternion.

The imaginary x+y√-1 may likewise be represented by the matrix

x y
-y x,

and the determinant of the matrix = the square of the modulus.



Paper 11: On a Class of Multiple Algebras †1

324. The object of this paper is to show what algebras express all the substitutions of two, of three, and of four letters; and to put these algebras into familiar forms. †2

It is evident that every substitution is a relative term. Thus, the transposition of A B to B A is in relative form (A:B)+(B:A), and the circular substitution inline image is (B:A)+(C:B)+(A:C). In this point of view, we see that substitutions may be added and multiplied by scalars, although the results will usually no longer be substitutions. A group of substitutions may, then, be linear expressions in an associative multiple algebra of a lower order than that of the group.

325. Of two letters, there are two substitutions (X:X)+(Y:Y) and (X:Y)+(Y:X). We may denote these by a and β respectively, so that taking A and B as indeterminate coefficients, the general expression of the algebra is A a+Bβ, or in the form of a matrix is

A B
B A.
Assume i and j such that i = 1/2(a+β)
j = 1/2(a-β)

Then the multiplication table of i and j is as follows:

i j
i i 0
j 0 j

The algebra is, therefore, a mixture of two ordinary simple algebras of B. Peirce's form (a1).

326. Of three letters, there are six substitutions. Let Α, Β, Γ, Δ, Ε, Ζ be indeterminate coefficients. Then, the general expression of the algebra is equivalent to

Α Β Γ Δ Ε Ζ
Γ Α Β + Ε Ζ Δ
Β Γ Α Ζ Δ Ε

Or, denoting the six substitutions by α, β, γ, Δ, ε, ζ we may write the general expression as Αα + Ββ + Γγ + Δδ + Εε + Ζζ. There is an equation between these substitutions, namely:

α+β+γ = Δ+ε+ζ.

Assume 5 relatives h,i,j,k,l, such that

-h = 1/3 (α+β+γ)

i = 1/2 (α-ε)

j = 1/3 (β-γ+Δ-ζ)

k = 1/4 (-β+γ+Δ-ζ)

l = 1/6 (2α-β-γ-Δ+2ε-ζ)

In matricular form,

h= 1
1
1
1
1
1
1
1
1
2i= 1
-1
0
-1
1
0
0
0
0
3j= 1
-1
0
1
-1
0
-2
+2
0
4k= 1
1
-2
-1
-1
+2
0
0
0
6l= 1
1
-2
1
1
-2
-2
-2
+4

The multiplication table of h,i,j,k,l, is as follows:

h i j k l
h h 0 0 0 0
i 0 i j 0 0
j 0 0 0 i j
k 0 k l 0 0
l 0 0 0 k l

The algebra is a mixture of ordinary single algebra (a1) with the algebra of Hamilton's biquaternions (g4).

327. In six letters, there are twenty-four substitutions. Using the capital Greek letters for indeterminate coefficients, the general linear expression in these substitutions is equivalent to

ΑΛΓΒ ΕΖΘΗ ΙΛΚΜ
ΓΒΑΔ + ΖΕΗΘ + ΜΚΛΙ
ΒΓΔΑ ΘΗΕΖ ΛΙΜΚ
ΔΑΒΓ ΗΘΖΕ ΚΜΙΛ
 
ΝΟΠΞ ΡΥΣΤ ΦΧΨΩ
+ ΟΝΞΠ + ΣΤΡΥ + ΩΨΧΦ
ΞΠΟΝ ΥΡΤΣ ΨΩΦΧ
ΠΞΝΟ ΤΣΥΡ ΞΦΩΨ

Using the twenty-four small Greek letters to denote the twenty-four substitutions, so that the general linear expression is Aα+Bβ+, etc., an attentive observation of the above scheme will show that the following equations subsist:

α+β+γ+δ = ε+ζ+η+θ = ι+κ+λ+μ = ν+ξ+ο+π= ρ+σ+τ+υ = φ+χ+ψ+ω

Also,

α+β=ν+ξ ε+ζ=ν+ο ι+κ=ν+π
α+γ=ρ+σ ε+η=ρ+τ ι+ι=ρ+ν
α+δ=φ+χ ε+θ=φ+ψ ι+μ=φ+ω

It is plain that these equations are all independent, and not difficult to see that there are no more. Since they are fourteen in number, a ten-fold algebra is required to express the twenty-four substitutions.

Assume the ten relatives h, i, j, k, l, m, n, o, p, q, such that

h = 1/4 (α+β+γ+δ)

i = 1/4 (ε+ζ-η-θ)

j = 1/4 (α-β+γ-δ)

k = 1/4 (ι-κ-λ+μ)

l = 1/4 (ι-κ+λ-μ)

m = 1/4 (ε-ζ-η+θ)

n = 1/4 (α+β-γ-δ)

o = 1/4 (α-β-γ+δ)

p = 1/4 (ι+κ-λ-μ)

q = 1/4 (ε-ζ+η-θ)

In matricular form, these are as follows (where + is written for +1 and - for -1):

4h= +
+
+
+
+
+
+
+
+
+
+
+
+
+
+
+
 
4i= +
+
-
-
+
+
-
-
-
-
+
+
-
-
+
+
4j= +
+
-
-
-
-
+
+
+
+
-
-
-
-
+
+
4k= +
+
-
-
-
-
+
+
-
-
+
+
+
+
-
-
 
4l= +
-
+
-
+
-
+
-
-
+
-
+
-
+
-
+
4m= +
-
+
-
-
+
-
+
+
-
+
-
-
+
-
+
4n= +
-
+
-
-
+
-
+
-
+
-
+
+
-
+
-
 
4o= +
-
-
+
+
-
-
+
-
+
+
-
-
+
+
-
4p= +
-
-
+
-
+
+
-
+
-
-
+
-
+
+
-
4q= +
-
-
+
-
+
+
-
-
+
+
-
+
-
-
+

The multiplication table is as follows:

h i j k l m n o p q
h h 0 0 0 0 0 0 0 0 0
i 0 i j k 0 0 0 0 0 0
j 0 0 0 0 i j k 0 0 0
k 0 0 0 0 0 0 0 i j k
l 0 l m n 0 0 0 0 0 0
m 0 0 0 0 l m n 0 0 0
n 0 0 0 0 0 0 0 l m n
o 0 o p q 0 0 0 0 0 0
p 0 0 0 0 o p q 0 0 0
q 0 0 0 0 0 0 0 o p q

The algebra is, therefore, a mixture of ordinary algebra with that of my nonions.

†1 Of course, the representation of quaternions as linear expressions in substitutions of three letters, the sum of the coefficients being zero, is equivalent to finding a theorem of plane geometry corresponding to each theorem of solid geometry expressible by quaternions. For instance, let the three letters which are interchanged be the coördinates x, y, z, of a point in space. Then, as above, let

α=
x,y,z
x,y,z

β=
z,x,y
x,y,z

γ=
y,z,x
x,y,z

δ=
x,z,y
x,y,z

ε=
y,x,z
x,y,z

ζ=
z,y,x
x,y,z

Thus, β and γ represent the operations of rotation through one-third of a circumference, the one forward, the other backward, about an axis passing through the origin and the point (1, 1, 1); while δ, ε, ζ represent three perversions with reference to axes passing through the origin and the points (0, 1, -1), (1, -1, 0), and (1, 0, -1), respectively. Quaternions may be represented thus:

1 = 1/3 (2α - β - γ)

i = -1/3 (2ε - δ - ζ)(√-1)

j = (1/√3) (β - γ)

k = (1/√3) (δ - ζ)(√-1).

We have here a new geometrical interpretation of quaternions. Since the sum of the coefficients of the substitutions is equal to zero in the values of every one of the quaternion elements, it follows that under this interpretation any quaternion operating upon any point brings it into the plane.

x+y+z = 0.

Hence, every quaternion equation has an interpretation relating to points in this plane. The reason why a quaternion, which has a four-fold multiplicity, is no more than adequate to expressing operations upon points in space, is that the operations are of such a nature that different ones may have the same effect upon single points. But a real quaternion has no greater multiplicity than the real and imaginary points of a plane; and the geometrical effects of different real quaternions upon points in the plane x+y+z = 0 under the new interpretation are different upon all points except the origin.

For the axes of x, y, z, in trilinear coordinates, take three lines meeting in one point and equally inclined to one another. To plot the point x = a+b√-1, y = c+d√-1, z = c+f√-1, plot the point x = a, y = c, z = e in blue, and the point x = b, y = d, d = f in red. Then the effects of the different quaternion elements upon points in the plane x+y+z = 0 are as follows: 1 leaves every point unchanged. The vector i reverses the position of a blue point with reference to the line z = 0 and changes it to red, and reverses the position of a red point with reference to the line x = y and changes it to blue. The vector j rotates every point through a quadrant round the origin in the direction from x = 0 to y = z, without changing the color. The vector k reverses the position of a blue point with reference to the line through the origin that bisects the angle between y = 0 and y = z and changes it to red, and reverses the position of a red point with reference to the line through the origin that bisects the angle between x = 0 and x = z and changes it to blue.



Paper 12: The Logic of Relatives †1

328. A dual relative term, such as "lover," "benefactor," "servant," is a common name signifying a pair of objects. Of the two members of the pair, a determinate one is generally the first, and the other the second; so that if the order is reversed, the pair is not considered as remaining the same.

329. Let A, B, C, D, etc., be all the individual objects in the universe; then all the individual pairs may be arrayed in a block, thus:

A:A A:B A:C A:D etc.
B:A B:B B:C B:D etc.
C:A C:B C:C C:D etc.
D:A D:B D:C D:D etc.
etc. etc. etc. etc. etc.

A general relative may be conceived as a logical aggregate of a number of such individual relatives. Let l denote "lover"; then we may write

l = ΣiΣj(l)ij(I:J)

where (l)ij is a numerical coefficient, whose value is 1 in case I is a lover of J, and 0 in the opposite case, and where the sums are to be taken for all individuals in the universe.

330. Every relative term has a negative (like any other term) which may be represented by drawing a straight line over the sign for the relative itself. The negative of a relative includes every pair that the latter excludes, and vice versa. Every relative has also a converse, produced by reversing the order of the members of the pair. Thus, the converse of "lover" is "loved." The converse may be represented by drawing a curved line over the sign for the relative, thus: . It is defined by the equation

()ij = (l)ji.

The following formulæ are obvious, but important:

l̿ = l $ = l
~ = $
(lb) = () (lb) = ().

331. Relative terms can be aggregated and compounded like others. Using + for the sign of logical aggregation, and the comma for the sign of logical composition (Boole's multiplication, here to be called non-relative or internal multiplication), we have the definitions

(l+b)ij = (l)ij + (b)ij

(l,b)ij = (l)ij × (b)ij.

The first of these equations, however, is to be understood in a peculiar way: namely, the + in the second member is not strictly addition, but an operation by which

0+0 = 0 0+1 = 1+0 = 1+1 = 1.

Instead of (l)ij+(b)ij, we might with more accuracy write

00(l)ij+(b)ij

The main formulæ of aggregation and composition are


If ls and bs, then l+bs.
If sl and sb, then sl,b.

 

If l+bs, then ls and bs.
If sl,b, then sl and sb.

 

(l+b),sl,s+b,s.
(l+s),(b+s) ⤙ l,b+s.

The subsidiary formulæ need not be given, being the same as in non-relative logic.

332. We now come to the combination of relatives. Of these, we denote two by special symbols; namely, we write

lb for lover of a benefactor,

and

lb for lover of everything but benefactors. †1

The former is called a particular combination, because it implies the existence of something loved by its relate and a benefactor of its correlate. The second combination is said to be universal, because it implies the non-existence of anything except what is either loved by its relate or a benefactor of its correlate. The combination lb is called a relative product, lb a relative sum. The l and b are said to be undistributed in both, because if ls, then lbsb and lbsb; and if bs, then lbls and lbls. †2

333. The two combinations are defined by the equations

(lb)ij = Σx(l)ix(b)xj

(lb)ij = Πx {(l)ix+(b)xj} †3

The sign of addition in the last formula has the same signification as in the equation defining non-relative multiplication. †4

334. Relative addition and multiplication are subject to the associative law. That is,

l†(bs) = (lb)†s

l(bs) = (lb)s.

Two formulæ so constantly used that hardly anything can be done without them are

l(bs) ⤙ lbs,

(lb)slbs.

The former asserts that whatever is lover of an object that is benefactor of everything but a servant, stands to everything but servants in the relation of lover of a benefactor. The latter asserts that whatever stands to any servant in the relation of lover of everything but its benefactors, is a lover of everything but benefactors of servants. The following formulæ are obvious and trivial:

ls+bs ⤙ (l+b)s †1

l,bs ⤙ (ls),(bs). †2

Unobvious and important, however, are these:

(l+b)sls+bs †1

(ls),(bs) ⤙ l,bs. †2

335. There are a number of curious development formulæ. †3 Such are

(l,b)s = Πp{l(s,p)+b(s,)}

l(b,s) = Πp{(l,p)b+(l,)s}

(l+b)†s = Σp{[l†(s+p)],[b†(s+)]}

l†(b+s) = Σp{[(l+p)†b],[(l+)†s]}

The summations and multiplications denoted by Σ and Π are to be taken non-relatively, and all relative terms are to be successively substituted for p.

336. The negatives of the combinations follow these rules:

~(l+b) = , ~(l,b) = +
~(lb) = ~(lb) =

337. The converses of combinations are as follows:

$(l+b) = + $(l,b) = ,
$(lb) = $(lb) =

338. †4 Individual dual relatives are of two types,

A:A and A:B.

Relatives containing no pair of an object with itself are called alio-relatives as opposed to self-relatives. The negatives of alio-relatives pair every object with itself. Relatives containing no pair of an object with anything but itself are called concurrents as opposed to opponents. The negatives of concurrents pair every object with every other.

339. There is but one relative which pairs every object with itself and with every other. It is the aggregate of all pairs, and is denoted by ∞. It is translated into ordinary language by "coexistent with." Its negative is 0. There is but one relative which pairs every object with itself and none with any other. It is

(A:A) + (B:B) + (C:C) + etc.;

is denoted by 1, and in ordinary language is "identical with —." Its negative, denoted by 𝖓, is "other than —," or "not."

340. No matter what relative term x may be, we have

0 ⤙ x x ⤙ ∞.

341. Hence, obviously

x + 0 = x x,∞ = x
x + ∞ = ∞ x,0 = 0.

The last formulæ hold for the relative operations; thus,

x † ∞ = ∞ x0 = 0.
∞ † x = ∞ 0x = 0.

The formulæ

x + 0 = x x,∞ = x

also hold if we substitute the relative operations, and also 1 for ∞, and 𝖓 for 0; thus,

x † 𝖓 = x x 1 = x.
𝖓 † x = x 1 x = x.

We have also

l + = ∞, l, = 0. †P1

To these partially correspond the following pair of highly important formulæ:

1 ⤙ l†$ †1 l$ ⤙ 𝖓. †2

342. The logic of relatives is highly multiform; it is characterized by innumerable immediate inferences, and by various distinct conclusions from the same sets of premisses. An example of the first character is afforded by Mr. Mitchell's F1v following from F1v'. †3 As an instance of the second, take the premisses,

Every man is a lover of an animal;

and

Every woman is a lover of a non-animal.

From these we can equally infer that

Every man is a lover of something which stands to each woman in the relation of not being the only thing loved by her,

and that

Every woman is a lover of something which stands to each man in the relation of not being the only thing loved by him.

The effect of these peculiarities is that this algebra cannot be subjected to hard and fast rules like those of the Boolian calculus; and all that can be done in this place is to give a general idea of the way of working with it. The student must at the outset disabuse himself of the notion that the chief instruments of algebra are the inverse operations. General algebra hardly knows any inverse operations. When an inverse operation is identical with a direct operation with an inverse quantity (as subtraction is the addition of the negative, and as division is multiplication by the reciprocal), it is useful; otherwise it is almost always useless. In ordinary algebra, we speak of the "principal value" of the logarithm, etc., which is a direct operation substituted for an indefinitely ambiguous inverse operation. The elimination and transposition in this algebra really does depend, however, upon formulæ quite analogous to the

x + (-x) = 0 x × (1/x) = 1,

of arithmetical algebra. These formulæ are

l, = 0 l$ ⤙ 𝖓
l+ = ∞ 1 ⤙ l†$.

For example, to eliminate s from the two propositions

1 ⤙ l 1 ⤙ s̆b,

we relatively multiply them in such an order as to bring the two s's together, and then apply the second of the above formulæ, thus:

1 ⤙ ls̆bl𝖓b.

This example shows the use of the association formulæ in bringing letters together. Other formulæ of great importance for this purpose are

(bl)sbls b(ls) ⤙ bls.

The distribution formulæ are also useful for this purpose.

343. When the letter to be eliminated has thus been replaced by one of the four relatives — 0, ∞, 1, 𝖓 — the replacing relative can often be got rid of by means of one of the formulæ

l+0 = l l,∞ = l
l†𝖓 = 𝖓†l = l l1 = 1l = l.

344. When we have only to deal with universal propositions, it will be found convenient so to transpose everything from subject to predicate as to make the subject 1. Thus, if we have given lb, we may relatively add $ to both sides; whereupon we have

1 ⤙ l†$b†$.

Every proposition will then be in one of the forms

1 ⤙ bl 1 ⤙ bl.

With a proposition of the form 1 ⤙ bl, we have the right (1) to transpose the terms, and (2) to convert the terms. Thus, the following are equivalent:

1 ⤙ bl
1 ⤙ lb 1 ⤙
1 ⤙ .

With a proposition of the form 1 ⤙ bl, we have only the right to convert the predicate giving 1 ⤙ .

With three terms, there are four forms of universal propositions, namely:

1 ⤙ lbs 1 ⤙ l(bs) 1 ⤙ lbs 1 ⤙ l b s.

Of these, the third is an immediate inference from the second.

345. By way of illustration, we may work out the syllogisms whose premisses are the propositions of the first order referred to in Note A. †1 Let a and c be class terms, and let β be a group of characters. Let p be the relative "possessing as a character." The non-relative terms are to be treated as relatives—a, for instance, being considered as "a coexistent with" and ă as "coexistent with a that is." Then, the six forms of affirmative propositions of the first order are

1 ⤙ ăp†β
1 ⤙ ă(p†β) 1 ⤙ (ăp
1 ⤙ ăp†β 1 ⤙ ăpβ
1 ⤙ ăpβ. †2

346. The various kinds of syllogism †3 are as follows:

1. Premisses: 1 ⤙ ăp†β 1 ⤙ p†β̅.

Convert one of the premisses and multiply,

1 ⤙ (ăp†β)($β̅†c) ăp†β$β̅†c
ăp†𝖓†căpc.

The treatment would be the same if one or both of the premisses were negative: that is, contained in place of p.

2. Premisses: 1 ⤙ ăp†β 1 ⤙ (p†β̅).

We have

1 ⤙ (ăp†β)($β̅†)c ⤙ (ăp)c.

The same with negatives.

3. Premisses: 1 ⤙ ă(p†β) 1 ⤙ (p†β̅).
1 ⤙ ă(p†β)($β̅†)că(p)c.

The same with negatives.

4. Premisses: 1 ⤙ ăp†β 1 ⤙ (p)β̅.
1 ⤙ (ăp†β)$β̅(c) ⤙ (ăp†β$β̅)(c) ⤙ (ăp)(c)

If one of the premisses, say the first, were negative, we should obtain a similar conclusion —

1 ⤙ (ă)(c);

but from this again p could be eliminated, giving

1 ⤙ ă†c, or āc.
5. Premisses: 1 ⤙ ă†(p†β) 1 ⤙ (p)β̅.
1 ⤙ ă†(p†β)β(c) ⤙ ăp(c).

If either premiss were negative, p could be eliminated, giving 1 ⤙ ăc, or some a is c.

6. Premisses: 1 ⤙ (ăp 1 ⤙ (p)β̅.
1 ⤙ (ăp)β$β̅(c) ⤙ (ăp)𝖓(c).
7. Premisses: 1 ⤙ ăp†β 1 ⤙ (c̆p†β̅.
1 ⤙ (ăp†β)($β̅†p̆c) ⤙ ăpp̆c.
8. Premisses: 1 ⤙ ă(p†β) 1 ⤙ c̆p†β̅.
1 ⤙ ă(p†β)($β̅†p̆c) ⤙ ă(pp̆c).
9. Premisses: 1 ⤙ (ăp 1 ⤙ c̆p†β̅.
1 ⤙ (ăp)β($β̅†p̆c) ⤙ (ăp)p̆c.

If one premiss is negative, we have the further conclusion 1 ⤙ ăc.

10. Premisses: 1 ⤙ ăp†β 1 ⤙ c̆p†β̅.
1 ⤙ (ăp†β)($β̅†p̆c) ⤙ ăpp̆c.
11. Premisses: 1 ⤙ ăp†β 1 ⤙ pβ̅
1 ⤙ (ăp†β)($β̅c) ⤙ (ăp)c.

We might also conclude

1 ⤙ ăp†𝖓c;

but this conclusion is an immediate inference from the other; for

(ăp)c ⤙ (ăp)(1†𝖓)c ⤙ (ăp)1†𝖓căp†𝖓c.

If one premiss is negative, we have the further conclusion 1 ⤙ ăc.

12. Premisses: 1 ⤙ ă(p†β) 1 ⤙ pβ̅.
1 ⤙ ă(p†β)($β̅c) ⤙ ă(pc).

If one premiss is negative, we have the further inference 1 ⤙ ăc.

13. Premisses: 1 ⤙ (ăp 1 ⤙ pβ̅.
1 ⤙ (ăp)β($β̅†c) ⤙ (ăp)(𝖓c).
14. Premisses: 1 ⤙ ăp†β 1 ⤙ pβ̅.
1 ⤙ (ăp†β)($β̅c) ⤙ ăpc.

If one premiss is negative, we have the further spurious †1 inference 1 ⤙ ă𝖓†c.

15. Premisses: 1 ⤙ ăpβ 1 ⤙ pβ̅.
1 ⤙ (ăpβ)($β̅c) ⤙ ăp(𝖓c).

We can also infer 1 ⤙ (ăp𝖓)c.

16. Premisses: 1 ⤙ ăp†β 1 ⤙ c̆pβ̅.
1 ⤙ (ăp†β)$β̅p̆c ⤙ (ăp)p̆c.

If one premiss is negative, we can further infer 1 ⤙ ăc.

17. Premisses: 1 ⤙ ă(p†β) 1 ⤙ c̆pβ̅.
1 ⤙ ă(p†β)$β̅p̆căpp̆c.

If one premiss is negative, we have the further spurious conclusion 1 ⤙ ă𝖓c.

18. Premisses: 1 ⤙ (ăp 1 ⤙ c̆pβ̅.
1 ⤙ (ăp)β$β̅p̆c ⤙ (ăp)𝖓p̆c.
19. Premisses: 1 ⤙ ăp†β 1 ⤙ c̆pβ̅.
1 ⤙ (ăp†β)$β̅p̆căpp̆c.

If one premiss is negative, we further conclude 1 ⤙ ă𝖓c.

20. Premisses: 1 ⤙ ăpβ 1 ⤙ c̆pβ̅.
1 ⤙ (ăpβ)$β̅p̆c ⤙ (ăp𝖓)p̆c.
21. Premisses: 1 ⤙ ăpβ 1 ⤙ c̆pβ̅.
1 ⤙ ăpβ$β̅p̆căp𝖓p̆c.

347. When we have to do with particular propositions, we have the proposition ∞ ~⤙ 0, or "something exists"; for every particular proposition implies this. Then every proposition can be put into one or other of the four forms

∞ ⤙ 0†l†0

∞ ⤙ (0†l)∞

∞ ⤙ 0†l

∞ ⤙ ∞l

Each of these propositions immediately follows from the one above it. The enveloped expressions which form the predicates have the remarkable property that each is either 0 or ∞ . This fact gives extraordinary freedom in the use of the formulæ. In particular, since if anything not zero is included under such an expression, the whole universe is included, it will be quite unnecessary to write the ∞ ⤙ which begins every proposition.

348. Suppose that f and g are general relatives signifying relations of things to times. Then, Dr. Mitchell's †1 six forms of two dimensional propositions appear thus:

F11 = 0†f†0

F1v = 0†f

Fu1 = ∞f†0

F1v' = (0†f)∞

Fu'1 = ∞(f†0)

Fuv = ∞f∞.

It is obvious that l † 0 ⤙ l, for

l † 0 ⤙ (l † 0)∞ ⤙ l † 0 ∞ ll † 𝖓 ⤙ l.

If then we have 0†f†0 as one premiss, and the other contains g, we may substitute for g the product (f,g).

gg,∞ ⤙ g,(0†f†0) ⤙ g,f.

349. From the two premisses

∞(f†0) and 0†g∞,

by the application of the formulæ

ls,(b) ⤙ (l,b)s

sl,(b) ⤙ s(l,b),

we have

{∞(f†0)},(0†g∞) ⤙ ∞{(f†0),g∞} ⤙ ∞(f,g)∞.

These formulæ give the first column of Dr. Mitchell's rule on page 90.

350. The following formulæ may also be applied:

1. (0†f†0),(0†g†0) = 0†(f,g)†0.
2. (0†f)∞ (0††0) ⤙ (0†f)(†0).
3. (0†f)∞ ∞(†0) = (0†f)(ğ†0)+(0†f)𝖓(†0).
4 (0†f)∞ (0†)∞ ⤙ (0†f)∞.
5. (0†f†0) (0†∞) = 0†(ğf,f)†0.
6. (0†f)∞ (0†∞) = (0†ğf,f)∞.
7. (0†f)∞,(0†g∞) = (0†f,g∞)∞.
8. (0†f∞) (0†g∞) = 0†(f,ğf)∞.
9. (0†f∞),(0†g∞) = 0†f∞,g∞.
10. (0†f†0)∞ = 0†(fğf,f)†0.
11. (0†f)∞ ∞ = (0†f)∞ (0†f)𝖓∞.
12. (0†f∞) ∞ = (0†f∞)+(0†f𝖓∞)
13. ∞f∞ ∞ = ∞f∞ + ∞f𝖓∞.

351. When the relative and non-relative operations occur together, the rules of the calculus become pretty complicated. In these cases, as well as in such as involve plural relations (subsisting between three or more objects), it is often advantageous to recur to the numerical coefficients mentioned in 329. Any proposition whatever is equivalent to saying that some complexus of aggregates †P1 and products of such numerical coefficients is greater than zero. Thus,

ΣiΣjlij > 0

means that something is a lover of something; and

ΠiΣjlij > 0

means that everything is a lover of something. We shall, however, naturally omit, in writing the inequalities, the > 0 which terminates them all; and the above two propositions will appear as

ΣiΣjlij and ΠiΣjlij.

352. The following are other examples:

ΠiΣj(l)ij(b)ij

means that everything is at once a lover and a benefactor of something.

ΠiΣj(l)ij(b)ji

means that everything is a lover of a benefactor of itself.

ΣiΣkΠj(lij + bjk)

means that there is something which stands to something in the relation of loving everything except benefactors of it. †1

353. Let α denote the triple relative "accuser to — of —," and ε the triple relative "excuser to — of —. Then,

ΣiΠjΣk(α)ijk(ε)jki

means that an individual i can be found, such, that taking any individual whatever, j, it will always be possible so to select a third individual, k, that i is an accuser to j of k, and j an excuser to k of i.

354. Let π denote "preferrer to — of —." Then,

ΠiΣjΣk(α)ijkjki + πkij)

means that, having taken any individual i whatever, it is always possible so to select two, j and k, that i is an accuser to j of k, and also is either excused by j to k or is something to which j is preferred by k.

355. When we have a number of premisses expressed in this manner, the conclusion is readily deduced by the use of the following simple rules. In the first place, we have

ΣiΠj ⤙ ΠjΣi.

In the second place, we have the formulæ

iφ(i)} {ΠjΨ(j)} = Πi{φ(i)·Ψ(i)}.

iφ(i)} {ΣjΨ(j)} ⤙ Σi{φ(i)·Ψ(i)}.

In the third place, since the numerical coefficients are all either zero or unity, the Boolian calculus is applicable to them.

356. The following is one of the simplest possible examples. Required to eliminate servant from these two premisses:

First premiss. There is somebody who accuses everybody to everybody, unless the latter is loved by some person that is servant of all not accused to him. †1

Second premiss. There are two persons, the first of whom excuses everybody to everybody, unless the unexcused be benefited by, without the person to whom he is unexcused being a servant of, the second.

These premisses may be written thus:

ΣhΠiΣjΠkhik + sjklji)

ΣuΣvΠxΠyuyx+yvbvx).

The second yields the immediate inference,

ΠxΣuΠyΣvuyx + yvbvx).

Combining this with the first, we have

ΣxΣuΣyΣvuyx+yvbvx)(axuv+syvlyu).

Finally, applying the Boolian calculus, we deduce the desired conclusion

ΣxΣuΣyΣvuyxxuvuyxlyuxuvbvx).

The interpretation of this is that either there is somebody excused by a person to whom he accuses somebody, or somebody excuses somebody to his (the excuser's) lover, or somebody accuses his own benefactor.

357. The procedure may often be abbreviated by the use of operations intermediate between Π and Σ. Thus, we may use Π', Π'', etc. to mean the products for all individuals except one, except two, etc. Thus,

Πij''(lij + bji)

will mean that every person except one is a lover of everybody except its benefactors, and at most two non-benefactors. In the same manner, Σ', Σ'', etc., will denote the sums of all products of two, of all products of three, etc. Thus,

Σi''(lii)

will mean that there are at least three things in the universe that are lovers of themselves. It is plain that if m < n, we have

Πm ⤙ Πn Σn ⤙ Σm.
imφi)(ΣjnΨj)⤙ Σin-mi·Ψi)
imφi)(ΠjnΨj)⤙ Πim+ni·Ψi)

358. Mr. Schlötel has written to the London Mathematical Society, †1 accusing me of having, in my Algebra of Logic, plagiarized from his writings. He has also written to me to inform me that he has read that Memoir with "heitere Ironie," and that Professor Drobisch, the Berlin Academy, and I constitute a "liederliche Kleeblatt," with many other things of the same sort. Up to the time of publishing my Memoir, I had never seen any of Mr. Schlötel's writings; I have since procured his Logik, †2 and he has been so obliging as to send me two cuttings from his papers, thinking, apparently, that I might be curious to see the passages that I had appropriated. But having examined these productions, I find no thought in them that I ever did, or ever should be likely to put forth as my own.



Paper 13: On the Algebra of Logic

A Contribution to the Philosophy of Notation †1

§1. Three Kinds of Signs †2

359. Any character or proposition either concerns one subject, two subjects, or a plurality of subjects. For example, one particle has mass, two particles attract one another, a particle revolves about the line joining two others. A fact concerning two subjects is a dual character or relation; but a relation which is a mere combination of two independent facts concerning the two subjects may be called degenerate, just as two lines are called a degenerate conic. In like manner a plural character or conjoint relation is to be called degenerate if it is a mere compound of dual characters.

360. A sign is in a conjoint relation to the thing denoted and to the mind. If this triple relation is not of a degenerate species, the sign is related to its object only in consequence of a mental association, and depends upon a habit. Such signs are always abstract and general, because habits are general rules to which the organism has become subjected. They are, for the most part, conventional or arbitrary. They include all general words, the main body of speech, and any mode of conveying a judgment. For the sake of brevity I will call them tokens. †3

361. But if the triple relation between the sign, its object, and the mind, is degenerate, then of the three pairs

sign object
sign mind
object mind

two at least are in dual relations which constitute the triple relation. One of the connected pairs must consist of the sign and its object, for if the sign were not related to its object except by the mind thinking of them separately, it would not fulfill the function of a sign at all. Supposing, then, the relation of the sign to its object does not lie in a mental association, there must be a direct dual relation of the sign to its object independent of the mind using the sign. In the second of the three cases just spoken of, this dual relation is not degenerate, and the sign signifies its object solely by virtue of being really connected with it. Of this nature are all natural signs and physical symptoms. I call such a sign an index, a pointing finger being the type of the class.

The index asserts nothing; it only says "There!" It takes hold of our eyes, as it were, and forcibly directs them to a particular object, and there it stops. Demonstrative and relative pronouns are nearly pure indices, because they denote things without describing them; so are the letters on a geometrical diagram, and the subscript numbers which in algebra distinguish one value from another without saying what those values are.

362. The third case is where the dual relation between the sign and its object is degenerate and consists in a mere resemblance between them. I call a sign which stands for something merely because it resembles it, an icon. Icons are so completely substituted for their objects as hardly to be distinguished from them. Such are the diagrams of geometry. A diagram, indeed, so far as it has a general signification, is not a pure icon; but in the middle part of our reasonings we forget that abstractness in great measure, and the diagram is for us the very thing. So in contemplating a painting, there is a moment when we lose the consciousness that it is not the thing, the distinction of the real and the copy disappears, and it is for the moment a pure dream — not any particular existence, and yet not general. At that moment we are contemplating an icon.

363. I have taken pains to make my distinction †P1 of icons, indices, and tokens clear, in order to enunciate this proposition: in a perfect system of logical notation signs of these several kinds must all be employed. Without tokens there would be no generality in the statements, for they are the only general signs; and generality is essential to reasoning. Take, for example, the circles by which Euler represents the relations of terms. They well fulfill the function of icons, but their want of generality and their incompetence to express propositions must have been felt by everybody who has used them. †1 Mr. Venn †2 has, therefore, been led to add shading to them; and this shading is a conventional sign of the nature of a token. In algebra, the letters, both quantitative and functional, are of this nature. But tokens alone do not state what is the subject of discourse; and this can, in fact, not be described in general terms; it can only be indicated. The actual world cannot be distinguished from a world of imagination by any description. Hence the need of pronoun and indices, and the more complicated the subject the greater the need of them. The introduction of indices into the algebra of logic is the greatest merit of Mr. Mitchell's system. †P1 He writes F1 to mean that the proposition F is true of every object in the universe, and Fu to mean that the same is true of some object. †3 This distinction can only be made in some such way as this. Indices are also required to show in what manner other signs are connected together. With these two kinds of signs alone any proposition can be expressed; but it cannot be reasoned upon, for reasoning consists in the observation that where certain relations subsist certain others are found, and it accordingly requires the exhibition of the relations reasoned within an icon. It has long been a puzzle how it could be that, on the one hand, mathematics is purely deductive in its nature, and draws its conclusions apodictically, while on the other hand, it presents as rich and apparently unending a series of surprising discoveries as any observational science. Various have been the attempts to solve the paradox by breaking down one or other of these assertions, but without success. The truth, however, appears to be that all deductive reasoning, even simple syllogism, involves an element of observation; namely, deduction consists in constructing an icon or diagram the relations of whose parts shall present a complete analogy with those of the parts of the object of reasoning, of experimenting upon this image in the imagination, and of observing the result so as to discover unnoticed and hidden relations among the parts. For instance, take the syllogistic formula,

All M is P
S is M
S is P.

This is really a diagram of the relations of S, M, and P. The fact that the middle term occurs in the two premisses is actually exhibited, and this must be done or the notation will be of no value. As for algebra, the very idea of the art is that it presents formulæ which can be manipulated, and that by observing the effects of such manipulation we find properties not to be otherwise discerned. In such manipulation, we are guided by previous discoveries which are embodied in general formulæ. These are patterns which we have the right to imitate in our procedure, and are the icons par excellence of algebra. The letters of applied algebra are usually tokens, but the x, y, z, etc., of a general formula, such as

(x+y)z = xz + yz,

are blanks to be filled up with tokens, they are indices of tokens. Such a formula might, it is true, be replaced by an abstractly stated rule (say that multiplication is distributive); but no application could be made of such an abstract statement without translating it into a sensible image.

364. In this paper, I purpose to develop an algebra adequate to the treatment of all problems of deductive logic, showing as I proceed what kinds of signs have necessarily to be employed at each stage of the development. I shall thus attain three objects. The first is the extension of the power of logical algebra over the whole of its proper realm. The second is the illustration of principles which underlie all algebraic notation. The third is the enumeration of the essentially different kinds of necessary inference; for when the notation which suffices for exhibiting one inference is found inadequate for explaining another, it is clear that the latter involves an inferential element not present to the former. Accordingly, the procedure contemplated should result in a list of categories of reasoning, the interest of which is not dependent upon the algebraic way of considering the subject. I shall not be able to perfect the algebra sufficiently to give facile methods of reaching logical conclusions: I can only give a method by which any legitimate conclusion may be reached and any fallacious one avoided. But I cannot doubt that others, if they will take up the subject, will succeed in giving the notation a form in which it will be highly useful in mathematical work. I even hope that what I have done may prove a first step toward the resolution of one of the main problems of logic, that of producing a method for the discovery of methods in mathematics.

§2. Non-Relative Logic

365. According to ordinary logic, a proposition is either true or false, and no further distinction is recognized. This is the descriptive conception, as the geometers say; the metric conception would be that every proposition is more or less false, and that the question is one of amount. At present we adopt the former view.

366. Let propositions be represented by quantities. Let v and f be two constant values, and let the value of the quantity representing a proposition be v if the proposition is true and be f if the proposition is false. Thus, x being a proposition, the fact that x is either true or false is written

(x - f)(v - x) = 0. †1
So (x - f)(v - y) = 0

will mean that either x is false or y is true. This may be said to be the same as 'if x is true, y is true.' A hypothetical proposition, generally, is not confined to stating what actually happens, but states what is invariably true throughout a universe of possibility. The present proposition is, however, limited to that one individual state of things, the Actual.

367. We are, thus, already in possession of a logical notation, capable of working syllogism. Thus, take the premisses, 'if x is true, y is true,' and 'if y is true, z is true.' These are written

(x - f)(v - y) = 0

(y - f)(v - z) = 0.

Multiply the first by (v - z) and the second by (x - f) and add. We get

(x - f)(v - f)(v - z) = 0,

or dividing by v - f, which cannot be 0,

(x - f)(v - z) = 0;

and this states the syllogistic conclusion, "if x is true, z is true."

368. But this notation shows a blemish in that it expresses propositions in two distinct ways, in the form of quantities, and in the form of equations; and the quantities are of two kinds, namely those which must be either equal to f or to v, and those which are equated to zero. To remedy this, let us discard the use of equations, and perform no operations which can give rise to any values other than f and v.

369. Of operations upon a simple variable, we shall need but one. For there are but two things that can be said about a single proposition, by itself; that it is true and that it is false,

x = v and x = f.

The first equation is expressed by x itself, the second by any function, φ, of x, fulfilling the conditions

φv = f φf = v.

The simplest solution of these equations is

φx = f + v - x.

A product of n factors of the two forms (x - f) and (v - y), if not zero, equals (v - f)n. Write P for the product. Then v - (P/((v - f)n-1)) is the simplest function of the variables which becomes v when the product vanishes and f when it does not. By this means any proposition relating to a single individual can be expressed.

370. If we wish to use algebraical signs with their usual significations, the meanings of the operations will entirely depend upon those of f and v. Boole †1 chose v = 1, f = 0. This choice gives the following forms:

f + v - x = 1 - x

which is best written .

v - ((x-f)(v-y)/(v-f)) = 1 - x + xy = ~(x).

v - ((v-x)(v-y)/(v-f)) = x + y - xy †2

v - ((v-x)(v-y)(v-z)/((v-f)2)) = x+y+z-xy-xz-yz+xyz

v - ((x-f)(y-f)/(v-f)) = 1-xy = ~(xy) †3

371. It appears to me that if the strict Boolian system is used, the sign + ought to be altogether discarded. Boole and his adherent, Mr. Venn (whom I never disagree with without finding his remarks profitable), prefer to write x+x̄y in place of ~(). I confess I do not see the advantage of this, for the distributive principle holds equally well when written

~()z = ~(~(xz)~(yz)) †4

~(~(xy)) = ~().~(). †5

The choice of v = 1, f = 0, is agreeable to the received measurement of probabilities. But there is no need, and many times no advantage, in measuring probabilities in this way. I presume that Boole, in the formation of his algebra, at first considered the letters as denoting propositions or events. As he presents the subject, they are class-names; but it is not necessary so to regard them. Take, for example, the equation

t = n + h f,

which might mean that the body of taxpayers is composed of all the natives, together with householding foreigners. We might reach the signification by either of the following systems of notation, which indeed differ grammatically rather than logically.


Sign.
Signification.
1st System.
Signification.
2nd System.
t
n
h
f
Taxpayer.
Native.
Householder.
Foreigner.
He is a Taxpayer.
He is a Native.
He is a Householder.
He is a Foreigner.

There is no index to show who the "He" of the second system is, but that makes no difference. To say that he is a taxpayer is equivalent to saying that he is a native or is a householder and a foreigner. In this point of view, the constants 1 and 0 are simply the probabilities, to one who knows, of what is true and what is false; and thus unity is conferred upon the whole system.

372. For my part, I prefer for the present not to assign determinate values to f and v, nor to identify the logical operations with any special arithmetical ones, leaving myself free to do so hereafter in the manner which may be found most convenient. Besides, the whole system of importing arithmetic into the subject is artificial, and modern Boolians do not use it. The algebra of logic should be self-developed, and arithmetic should spring out of logic instead of reverting to it. Going back to the beginning, let the writing of a letter by itself mean that a certain proposition is true. This letter is a token. There is a general understanding that the actual state of things or some other is referred to. This understanding must have been established by means of an index, and to some extent dispenses with the need of other indices. The denial of a proposition will be made by writing a line over it.

373. I have elsewhere †1 shown that the fundamental and primary mode of relation between two propositions is that which we have expressed by the form

v - ((x-f)(v-y)/(v-f)).

We shall write this xy,
which is also equivalent to (x-f)(v-y) = 0.

It is stated above that this means "if x is true, y is true." But this meaning is greatly modified by the circumstance that only the actual state of things is referred to.

374. To make the matter clear, it will be well to begin by defining the meaning of a hypothetical proposition, in general. What the usages of language may be does not concern us; language has its meaning modified in technical logical formulæ as in other special kinds of discourse. The question is what is the sense which is most usefully attached to the hypothetical proposition in logic? Now, the peculiarity of the hypothetical proposition is that it goes out beyond the actual state of things and declares what would happen were things other than they are or may be. The utility of this is that it puts us in possession of a rule, say that "if A is true, B is true," such that should we hereafter learn something of which we are now ignorant, namely that A is true, then, by virtue of this rule, we shall find that we know something else, namely, that B is true. There can be no doubt that the Possible, in its primary meaning, is that which may be true for aught we know, that whose falsity we do not know. †1 The purpose is subserved, then, if throughout the whole range of possibility, in every state of things in which A is true, B is true too. The hypothetical proposition may therefore be falsified by a single state of things, but only by one in which A is true while B is false. States of things in which A is false, as well as those in which B is true, cannot falsify it. If, then, B is a proposition true in every case throughout the whole range of possibility, the hypothetical proposition, taken in its logical sense, ought to be regarded as true, whatever may be the usage of ordinary speech. If, on the other hand, A is in no case true, throughout the range of possibility, it is a matter of indifference whether the hypothetical be understood to be true or not, since it is useless. But it will be more simple to class it among true propositions, because the cases in which the antecedent is false do not, in any other case, falsify a hypothetical. This, at any rate, is the meaning which I shall attach to the hypothetical proposition in general, in this paper.

375. The range of possibility is in one case taken wider, in another narrower; in the present case it is limited to the actual state of things. Here, therefore, the proposition

ab

is true if a is false or if b is true, but is false if a is true while b is false. But though we limit ourselves to the actual state of things, yet when we find that a formula of this sort is true by logical necessity, it becomes applicable to any single state of things throughout the range of logical possibility. For example, we shall see that from x ~⤙ y we can infer zx. This does not mean that because in the actual state of things x is true and y false, therefore in every state of things either z is false or x true; but it does mean that in whatever state of things we find x true and y false, in that state of things either z is false or x is true. In that sense, it is not limited to the actual state of things, but extends to any single state of things.

376. The first icon of algebra is contained in the formula of identity

xx.

This formula does not of itself justify any transformation, any inference. It only justifies our continuing to hold what we have held (though we may, for instance, forget how we were originally justified in holding it).

377. The second icon is contained in the rule that the several antecedents of a consequentia may be transposed; that is, that from

x ⤙ (yz)
we can pass to y ⤙ (xz).

This is stated in the formula

{x ⤙ (yz)} ⤙ {y ⤙ (xz)}.

Because this is the case, the brackets may be omitted, and we may write

yxz.

By the formula of identity

(xy) ⤙ (xy);

and transposing the antecedents

x ⤙ {(xy) ⤙ y}

or, omitting the unnecessary brackets

x ⤙ (xy) ⤙ y.

This is the same as to say that if in any state of things x is true, and if the proposition "if x, then y" is true, then in that state of things y is true. This is the modus ponens of hypothetical inference, and is the most rudimentary form of reasoning. †1

378. To say that (xx) is generally true is to say that it is so in every state of things, say in that in which y is true; so that we may write

y ⤙ (xx),

and then, by transposition of antecedents,

x ⤙ (yx),

or from x we may infer yx.

379. The third icon is involved in the principle of the transitiveness of the copula, which is stated in the formula

(xy) ⤙ (yz) ⤙ xz. †2

According to this, if in any case y follows from x and z from y, then z follows from x. †3 This is the principle of the syllogism in Barbara.

380. We have already seen that from x follows yx. Hence, by the transitiveness of the copula, if from yx follows z, then from x follows z, or from

(yx) ⤙ z
follows xz,
or {(yx) ⤙ z} ⤙ xz.

381. The original notation xy served without modification to express the pure formula of identity. An enlargement of the conception of the notation so as to make the terms themselves complex was required to express the principle of the transposition of antecedents; and this new icon brought out new propositions. The third icon introduces the image of a chain of consequence. We must now again enlarge the notation so as to introduce negation. We have already seen that if a is true, we can write xa, whatever x may be. Let b be such that we can write bx whatever x may be. Then b is false. We have here a fourth icon, which gives a new sense to several formulæ. Thus the principle of the interchange of antecedents is that from

x ⤙ (yz)
we can infer y ⤙ (xz).

Since z is any proposition we please, this is as much as to say that if from the truth of x the falsity of y follows, then from the truth of y the falsity of x follows.

382. Again the formula

x ⤙ {(xy) ⤙ y}

is seen to mean that from x, we can infer that anything we please follows from that things following from x, and a fortiori from everything following from x. This is, therefore, to say that from x follows the falsity of the denial of x; which is the principle of contradiction.

383. Again the formula of the transitiveness of the copula, or

{xy} ⤙ {(yz) ⤙ (xz)}

is seen to justify the inference

xy

.

The same formula justifies the modus tollens,

xy

.

So the formula {(yx) ⤙ z} ⤙ (xz) shows that from the falsity of yx the falsity of x may be inferred.

All the traditional moods of syllogism can easily be reduced to Barbara by this method.

384. A fifth icon is required for the principle of excluded middle and other propositions connected with it. One of the simplest formulæ of this kind is

{(xy) ⤙ x} ⤙ x.

This is hardly axiomatical. That it is true appears as follows. It can only be false by the final consequent x being false while its antecedent (xy) ⤙ x is true. If this is true, either its consequent, x, is true, when the whole formula would be true, or its antecedent xy is false. But in the last case the antecedent of xy, that is x, must be true. †P1

From the formula just given, we at once get

{(xy) ⤙ α} ⤙ x,

where the a is used in such a sense that (xy) ⤙ α means that from (xy) every proposition follows. With that understanding, the formula states the principle of excluded middle, that from the falsity of the denial of x follows the truth of x.

385. The logical algebra thus far developed contains signs of the following kinds:

First, tokens; signs of simple propositions, as t for 'He is a taxpayer,' etc.

Second, the single operative sign ⤙; also of the nature of a token.

Third, the juxtaposition of the letters to the right and left of the operative sign. This juxtaposition fulfils the function of an index, in indicating the connections of the tokens.

Fourth, the parentheses, subserving the same purpose.

Fifth, the letters α, β, etc. which are indices of no matter what tokens, used for expressing negation.

Sixth, the indices of tokens, x, y, z, etc., used in the general formulæ.

Seventh, the general formulæ themselves, which are icons, or exemplars of algebraic proceedings.

Eighth, the fourth icon which affords a second interpretation of the general formulæ.

386. We might dispense with the fifth and eighth species of signs — the devices by which we express negation — by adopting a second operational sign ~⤙, such that

x ~⤙ y

should mean that x = v, y = f. With this, we should require new indices of connections, and new general formulæ. Possibly this might be the preferable notation. We should thus have two operational signs but no sign of negation. The forms of Boolian algebra hitherto used, have either two operational signs and a special sign of negation, or three operational signs. One of the operational signs is in that case superfluous. Thus, in the usual notation we have

~(x+y) =

+ = ~(xy)

showing two modes of writing the same fact. The apparent balance between the two sets of theorems exhibited so strikingly by Schröder, arises entirely from this double way of writing everything. But while the ordinary system is not so analytically fitted to its purpose as that here set forth, the character of superfluity here, as in many other cases in algebra, brings with it great facility in working.

387. The general formulae given above are not convenient in practice. We may dispense with them altogether, as well as with one of the indices of tokens used in them, by the use of the following rules. A proposition of the form

xy

is true if x = f or y = v. It is only false if y = f and x = v. A proposition written in the form

x ~⤙ y

is true if x = v and y = f, and is false if either x = f or y = v. Accordingly, to find whether a formula is necessarily true substitute f and v for the letters and see whether it can be supposed false by any such assignment of values. Take, for example, the formula

(xy) ⤙ {(yz) ⤙ (xz)}.

To make this false we must take

(xy) = v

{(yz) ⤙ (xz)} = f.

The last gives

(yz) = v, (xz) = f, x = v, z = f.

Substituting these values in

(xy) = v (yz) = v
we have (vy) = v (yf) = v,

which cannot be satisfied together.

388. As another example, required the conclusion from the following premisses: Anyone I might marry would be either beautiful or plain; anyone whom I might marry would be a woman; any beautiful woman would be an ineligible wife; any plain woman would be an ineligible wife. Let

m be anyone whom I might marry,

b, beautiful,

p, plain,

w, woman,

i, ineligible.

Then the premisses are

m ⤙ (bf) ⤙ p,

mw,

wbi,

wpi.

Let x be the conclusion. Then,

m ⤙ (bf) ⤙ p ⤙ (mw) ⤙ (wbi) ⤙ (wpi) ⤙ x

is necessarily true. Now if we suppose m = v, the proposition can only be made false by putting w = v and either b or p = v. In this case the proposition can only be made false by putting i = v. If, therefore, x can only be made f by putting m = v, i = f, that is if x = (mi) the proposition is necessarily true.

In this method, we introduce the two special tokens of second intention f and v, we retain two indices of tokens x and y, and we have a somewhat complex icon, with a special prescription for its use.

389. A better method may be found as follows. We have seen that

x ⤙ (yz)
may be conveniently written xyz;
while (xy) ⤙ z

ought to retain the parenthesis. Let us extend this rule, so as to be more general, and hold it necessary always to include the antecedent in parenthesis.

Thus, let us write (x) ⤙ y

instead of xy. If now, we merely change the external appearance of two signs; namely, if we use the vinculum instead of the parenthesis, and the sign + in place of ⤙, we shall have

xy written +y

xyz written ++z

(xy) ⤙ z written ~(+)+z, †1 etc.

We may further write for x ~⤙ y, ~(+y) implying that x+y †2 is an antecedent for whatever consequent may be taken, and the vinculum becomes identified with the sign of negation. We may also use the sign of multiplication as an abbreviation, putting

xy = ~(+) = ~(x).

390. This subjects addition and multiplication to all the rules of ordinary algebra, and also to the following:

y+x = y y(x+) = y
x+ = v x̄x = f
xy+z = (x+z)(y+z).

391. To any proposition we have a right to add any expression at pleasure; also to strike out any factor of any term. The expressions for different propositions separately known may be multiplied together. These are substantially Mr. Mitchell's rules of procedure. †1 Thus the premisses of Barbara are

+y and +z.

Multiplying these, we get (+y)(+z) = +yz. Dropping and y we reach the conclusion +z.

§3. First-Intentional Logic of Relatives

392. The algebra of Boole affords a language by which anything may be expressed which can be said without speaking of more than one individual at a time. It is true that it can assert that certain characters belong to a whole class, but only such characters as belong to each individual separately. The logic of relatives considers statements involving two and more individuals at once. Indices are here required. Taking, first, a degenerate form of relation, we may write xiyj to signify that x is true of the individual i while y is true of the individual j. If z be a relative character zij will signify that i is in that relation to j. In this way we can express relations of considerable complexity. Thus, if

1, 2, 3,

4, 5, 6,

7, 8, 9,

are points in a plane, and l123 signifies that 1, 2, and 3 lie on one line, a well-known proposition of geometry †1 may be written

l159l267l348l147l258l369l123l456l789.

In this notation is involved a sixth icon.

393. We now come to the distinction of some and all, a distinction which is precisely on a par with that between truth and falsehood; that is, it is descriptive.

All attempts to introduce this distinction into the Boolian algebra were more or less complete failures until Mr. Mitchell †2 showed how it was to be effected. His method really consists in making the whole expression of the proposition consist of two parts, a pure Boolian expression referring to an individual and a Quantifying part saying what individual this is. Thus, if k means 'he is a king,' and h, 'he is happy,' the Boolian

(+h)

means that the individual spoken of is either not a king or is happy. Now, applying the quantification, we may write

Any (+h)

to mean that this is true of any individual in the (limited) universe, or

Some (+h)

to mean that an individual exists who is either not a king or is happy. So

Some (kh)

means some king is happy, and

Any (kh)

means every individual is both a king and happy. The rules for the use of this notation are obvious. The two propositions

Any (x) Any (y)

are equivalent to

Any (xy).

From the two propositions

Any (x) Some (y)

we may infer

Some (xy). †P1

Mr. Mitchell has also a very interesting and instructive extension of his notation for some and all, to a two-dimensional universe, that is, to the logic of relatives. Here, in order to render the notation as iconical as possible we may use Σ for some, suggesting a sum, and Π for all, suggesting a product. Thus Σixi means that x is true of some one of the individuals denoted by i or

Σixi = xi+xj+xk+etc. †1

In the same way, Πixi means that x is true of all these individuals, or

Πixi = xixjxk, etc. †2

If x is a simple relation, ΠiΠjxij means that every i is in this relation to every j, ΣiΠjxij that some one i is in this relation to every j, ΠjΣixij that to every j some i or other is in this relation, ΣiΣjxij that some i is in this relation to some j. It is to be remarked that Σixi and Πixi are only similar to a sum and a product; they are not strictly of that nature, because the individuals of the universe may be innumerable.

394. At this point, the reader would perhaps not otherwise easily get so good a conception of the notation as by a little practice in translating from ordinary language into this system and back again. Let lij mean that i is a lover of j, and bij that i is a benefactor of j. Then

ΠiΣjlijbij

means that everything is at once a lover and a benefactor of something; and

ΠiΣjlijbji

that everything is a lover of a benefactor of itself.

ΣiΣkΠj(lij + bjk)

means that there are two persons, one of whom loves everything except benefactors of the other (whether he loves any of these or not is not stated). Let gi mean that i is a griffin, and ci that i is a chimera, then

ΣiΠj(gilij + j)

means that if there be any chimeras there is some griffin that loves them all; while

ΣiΠjgi(lij + j)

means that there is a griffin and he loves every chimera that exists (if any exist). On the other hand,

ΠjΣigi(lij + j)

means that griffins exist (one, at least), and that one or other of them loves each chimera that may exist; and

ΠjΣi(gilij + j)

means that each chimera (if there is any) is loved by some griffin or other.

395. Let us express: every part of the world is either sometimes visited with cholera, and at others with small-pox (without cholera), or never with yellow fever and the plague together.

Let

cij mean the place i has cholera at the time j.

sij mean the place i has small-pox at the time j.

yij mean the place i has yellow fever at the time j.

pij mean the place i has plague at the time j.

Then we write

ΠiΣjΣkΠl(cijiksik+il+il)

Let us express this: one or other of two theories must be admitted, first, that no man is at any time unselfish or free, and some men are always hypocritical, and at every time some men are friendly to men to whom they are at other times inimical, or second, at each moment all men are alike either angels or fiends. Let

uij mean the man i is unselfish at the time j,

fij mean the man i is free at the time j,

hij mean the man i is hypocritical at the time j,

aij mean the man i is an angel at the time j,

dij mean the man i is a fiend at the time j,

pijk mean the man i is friendly at the time j, to the man k,

eijk the man i is an enemy at the time j to the man k;

1jm the two objects j and m are identical.

Then the proposition is

ΠiΣhΠjΣkΣlΣmΠnΠpΠq(ūijijhhjpkjlekmljm + apn + dqn)

396. We have now to consider the procedure in working with this calculus. It is far from being true that the only problem of deduction is to draw a conclusion from given premisses. On the contrary, it is fully as important to have a method for ascertaining what premisses will yield a given conclusion. There are besides other problems of transformation, where a certain system of facts is given, and it is required to describe this in other terms of a definite kind. Such, for example, is the problem of the fifteen young ladies, †1 and others relating to synthemes. †2 I shall, however, content myself here with showing how, when a set of premisses are given, they can be united and certain letters eliminated. Of the various methods which might be pursued, I shall here give the one which seems to me the most useful on the whole.

First, the different premisses having been written with distinct indices (the same index not used in two propositions) are written together, and all the Π's and Σ's are to be brought to the left. This can evidently be done, for

Πixi . Πjxj = ΠiΠjxixj

Σixi . Πjxj = ΣiΠjxixj

Σixi . Σjxj = ΣiΣjxixj

Second, without deranging the order of the indices of any one premiss, the Π's and Σ's belonging to different premisses may be moved relatively to one another, and as far as possible the Σ's should be carried to the left of the Π's. We have

ΠiΠjxij = ΠjΠixij

ΣiΣjxij = ΣjΣixij

and also ΣiΠjxiyj = ΠjΣixiyj

But this formula does not hold when the i and j are not separated. We do have, however,

ΣiΠjxij ⤙ ΠiΣjxij. †1

It will, therefore, be well to begin by putting the Σ's to the left, as far as possible, because at a later stage of the work they can be carried to the right but not to the left. For example, if the operators of the two premisses are ΠiΣjΠk and ΣxΠyΣz, we can unite them in either of the two orders

ΣxΠyΣzΠiΣjΠk

ΣxΠiΣjΠyΣzΠk,

and shall usually obtain different conclusions accordingly. There will often be room for skill in choosing the most suitable arrangement.

Third, it is next sometimes desirable to manipulate the Boolian part of the expression, and the letters to be eliminated can, if desired, be eliminated now. For this purpose they are replaced by relations of second intention, such as "other than," etc. If, for example, we find anywhere in the expression

aijkāxyz,

this may evidently be replaceable by

(nix+njy+nkz)

where, as usual, n means not or other than. This third step of the process is frequently quite indispensable, and embraces a variety of processes; but in ordinary cases it may be altogether dispensed with. †1

Fourth, the next step, which will also not commonly be needed, consists in making the indices refer to the same collections of objects, so far as this is useful. If the quantifying part, or Quantifier, contains Σx and we wish to replace the x by a new index i, not already in the Quantifier, and such that every x is an i, we can do so at once by simply multiplying every letter of the Boolian having x as an index by xi. Thus, if we have "some woman is an angel" written in the form Σwaw we may replace this by Σi(aiwi). It will be more often useful to replace the index of a Π by a wider one; and this will be done by adding i to every letter having x as an index. Thus, if we have "all dogs are animals, and all animals are vertebrates" written thus

Πdαd Πava

where a and α alike mean animal, it will be found convenient to replace the last index by i, standing for any object, and to write the proposition

Πi(āi+vi)

Fifth, †2 the next step consists in multiplying the whole Boolian part, by the modification of itself produced by substituting for the index of any Π any other index standing to the left of it in the Quantifier. Thus, for

ΣiΠjlij,
we can write ΣiΠjlijlii.

Sixth, the next step consists in the re-manipulation of the Boolian part, consisting, first, in adding to any part any term we like; second, in dropping from any part any factor we like, and third, in observing that

x = f, x+ = v,
so that xx̄y+z = z, (x++y)z = z.

Seventh, Π's and Σ's in the Quantifier whose indices no longer appear in the Boolian are dropped.

The fifth step will, in practice, be combined with part of the sixth and seventh. Thus, from ΣiΠjlij we shall at once proceed to Σilii if we like.

397. The following examples will be sufficient.

From the premisses Σiaibi and Πj(j+cj), eliminate b. We first write

ΣiΠjaibi(j+cj).

The distributive process gives

ΣiΠjai(bij + bicj).

But we always have a right to drop a factor or insert an additive term. We thus get

ΣiΠjai(bij + cj).

By the third process, we can, if we like, insert nij for bij. In either case, we identify j with i and get the conclusion

Σiaici.

Given the premisses:

ΣhΠiΣjΠkhik+sjklji)

ΣuΣvΠxΠyuyx + yvbvx).

Required to eliminate s. The combined premiss is

ΣuΣvΣhΠiΣjΠxΠkΠyhik+sjklji)(εuyx+yvbvx).

Identify k with v and y with j, and we get

ΣuΣvΣhΠiΣjΠxhiv+sjvlji)(εujx+jvbvx).

The Boolian part then reduces, so that the conclusion is

ΣuΣvΣhΠiΣjΠxhivεujxhivbvxujxlji).

§4. Second-Intentional Logic †1

398. Let us now consider the logic of terms taken in collective senses. Our notation, so far as we have developed it, does not show us even how to express that two indices, i and j, denote one and the same thing. We may adopt a special token of second intention, say 1, to express identity, and may write 1ij. But this relation of identity has peculiar properties. The first is that if i and j are identical, whatever is true of i is true of j. This may be written

ΠiΠj{1̅ij+i+xj}.

The use of the general index of a token, x, here, shows that the formula is iconical. The other property is that if everything which is true of i is true of j, then i and j are identical. This is most naturally written as follows: Let the token, q, signify the relation of a quality, character, fact, or predicate to its subject. Then the property we desire to express is

ΠiΠjΣk(1ij+kiqkj).

And identity is defined thus

1ij = Πk(qkiqkj+kikj).

That is, to say that things are identical is to say that every predicate is true of both or false of both. It may seem circuitous to introduce the idea of a quality to express identity; but that impression will be modified by reflecting that qkiqjk merely means that i and j are both within the class or collection k. If we please, we can dispense with the token q, by using the index of a token and by referring to this in the Quantifier just as subjacent indices are referred to. That is to say, we may write

1ij = Πx(xixj+ij).

399. The properties of the token q must now be examined. These may all be summed up in this, that taking any individuals i1, i2, i3, etc., and any individuals, j1, j2, j3, etc., there is a collection, class, or predicate embracing all the i's and excluding all the j's except such as are identical with some one of the i's. This might be written

αΠiα)(ΠβΠjβkαΣi'αlqkiα(kjβ + qli'αqljβ + li'αljβ),

where the i's and the i' 's are the same lot of objects. This notation presents indices of indices. The ΠαΠiα shows that we are to take any collection whatever of i's, and then any i of that collection. We are then to do the same with the j's. We can then find a quality k such that the i taken has it, and also such that the j taken wants it unless we can find an i that is identical with the j taken. The necessity of some kind of notation of this description in treating of classes collectively appears from this consideration: that in such discourse we are neither speaking of a single individual (as in the non-relative logic) nor of a small number of individuals considered each for itself, but of a whole class, perhaps an infinity of individuals. This suggests a relative term with an indefinite series of indices as xijkl . . . . Such a relative will, however, in most, if not in all cases, be of a degenerate kind and is consequently expressible as above. But it seems preferable to attempt a partial decomposition of this definition. In the first place, any individual may be considered as a class. This is written

ΠiΣkΠj qki(kj + 1ij).

This is the ninth icon. †1 Next, given any class, there is another which includes all the former excludes and excludes all the former includes. That is,

ΠlΣkΠi(qliki + liqki).

This is the tenth icon. Next, given any two classes, there is a third which includes all that either includes and excludes all that both exclude. That is

ΠlΠmΣkΠi(qliqki + qmiqki + limiki).

This is the eleventh icon. Next, given any two classes, there is a class which includes the whole of the first and any one individual of the second which there may be not included in the first and nothing else. That is,

ΠlΠmΠiΣkΠj{qli+mi+qki(qkj+lj)}.

This is the twelfth icon.

400. To show the manner in which these formulæ are applied let us suppose we have given that everything is either true of i or false of j. We write

Πk(qki + kj). †2

The tenth icon gives

ΠlΣk(qliki + liqki)(qljkj + ljqkj).

Multiplication of these two formulæ give

ΠlΣk(qkili + qljkj),

or, dropping the terms in k

Πl(li + qlj).

Multiplying this with the original datum and identifying l with k, we have

Πk(qkiqkj+kikj).

No doubt, a much more direct method of procedure could be found.

401. Just as q signifies the relation of predicate to subject, so we need another token, which may be written r, to signify the conjoint relation of a simple relation, its relate and its correlate. That is, rjαi is to mean that i is in the relation α to j. Of course, there will be a series of properties of r similar to those of q. But it is singular that the uses of the two tokens are quite different. Namely, the chief use of r is to enable us to express that the number of one class is at least as great as that of another. This may be done in a variety of different ways. Thus, we may write that for every a there is a b, in the first place, thus:

ΣαΠiΣjΠh{āi+bjrjαi(jαh+āh+1ih)}.

But, by an icon analogous to the eleventh, we have

ΠαΠβΣγΠuΠv(ruαvruγv + ruβvruγv + uαvuβvuγv).

From this, by means of an icon analogous to the tenth, we get the general formula

ΠαΠβΣγΠuΠv{ruαvruβvruγv + uγv(uαv + uβv)}.

For ruβv substitute au and multiply by the formula the last but two. Then, identifying u with h and v with j, we have

ΣαΠiΣhΠh{āi + bjrjαi(jαh + 1ih)}

a somewhat simpler expression. However, the best way to express such a proposition is to make use of the letter c as a token of a one-to-one correspondence. That is to say, c will be defined by the three formulæ, †1

ΠαΠuΠvΠw(α + uαv + uαw + 1vw)

ΠαΠuΠvΠw(α + uαw + rvαw †2 + 1uv)

ΠαΣuΣvΣw(cα + ruαvruαwvw + ruαwrvαwuv). †3

Making use of this token, we may write the proposition we have been considering in the form

ΣαΠiΣj cα(āi + bjrjαi).

402. In an appendix to his memoir on the logic of relatives, †1 De Morgan enriched the science of logic with a new kind of inference, the syllogism of transposed quantity. De Morgan was one of the best logicians that ever lived and unquestionably the father of the logic of relatives. Owing, however, to the imperfection of his theory of relatives, the new form, as he enunciated it, was a down-right paralogism, one of the premisses being omitted. But this being supplied, the form furnishes a good test of the efficacy of a logical notation. The following is one of De Morgan's examples: †2

Some X is Y,

For every X there is something neither Y nor Z;

Hence, something is neither X nor Z.

The first premiss is simply

Σαxαyα.

The second may be written

ΣαΠiΣj cα(i + rjαijj).

From these two premisses, little can be inferred. To get the above conclusion it is necessary to add that the class of X's is a finite collection †3; were this not necessary the following reasoning would hold good (the limited universe consisting of numbers); for it precisely conforms to De Morgan's scheme.

Some odd number is prime;

Every odd number has its square, which is neither prime nor even;

Hence, some number is neither odd nor even. †P1

Now, to say that a lot of objects is finite, is the same as to say that if we pass through the class from one to another we shall necessarily come round to one of those individuals already passed; that is, if every one of the lot is in any one-to-one relation to one of the lot, then to every one of the lot some one is in this same relation. This is written thus:

ΠβΠuΣvΣsΠt{β + u + xvruβv + xs(t + tβs)}

Uniting this with the two premisses and the second clause of the definition of c, we have

ΣaΣαΠβΠuΣvΣsΠiΣjΠtΠγΠeΠfΠgxayacα (i + rjαijj)

{β + u + xvruβv + xs(t + tβs)} (γ + eγg + fγw + 1ef).

We now substitute α for β and for γ, a for u and for e, j for t and for f, v for g. The factor in i is to be repeated, putting first s and then v for i. The Boolian part thus reduces to

(s + rjαsjj)cαxayaraαvxv rjαvjj1aj + rjαsjjxsj (v + rjavjj)(aαv + jαv1aj,

which, by the omission of factors, becomes

yaj1aj + jj.

Thus we have the conclusion

Σjjj.

403. It is plain that by a more iconical and less logically analytical notation this procedure might be much abridged. How minutely analytical the present system is, appears when we reflect that every substitution of indices of which nine were used in obtaining the last conclusion is a distinct act of inference. The annulling of (yaj1aj) makes ten inferential steps between the premisses and conclusion of the syllogism of transposed quantity.

§5. Note †1

403A. Under the third step 396, an example was given which is really a general formula of elimination. Namely, we have

aijk etc. āxyz etc. ⤙ 1̅ix + 1̅jy + 1̅kz etc.

and conversely

1ix1jy1zk etc. ⤙ aijk etc. + āxyz etc.

The principle of contradiction and of excluded middle might be considered as mere special cases of these formulæ. Namely, the latter give

aiāi ⤙ 1̅ii

1 ⤙ ai + āi

But the definition of identity is

v ⤙ 1ii iif;

whence

aiāif vai + āi.

403B. Under the head of the third step belongs the frequently necessary development of the Boolian by means of distribution formulæ. (This proceeding, and indeed the whole of the third step, might more properly have been made to follow the fourth.) The fundamental formulæ of distribution are

x(y+z) = xy+xz

x+yz = (x+y)(x+z).

The general development formulæ thence resulting are

Fx = xF1 + F0
= (x + F0)( + F1)

The following formulæ, which find continual application, are deducible from the above:

a+b = a + āb

xy+ = (x+)(+y)

(a+b)ca+bc

(a+x)(b+y) ⤙ a+b+xy.

403C. The fifth step is composed of two kinds of operations, the involution of the whole expression, and the identification and discrimination of indices.

403D. It is plain that any algebraical expression of a proposition may be multiplied into itself any number of times without ceasing to be true; and it will be found that such involution is essentially necessary in all difficult modes of inference. Consider, for example, the last formula but one of those last given,

(a+b)ca+bc.

This is not itself a distribution formula, but only an association formula, and therefore the deduction of a distribution formula from it is not a matter of the very utmost simplicity. If, however, we square the antecedent we have

(a+b)c = (a+b)c(a+b)c = c(a+b)c,

and then by the application of the association formula twice over we get

c(a+b)c ⤙ (ca+b)cca+bc;

so that we have proved

(a+b)cac+bc,

which is the main distribution formula, with the aid of which all the others are readily obtained.

Other much more striking examples of the utility of involution will present themselves in the course of this paper.

403E. In multiplying a proposition by itself, we have the right to choose any point from the beginning to the end of the quantifier and to identify in two factors all the indices to the left of this point while diversifying all to the right of it. When there is but one index, this is plainly true (see the formulæ in first step); for

Πiai = (Πiai)(Πiai) = (Πiai)(Πjaj)
Σiai = (Σiai)(Σiai) = (Σiai)(Σjaj).
Now iai)(Πjaj) = (a1a2a3 etc.)(a1a2a3 etc.)
while ΠiΠjaiaj = a1a1 · a1a2 · a1a3 etc.
× a2a1 · a2a2 · a2a3 etc.
× a3a1 · a3a2 · a3a3 etc.
× etc.
and Πiaiai = a1a1 · a2a2 · a3a3 · etc.

But since aa = a, all these are equivalent.

Also

iai)(Σjaj) = (a1+a2+a3+etc.)(a1+a2+a3+etc.)

while

ΣiΣjaiaj = a1a1+a1a2+a1a3 + etc.
+ a2a1+a2a2+a2a3 + etc.
+ a3a1+a3a2+a3a3 + etc.
+ etc.
and Σiaiai = a1a1+a2a2+a3a3 + etc.

The first two expressions are equivalent by the distribution principle. The third is equivalent to Σiai because aa = a, and thus all three are equivalent to one another.

The theorem enunciated having thus been proved true of every proposition having a single index, it only remains to show that if it be true of every proposition having n indices, it is true for every proposition having n + 1 indices.

Now let Φ and * stand either for Σ and + or for Π and X respectively, so that

Φiui = u1 * u2 * u3 * etc.

This may represent any proposition in n + 1 indices of which i is the first. When i is fixed in a certain individual as in u1, u2, etc., u becomes a proposition in n indices. Let u2 represent any legitimate square of u. Then the square of the whole expression with identification of i in the two factors is

Φiui2 = u12 * u22 * u32 * etc.

and since by hypothesis

u1u12 u2u22 u3u32 etc.

it follows that

Φiui ⤙ Φiui2.

Thus it is shown that the rule holds for every proposition in n + 1 indices, if the first of them is identified in the two factors. But according to the rule, the first cannot be diversified unless all are diversified; and all may be diversified, by an obvious extension of the formulæ relating to propositions in a single index. The theorem is therefore proved.

If the proposition is raised to a higher power, the diversifications toward the right may be represented as branchings from a stem, thus,

inline image

After the involution has once been performed, further identification and diversification of indices may be effected by applying the following formulæ.

For identifications For diversifications
ΠiΠjlij ⤙ Πilii Πilii ⤙ ΠiΣjlij
ΣiΠjlij ⤙ Σilii Σilii ⤙ ΣiΣjlij

403F. Besides these formulæ, the following are occasionally useful

Πimi = ΠiΠjmimj Σimi = ΣiΣj(mi+mj).

The following example shows the utility of changing the indices, even without involution. From the premiss

ΠiΠj(milij + iij)

let it be required to eliminate m. The immediate dropping of this token would only yield an identical proposition. But by separating the Boolian into factors and then changing the indices, we have

ΠiΠj(milij + iij)
= ΠiΠj(mi+ ij)(lij + i)
= ΠiΠj(mi+ ijj(lij + i)
= ΠiΠj(mi+ ijk(lik + i)
= ΠiΠjΠk(mi+ ij)(lik + i)
= ΠiΠjΠk(milik + iij)

Now dropping m we have

ΠiΠjΠk(lik + ij)

i.e. whatever is in the relation l to anything is in that relation to everything.

403G. An artifice which I have not included among the regular steps of the inferential procedure but which is occasionally useful, consists in taking the latter part of the quantifier away from its position at the head of the proposition and putting it before the part which alone it concerns. The formulæ governing this operation are

Πjaibij = aiΠibij Σjaibij = aiΣjbij
Πj(ai+bij) = ai + Πjbij Σj(ai+bij) = ai + Σjbij.

This transformation will generally be used in connection with the formulæ

aa + b and a + xb = a + xb. †1

403H. The special peculiarity of ordinary algebra has given us the false notion that inverse operations are the general means of solving algebraical problems. But the study of general algebra shows that inverse operations lead to determinate results only in special cases — that what is called "the general case" is in truth a mere form of speciality — and that the truly general method of elimination is by performing a direct operation which will give a constant result whatever the value of the variable. Thus, in ordinary algebra, it happens to be the case that every quantity has a reciprocal so that

(1/x) · x = 1.

So in logical algebra, the only way of eliminating any token is by means of the properties of special terms of second intention, such as

aijāxy ⤙ 1̅ix + 1jy aiāi ⤙ 0.

In order to bring these formulæ to bear it may be necessary to multiply the premiss into itself, to manipulate the indices, and use processes of association and of development by distribution; but whatever cannot be eliminated by these means cannot be eliminated at all.

403I. The universes of marks †2 to which the tokens q, r and others like these refer are, in reference to the combinations of objects to which they are attached, unlimited universes. (Compare the Thomist doctrine of angelic natures.) That is to say, every lot of objects has some quality common and peculiar to the objects composing it; every lot of pairs has some relation subsisting between the first and second members of each one of those pairs and between no others. In other regards, the universes are not unlimited; of characters familiar to us there is quite a limited number; of colours definable by Newton's diagram not all can [co-?]exist. In like manner, the universes of quantity, position, etc., of mathematics are unlimited universes. Of the objects of such a universe everything is true, which can be true; every proposition is true from which the unlimited universe cannot be eliminated without yielding a true proposition.

As a general rule, every proposition in Σαqα and Σαrα is true; but there are many exceptions.

To illustrate the application of this principle, consider the ninth icon. †1 This is written

ΠaΣκΠbqκa(κb + 1ab)

Now κ can only be eliminated from this without involution in two ways; first, eliminating qκa and qκb independently, we have

ΠaΠb(1̅ab + 1ab),

an identical proposition; second, identifying b with a, we get

Πa1aa,

another identical proposition. We next proceed to square the proposition; and we have

ΠaΣκΠbΠcqκa(κb + 1ab)(κc + 1ac).

We must identify κ in the factors or we should reach no result; and this forces us to identify a. But it is plain that we cannot from this square eliminate in any new way, and therefore we could not from any higher power, and consequently the original proposition is proved true.

Suppose, however, we had written that proposition

ΣκΠaΠbqκa(κb + 1ab).

Squaring this we have

ΣκΠaΠbΠcΠdqκaqκc(κb + 1ab)(κd + 1cd).

Now identify d with a, c with b, and we have

ΠaΠb1ab,

which is absurd. †1

403J. By the same principle, we can if we please solve the example of 400 as follows. Given the premiss

Πκ(qκi + κj):

We multiply the square of this by an identical proposition, thus,

ΠλΠκΠμ(qλi + λi)(qλj + λj)(qμi + μj)(qκi + κj).

Identifying μ with λ, we get

ΠλΠκ(qλi + λi)(qλi + λj)(qλi + λj)(qκi + κj).

We now introduce κi and qκj into the identical proposition wherever we please so long as they cannot be eliminated without making the proposition otherwise than identical; and this condition will be fulfilled so long as we do not introduce qκi or κj. Only Πκ must be changed to Σκ. We get then

ΠλΣκ(qλiκi + λi)(qλj + qκjλi)(qλi + λj)(qκi + κj).

This gives successively

ΠλΣq(qλiκiqλjκj + λiqκjλjqκi)

and Πλ(qλiqλj + λiλj).

But the same conclusion can be reached much more easily by identifying qκ with 1j, which we have a right to do on account of the universal quantifier. Thus from

ΠiΣjΠκ(qκi + κj)

we get

ΠiΣj(1ji + 1̅jj) = ΠiΣj1ji.

Hence the proposition

Πλ(qλiqλj + λiλj)

holds, because it becomes identical when i is substituted for j.

403K. Let us now consider some examples of a somewhat more difficult kind. Given the proposition

ΣκΠaΣbΠc{a + rκabyb(κcb + 1ac)},

that is, there is a relation κ such that whatever object be taken, either this is not an x or it stands in the relation k to some object which is a y and to which nothing except that x stands in that relation. Required to find all the propositions deducible from this by elimination of κ. The simple omission of factors gives

ΠaΣb(a + yb),

that is, there either is no x or there is a y. We get nothing further by identifying c with a. There is therefore no further conclusion without involution. Squaring, with identification of κ (without which we should plainly reach nothing new) we have

ΣκΠaΣbΠcΠdΣeΠf{a + rκabyb(κcb + 1ac)} {d + rκdeye(κfe + 1df)}.

The application to the Boolian of the common formula

(a + x)(b + y) ⤙ a + b + xy

gives

a + d + rκabrκdeybye(κcb + 1ac)(κfe + 1df).

We now identify f with a; when the first and last factors of the last term become

rκab(κae + 1ad).

But we have

rκabκae ⤙ 1̅be.

Thus, we reach the conclusion

ΠaΣbΠdΣe(a + d + 1ad +ybyebe);

that is, there are either not two x's different from one another or there are two y's different from one another. A little examination will show that no other conclusion could be reached by elimination from the square.

Cubing the original proposition, we reach in a similar way

ΣκΠaΣbΠcΠdΣeΠfΠgΣhΠia + d + g + ybyeyhrκabrκderκgh (κcb + 1ac)(κfe + 1df)(κih + 1gi)

Identifying f with a, i with d, g with c, the last term of the Boolian becomes

ybyeyhrκabrκderκch (κcb + 1ac)(κae + 1ad)(κdh + 1cd)

whence

ybyeyh (1̅bh + 1ac)(1̅be + 1ad)(1eh + 1cd),

whence again

ybyeyhbhbe1ch †1 + 1ac + 1ad + 1cd,

so that we have the conclusion

ΠaΣbΠcΠdΣeΣh

a + d + c + 1ac + 1ad + 1cd + ybyeyhbebheh,

that is, there either are not three x's all different from one another or there are three y's all different from one another.

It is plain that by raising to the n! power we should get any proposition of this form, and that no others could be obtained.

403L. Let us now seek all the propositions deducible, by elimination of κ, from

ΠκΣaΠbΣc ua(b + κab + rκcbac).

We have at once

Σaua.

The following proposition is universally true:

ΠαΠβΣγΠiΠj{γij(αij + βij) + rγijrαijrβij}.

To prove this, note first that if we eliminate γ at once, we have an identical proposition. If we raise the whole to a power, there will be no additional mode of elimination, unless i, j and γ be identified in different factors; but then all the indices must be alike and nothing will be changed. As a special case of this formula we put ui for rβij. This we can do, because

ΣβΠiΠj{uirβij + ūiβij}

which is true by the same reasoning as that just used. The product of these two formulæ gives

ΠαΣγΠiΠj{γij(αij + ūi) + rγijrαijui}.

Multiplying this twice into the first proposition, identifying κ with γ, i with a, and j with b, in one factor, and κ with γ, i with c, j with b in the other, we have

ΠaΣγΣaΠbΣc ua(b + γab + rγcbac) {rγab(αab+ūa) + rγabrαabua} {γcb(αcb + ūc) + rγcbrαcbuc}.

This gives

ΠαΣaΠbΣc ua(b + αab + rαcbucac).

We have universally

ΣδΠmΠn(rδmn1mn + δmnmn). †1

Multiplying this twice into the last proposition, identifying α with δ, and n with b, in both factors, m with a in one and with c in the other, we have

ΣaΠbΣc ua(b + 1̅ab + 1cbucac);

or since

1cbac ⤙ 1̅ab

ΣaΠb ua(b+ 1̅ab);

and identifying b with a

Σauaa.

Again, we have universally

ΣδΠmΠnrδmn. †2

Multiplying this twice into the proposition

ΠαΣaΠbΣc ua(b + αab + rαcbucac)

with the same identifications as before, we get

ΣaΠbΣc ua(b + ucac),

that is, there is a u, and if there be a w there is a second u. This last proposition shows, by the formula

ΣaΣc xaxcac ⤙ ΠaΣc(a + xcac)

that the original proposition may be written in the form

ΠαΣaΠbΣc ua {b + ucac(αab + rαcb)}

403M. Sixth step. The step numbered fifth 396 may more conveniently be separated into two. The first of these somewhat resembles the last; it is a sort of development or setting forth in detail the premiss, but instead of being founded upon a distribution formula it consists in raising the whole premiss to a power or multiplying it into itself. At each such multiplication any of the indices may be changed to new ones, their order in the quantifier being determined by the rules of the second step. To prove that this can be done we begin by confining our attention to the first index of the quantifier. The proposition is then either of the form Πiai or Σiai. We have obviously

Πiai = a1a2a3 etc. = ΠiΠjaiaj
Σiai = a1 + a2 + a3 + etc.
= (a1+a2+a3+etc.)(a1+a2+a3+etc.)
= ΣiΣjaiaj.

Next consider the first two indices. The proposition is of one of the four forms

ΠiΠjaij ΣiΠjaij ΠiΣjaij ΣiΣjaij

For the first and last of these, we have only to apply the formulæ just obtained, with the first two [of step two in 396]. That is,

ΠiΠjaij = ΠjΠiaij = ΠiΠkΠjaijakj
= ΠjΠiΠkaijakj = ΠjΠlΠiΠkaijakl.
ΣiΣjaij = ΣjΣiaij = ΣiΣkΣjaijakj
= ΣiΣjΣkΣlaijakl.

For the other two forms the proceeding is not more difficult. By the formulæ for a single index

Σijaij) = ΣiΣkjaij)(Πjakj) = ΣiΣkΠjaijakj

Πijaij) = ΠiΠkjaij)(Σjakj)



Paper 14: The Critic of Arguments †1

§1. Exact Thinking

404. "Critic" is a word used by Locke in English, by Kant in German, and by Plato in Greek, to signify the art of judging, being formed like "logic." I should shrink from heading my papers Logic, because logic, as it is set forth in the treatises, is an art far worse than useless, making a man captious about trifles and neglectful of weightier matters, condemning every inference really valuable and admitting only such as are really childish.

It is naughty to do what mamma forbids;

Now, mamma forbids me to cut off my hair:

Therefore, it would be naughty for me to cut off my hair.

This is the type of reasoning to which the treatises profess to reduce all the reasonings which they approve. Reasoning from authority does, indeed, come to that, and in a broad sense of the word authority, such reasoning only. This reminds us that the logic of the treatises is, in the main, a heritage from the ages of faith and obedience, when the highest philosophy was conceived to lie in making everything depend upon authority. Though few men and none of the less sophisticated minds of the other sex ever, nowadays, plunge into the darkling flood of the medieval commentaries, and fewer still dive deep enough to touch bottom, everybody has received the impression they are full of syllogistic reasoning; and this impression is correct. The syllogistic logic truly reflects the sort of reasoning in which the men of the middle ages sincerely put their trust; and yet it is not true that even scholastic theology was sufficiently prostrate before its authorities to have possibly been, in the main, a product of ordinary syllogistic thinking. Nothing can be imagined more strongly marked in its distinctive character than the method of discussion of the old doctors. Their one recipe for any case of difficulty was a distinction. That drawn, they would proceed to show that the difficulties were in force against every member of it but one. Therein all their labor of thinking lies, and thence comes all that makes their philosophy what it is. Without pretending, then, to pronounce the last word on the character of their thought, we may, at least, say it was not, in their sense, syllogistic; since in place of syllogisms it is rather characterised by the use of such forms as the following:

Everything is either P or M,

S is not M;

S is P.

This is commonly called disjunctive reasoning; but, for reasons which it would be too long to explain in full, I prefer to term it dilemmatic reasoning. Such modes of inference are, essentially, of the same character as the dilemma. Indeed, the regular stock example of the dilemma (for the logicians, in their gregariousness, follow their leader even down to the examples), though we find it set down in the second-century commonplace-book of Aulus Gellius, has quite the ring of a scholastic disquisition. The question, in this example, is, ought one to take a wife? In answering it, we first distinguish in regard to wives (and I seem to hear the Doctor subtillissimus saying: primo distinguendum est de hoc nomine uxor). A wife may mean a plain or a pretty wife. Now, a plain wife does not satisfy her husband; so one ought not to take a plain wife. But a pretty wife is a perpetual source of jealousy; so one ought still less to take a pretty wife. In sum, one ought to take no wife, at all. It may seem strange that the dilemma is not mentioned in a single medieval logic. It first appears in the De Dialectica of Rudolph Agricola. †1 But it should surprise nobody that the most characteristic form of demonstrative reasoning of those ages is left unnoticed in their logical treatises. The best of such works, at all epochs, though they reflect in some measure contemporary modes of thought, have always been considerably behind their times. For the methods of thinking that are living activities in men are not objects of reflective consciousness. They baffle the student, because they are a part of himself.

"Of thine eye I am eye-beam,"

says Emerson's sphynx. The methods of thinking men consciously admire are different from, and often, in some respects, inferior to those they actually employ. Besides, it is apparent enough, even to one who only knows the works of the modern logicians, that their predecessors can have been little given to seeing out of their own eyes, since, had they been so, their sequacious successors would have been religiously bound to follow suit.

405. One has to confess that writers of logic-books have been, themselves, with rare exceptions, but shambling reasoners. How wilt thou say to thy brother, Let me pull out the mote out of thine eye; and behold, a beam is in thine own eye? I fear it has to be said of philosophers at large, both small and great, that their reasoning is so loose and fallacious, that the like in mathematics, in political economy, or in physical science, would be received in derision or simple scorn. When, in my teens, I was first reading the masterpieces of Kant, Hobbes, and other great thinkers, my father, who was a mathematician, and who, if not an analyst of thought, at least never failed to draw the correct conclusion from given premisses, unless by a mere slip, would induce me to repeat to him the demonstrations of the philosophers, and in a very few words would usually rip them up and show them empty. In that way, the bad habits of thinking that would otherwise have been indelibly impressed upon me by those mighty powers, were, I hope, in some measure, overcome. Certainly, I believe the best thing for a fledgling philosopher is a close companionship with a stalwart practical reasoner.

406. How often do we hear it said that the study of philosophy requires hard thinking! But I am rather inclined to think a man will never begin to reason well about such subjects, till he has conquered the natural impulse to making spasmodic efforts of mind. In mathematics, the complexity of the problems renders it often a little difficult to hold all the different elements of our mental diagrams in their right places. In a certain sense, therefore, hard thinking is occasionally requisite in that discipline. But metaphysical philosophy does not present any such complications, and has no work that hard thinking can do. What is needed above all, for metaphysics, is thorough and mature thinking; and the particular requisite to success in the critic of arguments is exact and diagrammatic thinking.

407. †1 To illustrate my meaning, and at the same time to justify myself, in some degree, for conceding all I have to the prejudice of logicians, I will devote the residue of the space which I can venture to occupy today, to the examination of a statement which has often been made by logicians, and often dissented from, but which I have never seen treated otherwise than as a position quite possible for a reputable logician. I mean the statement that the principle of identity is the necessary and sufficient condition of the validity of all affirmative syllogisms, and that the principles of contradiction and excluded middle, constitute the additional necessary and sufficient conditions for the validity of negative syllogisms. The principle of identity, expressed by the formula "A is A," states that the relation of subject to predicate is a relation which every term bears to itself. The principle of contradiction, expressed by the formula "A is not not A," might be understood in three different senses; first, that any term is in the relation of negation to whatever term is in that relation to it, which is as much as to say that the relation of negation is its own converse; second, that no term is in the relation of negation to itself; third, that every term is in the relation of negation to everything but itself. But the first meaning is the best, since from it the other two readily follow as corollaries. The principle of excluded middle, expressed by the formula "Not not A is A," may also be understood in three senses; first, that every term, A, is predicable of anything that is in the relation of negation to a term which is in the same relation to it, A; second, that the objects of which any term, A, is predicable together with those of which the negative of A is predicable together make up all the objects possible; third, that every term, A, is predicable of whatever is in the relation of negation to everything but A. But, as before, the first meaning is to be preferred, since from it the others are immediately deducible.

408. There is but one mood of universal affirmative syllogism. It is called Barbara, and runs thus:

Any M is P,

Any S is M;

∴ Any S is P.

Now the question is, what one of the properties of the relation of subject to predicate is it, with the destruction of which alone this form of inference ceases invariably to yield a true conclusion from true premisses? To find that out the obvious way is to destroy all the properties of the relation in question, so as to make it an entirely different relation, and then note what condition this relation must satisfy in order to make the inference valid. Putting loves in place of is, we get:

M loves P,

S loves M;

S loves P.

That this should be universally true, it is necessary that every lover should love whatever his beloved loves. A relation of which the like is true is called a transitive relation. Accordingly, the condition of the validity of Barbara is that the relation expressed by the copula should be a transitive relation. This statement was first accurately made by De Morgan †1; but it is in substantial agreement with the doctrine of Aristotle. The analogue of the principle of identity, when loves is the copula of the proposition, is that everybody loves himself. This would plainly not suffice of itself to make the inferential form valid; nor would its being false prevent that form from being valid, provided loving were a transitive relation. Thus, by a little exact thinking, the principle of identity is clearly seen to be neither a sufficient nor a necessary condition for the truth of Barbara.

409. Let us now examine the negative syllogisms. The simplest of these is Celarent, which runs as follows:

Any M is not P,

Any S is M;

∴ Any S is not P

Let us substitute injures for is not. Then the form becomes

Every M injures P,

Every S is M;

∴ Every S injures P.

This is a good inference, still, no matter what sort of relation injuring is. Consequently, this syllogism is dependent upon no property of negation, except that it expresses a relation. Let us, in the last form, substitute loves for is. Then, we get

M injures P,

S loves M;

S injures P.

In order that this should hold good irrespective of the nature of the relation of injuring, it is necessary that nobody should love anybody but himself. A relation of that sort is called a sibi-relation or concurrency. †1 The necessary and sufficient condition of the validity of Celarent is, then, that the copula should express a sibi-relation. This is not what the principle of identity expresses. Of course, every sibi-relation is transitive.

410. The next simplest of the universal negative syllogisms is Camestres, which runs thus:

Any M is P,

Any S is not P;

∴ Any S is not M.

Substitute injures for is not, and we get,

Every M is a P,

Every S injures every P;

∴ Every S injures every M.

This obviously holds because the injuring is to every one of the class injured. It would not do to reason,

Every M is a P,

Every S injures a P;

∴ Every S injures an M.

We see, then, that the principal reason of the validity of Camestres is that by not, we mean not any, and not not some. In logical lingo, this is expressed by saying that negative predicates are distributed. But the condition that the copula expresses a sibi-relation is also involved.

411. The remaining universal negative syllogisms of the old enumeration, Celantes and Cesare, depend upon one principle. They are:

Celantes Cesare
Any M is not P, Any M is not P,
Any S is M; Any S is P;
∴ Any P is not S. ∴ Any S is not M.

Substituting fights for is not, we get

Every M fights every P,

Every S is M;

∴ Every P fights every S.

Every M fights every P,

Every S is P;

∴ Every S fights every M.

What is requisite to the validity of these inferences is plainly that the relation expressed by fights should be its own converse, or that everything should fight whatever fights it. This is the analogue of the principle of contradiction.

412. We see, then, that the principles of universal syllogism of the ordinary sort are that the copula expresses a sibi-relation, not that it expresses an agreement, which is what the principle of identity states, and that the negative is its own converse, which is the law of contradiction.

413. The authors who say that the principle of identity governs affirmative syllogism give no proof of what they allege. We are expected to see it by "hard thinking." I fancy I can explain what this process of "hard thinking" is. By a spasm induced by self-hypnotisation you throw yourself into a state of mental vacancy. In this state the formula "A is A" loses its definite signification and seems quite empty. Being empty it is regarded as wonderfully lofty and precious. Fired into enthusiasm by the contemplation of it, the subject, with one wild mental leap, throws himself into the belief that it must rule all human reason. Consequently, it is the principle of syllogism. If this is, as I suspect, what hard thinking means, it is of no use in philosophy.

414. As for the principle of excluded middle, the only syllogistic forms it governs are the dilemmatic ones.

Any not P is M,

Any S is not M;

∴ Any S is P.

Putting admiring for not, we have:

Everything admiring every P is an M,

Every S admires every M;

∴ Every S is a P.

To make this good, it must be that the only person who admires everybody that admires a given person is that person. This is the analogue of "everything not not A is A," which is the principle of excluded middle.

§2. The Reader is Introduced to Relatives †1

415. There is a melancholy book entitled Astronomy Without Mathematics. The author, an F. R. A. S., presumably knew something of astronomy; therefore, I pity him. I think I hear his groans and maledictions, as he wrote the book, over the initial lie to which he had committed himself, that it is possible to convey any idea of the science of astronomy without making use of mathematics. He could tell roughly how the planets go round the sun, and make his readers think they knew what the error of the ancient system was (namely, that all went round the earth — really, no error), and could set down surprising figures about the stars (beaten, however, by Buddhistic numbers both in magnitude and in intellectual value). A book so made might well have been called "The Story of the Heavens" (in anticipation of Dr. Ball's splendid volume, which, promising little, performs much), but it was not the "astronomy" stipulated for in the title page. When, in a neighbor's house yesterday, my eye lit upon that book, I shuddered. For I too have engaged myself by the title of these papers to produce something of solid value to my readers; but, thank God, I have not agreed to do it without the use of mathematics. I came home and pondered; and have decided that, in order to fulfill legitimate expectations, I must begin with a few chapters upon certain dry and somewhat technical matters that underlie the more interesting questions concerning reasoning. Do not fear a repetition of matter to be found in common textbooks. I shall suppose the reader to be acquainted with what is contained in Dr. Watts's Logick, a book very cheap and easily procured, and far superior to the treatises now used in colleges, being the production of a man distinguished for good sense. I mean to bring out a reprint of it, with extensive annotations, whenever I can find an eligible publisher. †1 Though a life-long student of reasonings, I know no way of giving the reader the benefit of what I ought to have learned, without asking him to go through with some irksome preliminary thinking about relations.
For this subject, although always recognised as an integral part of logic, has been left untouched on account of its intricacy. It is as though a geographer, finding the whole United States, its topography, its population, its industries, etc., too vast for convenient treatment, were to content himself with a description of Nantucket. This comparison hardly, if at all, exaggerates the inadequacy of a theory of reasoning that takes no account of relative terms.

416. A relation is a fact about a number of things. Thus the fact that a locomotive blows off steam constitutes a relation, or more accurately a relationship (the Century Dictionary, under relation, 3, gives the terminology. †2 See also relativity, etc.) between the locomotive and the steam. In reality, every fact is a relation. Thus, that an object is blue consists of the peculiar regular action of that object on human eyes. This is what should be understood by the "relativity of knowledge."

417. Not only is every fact really a relation, but your thought of the fact implicitly represents it as such. Thus, when you think "this is blue," the demonstrative "this" shows you are thinking of something just brought up to your notice; while the adjective shows that you recognise a familiar idea as applicable to it. Thus, your thought, when explicated, develops into the thought of a fact concerning this thing and concerning the character of blueness. Still, it must be admitted that, antecedently to the unwrapping of your thought, you were not actually thinking of blueness as a distinct object, and therefore were not thinking of the relation as a relation. †P1 There is an aspect of every relation under which it does not appear as a relation. Thus, the blowing off of steam by a locomotive may be regarded as merely an action of the locomotive, the steam not being conceived to be a thing distinct from the engine. This aspect we enphrase in saying, "the engine blows."

418. Thus, the question whether a fact is to be regarded as referring to a single thing or to more is a question of the form of proposition under which it suits our purpose to state the fact. Consider any argument concerning the validity of which a person might conceivably entertain for a moment some doubt. For instance, let the premiss be that from either of two provinces of a certain kingdom it is possible to proceed to any province by floating down the only river the kingdom contains, combined with a land-journey within the boundaries of one province; and let the conclusion be that the river, after touching every province in the kingdom, must again meet the one which it first left. Now, in order to show that this inference is (or that it is not) absolutely necessary, it is requisite to have something analogous to a diagram with different series of parts, the parts of each series being evidently related as those provinces are said to be, while in the different series something corresponding to the course of the river has all the essential variations possible; and this diagram must be so contrived that it is easy to examine it and find out whether the course of the river is in truth in every case such as is here proposed to be inferred. Such a diagram has got to be either auditory or visual, the parts being separated in the one case in time, in the other in space. But in order completely to exhibit the analogue of the conditions of the argument under examination, it will be necessary to use signs or symbols repeated in different places and in different juxtapositions, these signs being subject to certain "rules," that is, certain general relations associated with them by the mind. Such a method of forming a diagram is called algebra. All speech is but such an algebra, the repeated signs being the words, which have relations by virtue of the meanings associated with them. What is commonly called logical algebra differs from other formal logic only in using the same formal method with greater freedom. I may mention that unpublished studies †1 have shown me that a far more powerful method of diagrammatisation than algebra is possible, being an extension at once of algebra and of Clifford's method of graphs; but I am not in a situation to draw up a statement of my researches.

419. Diagrams and diagrammatoidal figures are intended to be applied to the better understanding of states of things, whether experienced, or read of, or imagined. Such a figure cannot, however, show what it is to which it is intended to be applied; nor can any other diagram avail for that purpose. The where and the when of the particular experience, or the occasion or other identifying circumstance of the particular fiction to which the diagram is to be applied, are things not capable of being diagrammatically exhibited. Describe and describe and describe, and you never can describe a date, a position, or any homaloidal quantity. You may object that a map is a diagram showing localities; undoubtedly, but not until the law of the projection is understood, nor even then unless at least two points on the map are somehow previously identified with points in nature. Now, how is any diagram ever to perform that identification? If a diagram cannot do it, algebra cannot: for algebra is but a sort of diagram; and if algebra cannot do it, language cannot: for language is but a kind of algebra. It would, certainly, in one sense be extravagant to say that we can never tell what we are talking about; yet, in another sense, it is quite true. The meanings of words ordinarily depend upon our tendencies to weld together qualities and our aptitudes to see resemblances, or, to use the received phrase, upon associations by similarity; while experience is bound together, and only recognisable, by forces acting upon us, or, to use an even worse chosen technical term, by means of associations by contiguity. Two men meet on a country road. One says to the other, "that house is on fire." "What house?" "Why, the house about a mile to my right." Let this speech be taken down and shown to anybody in the neighboring village, and it will appear that the language by itself does not fix the house. But the person addressed sees where the speaker is standing, recognises his right hand side (a word having a most singular mode of signification) estimates a mile (a length having no geometrical properties different from other lengths), and looking there, sees a house. It is not the language alone, with its mere associations of similarity, but the language taken in connection with the auditor's own experiential associations of contiguity, which determines for him what house is meant. It is requisite then, in order to show what we are talking or writing about, to put the hearer's or reader's mind into real, active connection with the concatenation of experience or of fiction with which we are dealing, and, further, to draw his attention to, and identify, a certain number of particular points in such concatenation. If there be a reader who cannot understand my writings, let me tell him that no straining of his mind will help him: his whole difficulty is that he has no personal experience of the world of problems of which I am talking, and he might as well close the book until such experience comes. That the diagrammatisation is one thing and the application of the diagram quite another, is recognised obscurely in the structure of such languages as I am acquainted with, which distinguishes the subjects and predicates of propositions. The subjects are the indications of the things spoken of, the predicates, words that assert, question, or command whatever is intended. †1 Only, the shallowness of syntax is manifest in its failing to recognise the impotence of mere words, and especially of common nouns, to fulfil the function of a grammatical subject. Words like this, that, lo, hallo, hi there, have a direct, forceful action upon the nervous system, and compel the hearer to look about him; and so they, more than ordinary words, contribute towards indicating what the speech is about. But this is a point that grammar and the grammarians (who, if they are faithfully to mirror the minds of the language-makers, can hardly be scientific analysts) are so far from seeing as to call demonstratives, such as that and this, pronouns — a literally preposterous designation, for nouns may more truly be called pro-demonstratives. †1

420. If upon a diagram we mark two or more points to be identified at some future time with objects in nature, †P1 so as to give the diagram at that future time its meaning; or if in any written statement we put dashes in place of two or more demonstratives or pro-demonstratives, the professedly incomplete representation resulting may be termed a relative rhema. It differs from a relative term only in retaining the "copula," or signal of assertion. If only one demonstrative or prodemonstrative is erased, the result is a non-relative rhema. For example, "— buys — from — for the price —," is a relative rhema; it differs in a merely secondary way from

"— is bought by — from — for —,"
from "— sells — to — for —,"
and from "— is paid by — to — for —."
On the other hand, "— is mortal" is a non-relative rhema.

421. A rhema is somewhat closely analogous to a chemical atom or radicle with unsaturated bonds. †2 A non-relative rhema is like a univalent radicle; it has but one unsaturated bond. A relative rhema is like a multivalent radicle. The blanks of a rhema can only be filled by terms, or, what is the same thing, by "something which" (or the like) followed by a rhema; or, two can be filled together by means of "itself" or the like. So, in chemistry, unsaturated bonds can only be saturated by joining two of them, which will usually, though not necessarily, belong to different radicles. If two univalent radicles are united, the result is a saturated compound. So, two non-relative rhemas being joined give a complete proposition. Thus, to join "— is mortal" and "— is a man," we have "X is mortal and X is a man," or some man is mortal. So likewise, a saturated compound may result from joining two bonds of a bivalent radicle; †P2 and, in the same way, the two blanks of a dual rhema may be joined to make a complete proposition. Thus, "— loves —," "X loves X," or something loves itself. A univalent radicle united to a bivalent radicle gives a univalent radicle (as H-O-); and, in like manner, a non-relative rhema, joined to a dual rhema, gives a non-relative rhema. Thus, "— is mortal" joined to "— loves —" gives "— loves something that is mortal," which is a non-relative rhema, since it has only one blank. Two, or any number of bivalent radicles united, give a bivalent radicle (as-O-O-S-O-O-), and so two or more dual rhemata give a dual rhema; as "— loves somebody that loves somebody that serves somebody that loves —." Non-relative and dual rhemata only produce rhemata of the same kind, so long as the junctions are by twos; but junctions of triple rhemata (or junctions of dual rhemata by threes), will produce all higher orders. Thus, "— gives — to —" and "— takes — from —," give "— gives — to somebody who takes — from —," a quadruple rhema. This joined to another quadruple rhema, as "— sells — to — for —," gives the sextuple rhema "— gives — to somebody who takes — from somebody who sells — to — for —." Accordingly, all rhemata higher than the dual may be considered as belonging to one and the same order; and we may say that all rhemata are either singular, dual, or plural. †1

422. Such, at least, is the doctrine I have been teaching for twenty-five years, and which, if deeply pondered, will be found to enwrap an entire philosophy. †2 Kant taught that our fundamental conceptions are merely the ineluctable ideas of a system of logical forms; nor is any occult transcendentalism requisite to show that this is so, and must be so. Nature only appears intelligible so far as it appears rational, that is, so far as its processes are seen to be like processes of thought. I must take this for granted, for I have no space here to argue it. It follows that if we find three distinct and irreducible forms of rhemata, the ideas of these should be the three elementary conceptions of metaphysics. That there are three elementary forms of categories is the conclusion of Kant, to which Hegel subscribes; and Kant seeks to establish this from the analysis of formal logic. Unfortunately, his study of that subject was so excessively superficial that his argument is destitute of the slightest value. Nevertheless, his conclusion is correct; for the three elements permeate not only the truths of logic, but even to a great extent the very errors of the profounder logicians. I shall return to them next week. †1 I will only mention here that the ideas which belong to the three forms of rhemata are firstness, secondness, thirdness; firstness, or spontaneity; secondness, or dependence; thirdness, or mediation.

423. But Mr. A. B. Kempe, in his important memoir on the "Theory of Mathematical Forms," †P1 presents an analysis which amounts to a formidable objection to my views. He makes diagrams of spots connected by lines; and it is easy to prove that every possible system of relationship can be so represented, although he does not perceive the evidence of this. But he shows (§68) that every such form can be represented by spots indefinitely varied, some of them being connected by lines, all of the same kind. He thus represents every possible relationship by a diagram consisting of only two different kinds of elements, namely, spots and lines between pairs of spots. Having examined this analysis attentively, I am of opinion that it is of extraordinary value. It causes me somewhat to modify my position, but not to surrender it. For, in the first place, it is to be remarked that Mr. Kempe's conception depends upon considering the diagram purely in its self-contained relations, the idea of its representing anything being altogether left out of view; while my doctrine depends upon considering how the diagram is to be connected with nature. It is not surprising that the idea of thirdness, or mediation, should be scarcely discernible when the representative character of the diagram is left out of account. In the second place, while it is not in the least necessary that the spots should be of different kinds, so long as each is distinguishable †P2 from the others, yet it is necessary that the connections between the spots should be of two different kinds, which, in Mr. Kempe's diagrams, appear as lines and as the absence of lines. Thus, Mr. Kempe has, and must have, three kinds of elements in his diagrams, namely, one kind of spots, and two kinds of connections of spots. In the third place, the spots, or units, as he calls them, involve the idea of firstness; the two-ended lines, that of secondness; the attachment of lines to spots, that of mediation.

424. My position has been modified by the study of Mr. Kempe's analysis. For, having a perfect algebra for dual relations, by which, for instance, I could express that "A is at once lover of B and servant of C," I declared that this was inadequate for the expression of plural relations; since to say that A gives B to C is to say more than that A gives something to C, and gives to somebody B, which is given to C by somebody. But Mr. Kempe (§330) virtually shows that my algebra is perfectly adequate to expressing that A gives B to C; since I can express each of the following relations:

In a certain act, D, something is given by A;

In the act, D, something is given to C;

In the act, D, to somebody is given B.

This is accomplished by adding to the universe of concrete things the abstraction "this action." But I remark that the diagram fails to afford any formal representation of the manner in which this abstract idea is derived from the concrete ideas. Yet it is precisely in such processes that the difficulty of all difficult reasoning lies. We have an illustration of this in the circumstance that I was led into an error about the capability of my own algebra for want of just the idea that process would have supplied. The process consists, psychologically, in catching one of the transient elements of thought upon the wing and converting it into one of the resting places of the mind. The difference between setting down spots in a diagram to represent recognised objects, and making new spots for the creations of logical thought, is huge. To include this last as one of the regular operations of logical algebra is to make an intrinsic transmutation of that algebra. What that mutation was I had already shown before Mr. Kempe's memoir appeared.



Paper 15: The Regenerated Logic †1

425. The appearance of Schröder's Exact Logic †P1 has afforded much gratification to all those homely thinkers who deem the common practice of designating propositions as "unquestionable," "undoubtedly true," "beyond dispute," etc., which are known to the writer who so designates them to be doubted, or perhaps even to be disputed, by persons who with good mental capacities have spent ten or more years of earnest endeavor in fitting themselves to judge of matters such as those to which the propositions in question relate, to be no less heinous an act than a trifling with veracity, and who opine that questions of logic ought not to be decided upon philosophical principles, but on the contrary, that questions of philosophy ought to be decided upon logical principles, these having been themselves settled upon principles derived from the only science in which there has never been a prolonged dispute relating to the proper objects of that science. Among those homely thinkers the writer of this review is content to be classed.

Why should we be so much gratified by the appearance of a single book? Do we anticipate that this work is to convince the philosophical world? By no means; because we well know that prevalent philosophical opinions are not formed upon the above principles, nor upon any approach to them. A recent little paper by an eminent psychologist concludes with the remark that the verdict of a majority of four of a jury, provided the individual members would form their judgments independently, would have greater probability of being true than the unanimous verdict now is. Certainly, this may be assented to; for the present verdict is not so much an opinion as a resultant of psychical and physical forces. But the remark seemed to me a pretty large concession from a man imbued with the idea of the value of modern opinion about philosophical questions formed according to that scientific method which the Germans and their admirers regard as the method of modern science — I mean, that method which puts great stress upon cooperation and solidarity of research even in the early stages of a branch of science, when independence of thought is the wholesome attitude, and gregarious thought is really sure to be wrong. For, as regards the verdict of German university professors, which, excepting at epochs of transition, has always presented a tolerable approach to unanimity upon the greater part of fundamental questions, it has always been made up as nearly as possible in the same way that the verdict of a jury is made up. Psychical forces, such as the spirit of the age, early inculcations, the spirit of loyal discipline in the general body, and that power by virtue of which one man bears down another in a negotiation, together with such physical forces as those of hunger and cold, are the forces which are mainly operative in bringing these philosophers into line; and none of these forces have any direct relation to reason. Now, these men write the larger number of those books which are so thorough and solid that every serious inquirer feels that he is obliged to read them; and his time is so engrossed by their perusal that his mind has not the leisure to digest their ideas and to reject them. Besides, he is somewhat overawed by their learning and thoroughness. This is the way in which certain opinions — or rather a certain verdict — becomes prevalent among philosophical thinkers everywhere; and reason takes hardly the leading part in the performance. It is true, that from time to time, this prevalent verdict becomes altered, in consequence of its being in too violent opposition with the changed spirit of the age; and the logic of history will usually cause such a change to be an advance toward truth in some respect. But this process is so slow, that it is not to be expected that any rational opinion about logic will become prevalent among philosophers within a generation, at least.

Nevertheless, hereafter, the man who sets up to be a logician without having gone carefully through Schröder's Logic will be tormented by the burning brand of false pretender in his conscience, until he has performed that task; and that task he cannot perform without acquiring habits of exact thinking which shall render the most of the absurdities which have hitherto been scattered over even the best of the German treatises upon logic impossible for him. Some amelioration of future treatises, therefore, though it will leave enough that is absurd, is to be expected; but it is not to be expected that those who form their opinions about logic or philosophy rationally, and therefore not gregariously, will ever comprise the majority even of philosophers. But opinions thus formed, and among such those formed by thoroughly informed and educated minds, are the only ones which need cause the homely thinker any misgiving concerning his own.

426. It is a remarkable historical fact that there is a branch of science in which there has never been a prolonged dispute concerning the proper objects of that science. It is the mathematics. Mistakes in mathematics occur not infrequently, and not being detected give rise to false doctrine, which may continue a long time. Thus, a mistake in the evaluation of a definite integral by Laplace, in his Mécanique céleste, led to an erroneous doctrine about the motion of the moon which remained undetected for nearly half a century. But after the question had once been raised, all dispute was brought to a close within a year. So, several demonstrations in the first book of Euclid, notably that of the sixteenth proposition, are vitiated by the erroneous assumption that a part is necessarily less than its whole. These remained undetected until after the theory of the non-Euclidean geometry had been completely worked out; but since that time, no mathematician has defended them; nor could any competent mathematician do so, in view of Georg Cantor's, †1 or even of Cauchy's discoveries. Incessant disputations have, indeed, been kept up by a horde of undisciplined minds about quadratures, cyclotomy, the theory of parallels, rotation, attraction, etc. But the disputants are one and all men who cannot discuss any mathematical problem without betraying their want of mathematical power and their gross ignorance of mathematics at every step. Again, there have been prolonged disputes among real mathematicians concerning questions which were not mathematical or which had not been put into mathematical form. Instances of the former class are the old dispute about the measure of force, and that lately active concerning the number of constants of an elastic body; and there have been sundry such disputes about mathematical physics and probabilities. Instances of the latter class are the disputes about the validity of reasonings concerning divergent series, imaginaries, and infinitesimals. But the fact remains that concerning strictly mathematical questions, and among mathematicians who could be considered at all competent, there has never been a single prolonged dispute.

It does not seem worth while to run through the history of science for the sake of the easy demonstration that there is no other extensive branch of knowledge of which the same can be said.

Nor is the reason for this immunity of mathematics far to seek. It arises from the fact that the objects which the mathematician observes and to which his conclusions relate are objects of his mind's own creation. Hence, although his proceeding is not infallible — which is shown by the comparative frequency with which mistakes are committed and allowed — yet it is so easy to repeat the inductions upon new instances, which can be created at pleasure, and extreme cases can so readily be found by which to test the accuracy of the processes, that when attention has once been directed to a process of reasoning suspected of being faulty, it is soon put beyond all dispute either as correct or as incorrect.

427. Hence, we homely thinkers believe that, considering the immense amount of disputation there has always been concerning the doctrines of logic, and especially concerning those which would otherwise be applicable to settle disputes concerning the accuracy of reasonings in metaphysics, the safest way is to appeal for our logical principles to the science of mathematics, where error can only long go unexploded on condition of its not being suspected.

This double assertion, first, that logic ought to draw upon mathematics for control of disputed principles, and second that ontological philosophy ought in like manner to draw upon logic, is a case under a general assertion which was made by Auguste Comte, †1 namely, that the sciences may be arranged in a series with reference to the abstractness of their objects; and that each science draws regulating principles from those superior to it in abstractness, while drawing data for its inductions from the sciences inferior to it in abstractness. So far as the sciences can be arranged in such a scale, these relationships must hold good. For if anything is true of a whole genus of objects, this truth may be adopted as a principle in studying every species of that genus. While whatever is true of a species will form a datum for the discovery of the wider truth which holds of the whole genus. Substantially the following scheme of the sciences †2 is given in the Century Dictionary †3:

Mathematics

Philosophy
Logic
Metaphysics.
Science of Time Geometry
Nomological Psychics
Nomological Physics


Molar
Molecular
Ethereal
Classificatory Psychics
Classificatory Physics


Chemistry
Biology, or the chemistry of
protoplasms
Descriptive Psychics Descriptive Physics

Practical Science.

Perhaps each psychical branch ought to be placed above the corresponding physical branch. However, only the first three branches concern us here.

428. Mathematics is the most abstract of all the sciences. For it makes no external observations, nor asserts anything as a real fact. When the mathematician deals with facts, they become for him mere "hypotheses"; for with their truth he refuses to concern himself. The whole science of mathematics is a science of hypotheses; so that nothing could be more completely abstracted from concrete reality. Philosophy is not quite so abstract. For though it makes no special observations, as every other positive science does, yet it does deal with reality. It confines itself, however, to the universal phenomena of experience; and these are, generally speaking, sufficiently revealed in the ordinary observations of every-day life. I would even grant that philosophy, in the strictest sense, confines itself to such observations as must be open to every intelligence which can learn from experience. Here and there, however, metaphysics avails itself of one of the grander generalisations of physics, or more often of psychics, not as a governing principle, but as a mere datum for a still more sweeping generalisation. But logic is much more abstract even than metaphysics. For it does not concern itself with any facts not implied in the supposition of an unlimited applicability of language.

Mathematics is not a positive science; for the mathematician holds himself free to say that A is B or that A is not B, the only obligation upon him being, that as long as he says A is B, he is to hold to it, consistently. But logic begins to be a positive science; since there are some things in regard to which the logician is not free to suppose that they are or are not; but acknowledges a compulsion upon him to assert the one and deny the other. Thus, the logician is forced by positive observation to admit that there is such a thing as doubt, that some propositions are false, etc. But with this compulsion comes a corresponding responsibility upon him not to admit anything which he is not forced to admit.

429. Logic may be defined as the science of the laws of the stable establishment of beliefs. Then, exact logic will be that doctrine of the conditions of establishment of stable belief which rests upon perfectly undoubted observations and upon mathematical, that is, upon diagrammatical, or, iconic, thought. We, who are sectaries of "exact" logic, and of "exact" philosophy, in general, maintain that those who follow such methods will, so far as they follow them, escape all error except such as will be speedily corrected after it is once suspected. For example, the opinions of Professor Schröder and of the present writer diverge as much as those of two "exact" logicians well can; and yet, I think, either of us would acknowledge that, however serious he may hold the errors of the other to be, those errors are, in the first place, trifling in comparison with the original and definite advance which their author has, by the "exact" method, been able to make in logic, that in the second place, they are trifling as compared with the errors, obscurities, and negative faults of any of those who do not follow that method, and in the third place, that they are chiefly, if not wholly, due to their author not having found a way to the application of diagrammatical thought to the particular department of logic in which they occur.

430. "Exact" logic, in its widest sense, will (as I apprehend) consist of three parts. †1 For it will be necessary, first of all, to study those properties of beliefs which belong to them as beliefs, irrespective of their stability. This will amount to what Duns Scotus †2 called speculative grammar. For it must analyse an assertion into its essential elements, independently of the structure of the language in which it may happen to be expressed. It will also divide assertions into categories according to their essential differences. The second part will consider to what conditions an assertion must conform in order that it may correspond to the "reality," that is, in order that the belief it expresses may be stable. This is what is more particularly understood by the word logic. It must consider, first, necessary, and second, probable reasoning. Thirdly, the general doctrine must embrace the study of those general conditions under which a problem presents itself for solution and those under which one question leads on to another. As this completes a triad of studies, or trivium, we might, not inappropriately, term the last study Speculative rhetoric. This division was proposed in 1867 †3 by me, but I have often designated this third part as objective logic.

431. Dr. Schröder's Logic is not intended to cover all this ground. It is not, indeed, as yet complete; and over five hundred pages may be expected yet to appear. But of the seventeen hundred and sixty-six pages which are now before the public, only an introduction of one hundred and twenty-five pages rapidly examines the speculative grammar, while all the rest, together with all that is promised, is restricted to the deductive branch of logic proper. By the phrase "exact logic" upon his title-page, he means logic treated algebraically. Although such treatment is an aid to exact logic, as defined on the last page, it is certainly not synonymous with it. The principal utility of the algebraic treatment is stated by him with admirable terseness: it is "to set this discipline free from the fetters in which language, by force of custom, has bound the human mind." †1 Upon the algebra may, however, be based a calculus, by the aid of which we may in certain difficult problems facilitate the drawing of accurate conclusions. A number of such applications have already been made; and mathematics has thus been enriched with new theorems. But the applications are not so frequent as to make the elaboration of a facile calculus one of the most pressing desiderata of the study. Professor Schröder has done a great deal in this direction; and of course his results are most welcome, even if they be not precisely what we should most have preferred to gain.

432. The introduction, which relates to first principles, while containing many excellent observations, is somewhat fragmentary and wanting in a unifying idea; and it makes logic too much a matter of feeling. †2 It cannot be said to belong to exact logic in any sense. Thus, under β (Vol. I., p. 2) the reader is told that the sciences have to suppose, not only that their objects really exist, but also that they are knowable and that for every question there is a true answer and but one. But, in the first place, it seems more exact to say that in the discussion of one question nothing at all concerning a wholly unrelated question can be implied. And, in the second place, as to an inquiry presupposing that there is some one truth, what can this possibly mean except it be that there is one destined upshot to inquiry with reference to the question in hand —one result, which when reached will never be overthrown? Undoubtedly, we hope that this, or something approximating to this, is so, or we should not trouble ourselves to make the inquiry. But we do not necessarily have much confidence that it is so. Still less need we think it is so about the majority of the questions with which we concern ourselves. But in so exaggerating the presupposition, both in regard to its universality, its precision, and the amount of belief there need be in it, Schröder merely falls into an error common to almost all philosophers about all sorts of "presuppositions." Schröder (under {e}, p. 5) undertakes to define a contradiction in terms without having first made an ultimate analysis of the proposition. The result is a definition of the usual peripatetic type; that is, it affords no analysis of the conception whatever. It amounts to making the contradiction in terms an ultimate unanalysable relation between two propositions — a sort of blind reaction between them. He goes on (under {z}, p. 9) to define, after Sigwart, logical consequentiality, as a compulsion of thought. Of course, he at once endeavors to avoid the dangerous consequences of this theory, by various qualifications. But all that is to no purpose. Exact logic will say that C's following logically from A is a state of things which no impotence of thought can alone bring about, unless there is also an impotence of existence for A to be a fact without C being a fact. Indeed, as long as this latter impotence exists and can be ascertained, it makes little or no odds whether the former impotence exists or not. And the last anchor-hold of logic he makes (under {i}) to lie in the correctness of a feeling! If the reader asks why so subjective a view of logic is adopted, the answer seems to be (under β, p. 2), that in this way Sigwart escapes the necessity of founding logic upon the theory of cognition. By the theory of cognition is usually meant an explanation of the possibility of knowledge drawn from principles of psychology. Now, the only sound psychology being a special science, which ought itself to be based upon a wellgrounded logic, it is indeed a vicious circle to make logic rest upon a theory of cognition so understood. But there is a much more general doctrine to which the name theory of cognition might be applied. Namely, it is that speculative grammar, or analysis of the nature of assertion, which rests upon observations, indeed, but upon observations of the rudest kind, open to the eye of every attentive person who is familiar with the use of language, and which, we may be sure, no rational being, able to converse at all with his fellows, and so to express a doubt of anything, will ever have any doubt. Now, proof does not consist in giving superfluous and superpossible certainty to that which nobody ever did or ever will doubt, but in removing doubts which do, or at least might at some time, arise. A man first comes to the study of logic with an immense multitude of opinions upon a vast variety of topics; and they are held with a degree of confidence, upon which, after he has studied logic, he comes to look back with no little amusement. There remains, however, a small minority of opinions that logic never shakes; and among these are certain observations about assertions. The student would never have had a desire to learn logic if he had not paid some little attention to assertion, so as at least to attach a definite signification to assertion. So that, if he has not thought more accurately about assertions, he must at least be conscious, in some out-of-focus fashion, of certain properties of assertion. When he comes to the study, if he has a good teacher, these already dimly recognised facts will be placed before him in accurate formulation, and will be accepted as soon as he can clearly apprehend their statements.

433. Let us see what some of these are. When an assertion is made, there really is some speaker, writer, or other signmaker who delivers it; and he supposes there is, or will be, some hearer, reader, or other interpreter who will receive it. It may be a stranger upon a different planet, an æon later; or it may be that very same man as he will be a second after. In any case, the deliverer makes signals to the receiver. Some of these signs (or at least one of them) are supposed to excite in the mind of the receiver familiar images, pictures, or, we might almost say, dreams — that is, reminiscences of sights, sounds, feelings, tastes, smells, or other sensations, now quite detached from the original circumstances of their first occurrence, so that they are free to be attached to new occasions. The deliverer is able to call up these images at will (with more or less effort) in his own mind; and he supposes the receiver can do the same. For instance, tramps have the habit of carrying bits of chalk and making marks on the fences to indicate the habits of the people that live there for the benefit of other tramps who may come on later. If in this way a tramp leaves an assertion that the people are stingy, he supposes the reader of the signal will have met stingy people before, and will be able to call up an image of such a person attachable to a person whose acquaintance he has not yet made. Not only is the outward significant word or mark a sign, but the image which it is expected to excite in the mind of the receiver will likewise be a sign — a sign by resemblance, or, as we say, an icon — of the similar image in the mind of the deliverer, and through that also a sign of the real quality of the thing. This icon is called the predicate of the assertion. But instead of a single icon, or sign by resemblance of a familiar image or "dream," evocable at will, there may be a complexus of such icons, forming a composite image of which the whole is not familiar. But though the whole is not familiar, yet not only are the parts familiar images, but there will also be a familiar image of its mode of composition. In fact, two types of complication will be sufficient. For example, one may be conjunctive and the other disjunctive combination. Conjunctive combination is when two images are both to be used at once; and disjunctive when one or other is to be used. (This is not the most scientific selection of types; but it will answer the present purpose.) The sort of idea which an icon embodies, if it be such that it can convey any positive information, being applicable to some things but not to others, is called a first intention. The idea embodied by an icon which cannot of itself convey any information, being applicable to everything or to nothing, but which may, nevertheless, be useful in modifying other icons, is called a second intention.

434. The assertion which the deliverer seeks to convey to the mind of the receiver relates to some object or objects which have forced themselves upon his attention; and he will miss his mark altogether unless he can succeed in forcing those very same objects upon the attention of the receiver. No icon can accomplish this, because an icon does not relate to any particular thing; nor does its idea strenuously force itself upon the mind, but often requires an effort to call it up. Some such sign as the word this, or that, or hullo, or hi, which awakens and directs attention must be employed. A sign which denotes a thing by forcing it upon the attention is called an index. An index does not describe the qualities of its object. An object, in so far as it is denoted by an index, having thisness, and distinguishing itself from other things by its continuous identity and forcefulness, but not by any distinguishing characters, may be called a hecceity. A hecceity in its relation to the assertion is a subject thereof. An assertion may have a multitude of subjects; but to that we shall return presently.

435. Neither the predicate, nor the subjects, nor both together, can make an assertion. The assertion represents a compulsion which experience, meaning the course of life, brings upon the deliverer to attach the predicate to the subjects as a sign of them taken in a particular way. This compulsion strikes him at a certain instant; and he remains under it forever after. It is, therefore, different from the temporary force which the hecceities exert upon his attention. This new compulsion may pass out of mind for the time being; but it continues just the same, and will act whenever the occasion arises, that is, whenever those particular hecceities and that first intention are called to mind together. It is, therefore, a permanent conditional force, or law. The deliverer thus requires a kind of sign which shall signify a law that to objects of indices an icon appertains as sign of them in a given way. Such a sign has been called a symbol. It is the copula of the assertion.

436. Returning to the subjects, it is to be remarked that the assertion may contain the suggestion, or request, that the receiver do something with them. For instance, it may be that he is first to take any one, no matter what, and apply it in a certain way to the icon, that he is then to take another, perhaps this time a suitably chosen one, and apply that to the icon, etc. For example, suppose the assertion is: "Some woman is adored by all catholics." The constituent icons are, in the probable understanding of this assertion, three, that of a woman, that of a person, A, adoring another, B, and that of a non-catholic. We combine the two last disjunctively, identifying the noncatholic with A; and then we combine this compound with the first icon conjunctively, identifying the woman with B. The result is the icon expressed by, "B is a woman, and moreover, either A adores B or else A is a non-catholic." The subjects are all the things in the real world past and present. From these the receiver of the assertion is suitably to choose one to occupy the place of B; and then it matters not what one he takes for A. A suitably chosen object is a woman, and any object, no matter what, adores her, unless that object be a noncatholic. This is forced upon the deliverer by experience; and it is by no idiosyncrasy of his; so that it will be forced equally upon the receiver.

437. Such is the meaning of one typical assertion. An assertion of logical necessity is simply one in which the subjects are the objects of any collection, no matter what. The consequence is, that the icon, which can be called up at will, need only to be called up, and the receiver need only ascertain by experiment whether he can distribute any set of indices in the assigned way so as to make the assertion false, in order to put the truth of the assertion to the test. For example, suppose the assertion of logical necessity is the assertion that from the proposition, "Some woman is adored by all catholics," it logically follows that "Every catholic adores some woman." That is as much as to say that, for every imaginable set of subjects, either it is false that some woman is adored by all catholics or it is true that every catholic adores some woman. We try the experiment. In order to avoid making it false that some woman is adored by all catholics, we must choose our set of indices so that there shall be one of them, B, such that, taking any one, A, no matter what, B is a woman, and moreover either A adores B or else A is a non-catholic. But that being the case, no matter what index, A, we may take, either A is a noncatholic or else an index can be found, namely, B, such that B is a woman, and A adores B. We see, then, by this experiment, that it is impossible so to take the set of indices that the proposition of consecution shall be false. The experiment may, it is true, have involved some blunder; but it is so easy to repeat it indefinitely, that we readily acquire any desired degree of certitude for the result.

438. It will be observed that this explanation of logical certitude depends upon the fact of speculative grammar that the predicate of a proposition, being essentially of an ideal nature, can be called into the only kind of existence of which it is capable, at will.

439. A not unimportant dispute has raged for many years as to whether hypothetical propositions (by which, according to the traditional terminology, I mean any compound propositions, and not merely those conditional propositions to which, since Kant, the term has often been restricted) and categorical propositions are one in essence. Roughly speaking, English logicians maintain the affirmative, Germans the negative. Professor Schröder is in the camp of the latter, I in that of the former.

440. I have maintained since 1867 that there is but one primary and fundamental logical relation, that of illation, expressed by ergo. A proposition, for me, is but an argumentation divested of the assertoriness of its premiss and conclusion. This makes every proposition a conditional proposition at bottom. In like manner a "term," or class-name, is for me nothing but a proposition with its indices or subjects left blank, or indefinite. The common noun happens to have a very distinctive character in the Indo-European languages. In most other tongues it is not sharply discriminated from a verb or participle. "Man," if it can be said to mean anything by itself, means "what I am thinking of is a man." This doctrine, which is in harmony with the above theory of signs, gives a great unity to logic; but Professor Schröder holds it to be very erroneous.

441. Cicero †1 and other ancient writers †2 mention a great dispute between two logicians, Diodorus and Philo, in regard to the significance of conditional propositions. †3 This dispute has continued to our own day. The Diodoran view seems to be the one which is natural to the minds of those, at least, who speak the European languages. How it may be with other languages has not been reported. The difficulty with this view is that nobody seems to have succeeded in making any clear statement of it that is not open to doubt as to its justice, and that is not pretty complicated. The Philonian view has been preferred by the greatest logicians. Its advantage is that it is perfectly intelligible and simple. Its disadvantage is that it produces results which seem offensive to common sense.

442. In order to explain these positions, it is best to mention that possibility may be understood in many senses; but they may all be embraced under the definition that that is possible which, in a certain state of information, is not known to be false. By varying the supposed state of information all the varieties of possibility are obtained. Thus, essential possibility is that which supposes nothing to be known except logical rules. Substantive possibility, on the other hand, supposes a state of omniscience. Now the Philonian logicians have always insisted upon beginning the study of conditional propositions by considering what such a proposition means in a state of omniscience; and the Diodorans have, perhaps not very adroitly, commonly assented to this order of procedure. Duns Scotus †1 terms such a conditional proposition a "consequentia simplex de inesse." According to the Philonians, "If it is now lightening it will thunder," understood as a consequence de inesse, means "It is either not now lightening or it will soon thunder." According to Diodorus, and most of his followers (who seem here to fall into a logical trap), it means "It is now lightening and it will soon thunder."

443. Although the Philonian views lead to such inconveniences as that it is true, as a consequence de inesse, that if the Devil were elected president of the United States, it would prove highly conducive to the spiritual welfare of the people (because he will not be elected), yet both Professor Schröder and I prefer to build the algebra of relatives upon this conception of the conditional proposition. The inconvenience, after all, ceases to seem important, when we reflect that, no matter what the conditional proposition be understood to mean, it can always be expressed by a complexus of Philonian conditionals and denials of conditionals. It may, however, be suspected that the Diodoran view has suffered from incompetent advocacy, and that if it were modified somewhat, it might prove the preferable one.

444. The consequence de inesse, "if A is true, then B is true," is expressed by letting i denote the actual state of things, Ai mean that in the actual state of things A is true, and Bi mean that in the actual state of things B is true, and then saying "If Ai is true then Bi is true," or, what is the same thing, "Either Ai is not true or Bi is true." But an ordinary Philonian conditional is expressed by saying, "In any possible state of things, i, either Ai is not true, or Bi is true."

445. Now let us express the categorical proposition, "Every man is wise." Here, we let mi mean that the individual object i is a man, and wi mean that the individual object i is wise. Then, we assert that, "taking any individual of the universe, i, no matter what, either that object, i, is not a man or that object, i, is wise"; that is, whatever is a man is wise. That is, "whatever i can indicate, either mi is not true or wi is true. The conditional and categorical propositions are expressed in precisely the same form; and there is absolutely no difference, to my mind, between them. The form of relationship is the same.

446. I find it difficult to state Professor Schröder's objection to this, because I cannot find any clear-cut, unitary conception governing his opinion. More than once in his first volume promises are held out that §28, the opening section of the second volume, shall make the matter plain. But when the second volume was published, all we found in that section was, as far as repeated examination has enabled me to see, as follows. First, hypothetical propositions, unlike categoricals, essentially involve the idea of time. When this is eliminated from the assertion, they relate only to two possibilities, what always is and what never is. Second, a categorical is always either true or false; but a hypothetical is either true, false, or meaningless. Thus, "this proposition is false" is meaningless; and another example is, "the weather will clear as soon as there is enough sky to cut a pair of trousers." Third, the supposition of negation is forced upon us in the study of hypotheticals, never in that of categoricals. Such are Schröder's arguments, to which I proceed to reply.

As to the idea of time, it may be introduced; but to say that the range of possibility in hypotheticals is always a unidimensional continuum is incorrect. "If you alone trump a trick in whist, you take it." The possibilities are that each of the four players plays any one of the four suits. There are 216 different possibilities. Certainly, the universe in hypotheticals is far more frequently finite than in categoricals. Besides, it is an ignoratio elenchi to drag in time, when no logician of the English camp has ever alleged anything about propositions involving time. That is not the question.

Every proposition is either true or false, and something not a proposition, when considered as a proposition, is, from the Philonian point of view, true. To be objectionable, a proposition must assert something; if it is merely neutral, it is not positively objectionable, that is, it is not false. "This proposition is false," far from being meaningless, is self-contradictory. That is, it means two irreconcilable things. That it involves contradiction (that is, leads to contradiction if supposed true), is easily proved. For if it be true, it is true; while if it be true, it is false. Every proposition besides what it explicitly asserts, tacitly implies its own truth. The proposition is not true unless both, what it explicitly asserts and what it tacitly implies, are true. This proposition, being self-contradictory, is false; and hence, what it explicitly asserts is true. But what it tacitly implies (its own truth) is false. †1 The difficulty about the proposition concerning the piece of blue sky is not a logical one, at all. It is no more senseless than any proposition about a "red odor" which might be a term of a categorical.

The fact stated about negation is only true of the sorts of propositions which are commonly put into categorical and hypothetical shapes, and has nothing to do with the essence of the propositions. In a paper "On the Validity of the Laws of Logic" in the Journal of Speculative Philosophy, Vol. II., †2 I have given a sophistical argument that black is white, which shows in the domain of categoricals the phenomena to which Professor Schröder refers as peculiar to hypotheticals.

The consequentia de inesse is, of course, the extreme case where the conditional proposition loses all its proper signification, owing to the absence of any range of possibilities. The conditional proper is, "In any possible case, i, either Ai is not true, or Bi is true." In the consequence de inesse the meaning sinks to, "In the true state of things, i, either Ai is not true or Bi is true."

447. My general algebra of logic (which is not that algebra of dual relations, likewise mine, which Professor Schröder prefers, although in his last volume he often uses this general algebra) consists in simply attaching indices to the letters of an expression in the Boolian algebra, making what I term a Boolian, and prefixing to this a series of "quantifiers," which are the letters Π and Σ, each with an index attached to it. Such a quantifier signifies that every individual of the universe is to be substituted for the index the Π or Σ carries, and that the non-relative product or aggregate of the results is to be taken.

448. Properly to express an ordinary conditional proposition the quantifier Π is required. In 1880, three years before I developed that general algebra, I published a paper containing a chapter on the algebra of the copula †1 (a subject I have since worked out completely in manuscript.) †2 I there noticed the necessity of such quantifiers properly to express conditional propositions; but the algebra of quantifiers not being at hand, I contented myself with considering consequences de inesse. Some apparently paradoxical results were obtained. Now Professor Schröder seems to accept these results as holding good in the general theory of hypotheticals; and then, since such results are in strong contrast with the doctrine of categoricals, he infers, in §45 of his Vol. II., a great difference between hypotheticals and categoricals. But the truth simply is that such hypotheticals want the characteristic feature of conditionals, that of a range of possibilities.

449. In connexion with this point, I must call attention to a mere algebraical difference between Schröder and me. I retain Boole's idea that there are but two values in the system of logical quantity. This harmonises with my use of the general algebra. Any two numbers may be selected to represent those values. I prefer 0 and a positive logarithmic ∞. To express that something is A and something is not A, I write:

∞ = ΣiAi ∞ = ΣjĀj

or, what is the same thing:

ΣiAi > 0 ΣjĀj > 0.

I have no objection to writing, as a mere abbreviation, which may, however, lead to difficulties, if not interpreted:

A > 0 Ā > 0.

But Professor Schröder understands these formulæ literally, and accordingly rejects Boole's conception of two values. He does not seem to understand my mode of apprehending the matter; and hence considers it a great limitation of my system that I restrict myself to two values. In fact, it is a mere difference of algebraical form of conception. I very much prefer the Boolian idea as more simple, and more in harmony with the general algebra of logic.

450. Somewhat intimately connected with the question of the relation between categoricals and hypotheticals is that of the quantification of the predicate. This is the doctrine that identity, or equality, is the fundamental relation involved in the copula. Holding as I do that the fundamental relation of logic is the illative relation, and that only in special cases does the premiss follow from the conclusion, I have in a consistent and thoroughgoing manner opposed the doctrine of the quantification of the predicate. †1 Schröder seems to admit some of my arguments; but still he has a very strong penchant for the equation.

Were I not opposed to the quantification of the predicate, I should agree with Venn †2 that it was a mistake to replace Boole's operation of arithmetical addition by the operation of logical aggregation, as most Boolians now do. I should consider the "principle of duality" †3 rather an argument against than for our modern practice. The algebra of dual relatives would be almost identical with the theory of matrices were addition retained; and this would be a great advantage.

451. It is Schröder's predilection for equations which motives his preference for the algebra of dual relatives, namely, the fact that in that algebra, even a simple undetermined inequality can be expressed as an equation. I think, too, that that algebra has merits; it certainly has uses to which Schröder seldom puts it. Yet, after all, it has too much formalism to greatly delight me — too many bushels of chaff per grain of wheat. I think Professor Schröder likes algebraic formalism better, or dislikes it less, than I.

He looks at the problems of logic through the spectacles of equations, and he formulates them, from that point of view, as he thinks, with great generality; but, as I think, in a narrow spirit. The great thing, with him, is to solve a proposition, and get a value of x, that is, an equation of which x forms one member without occurring in the other. How far such equation is iconic, that is, has a meaning, or exhibits the constitution of x, he hardly seems to care. He prefers general values to particular roots. Why? I should think the particular root alone of service, for most purposes, unless the general expressions were such that particular roots could be deduced from it — particular instances, I mean, showing the constitution of x. In most instances, a profitable solution of a mathematical problem must consist, in my opinion, of an exhaustive examination of special cases; and quite exceptional are those fortunate problems which mathematicians naturally prefer to study, where the enumeration of special cases, together with the pertinent truths about them, flow so naturally from the general statement as not to require separate examination.

I am very far from denying the interest and value of the problems to which Professor Schröder has applied himself; though there are others to which I turn by preference. Certainly, he has treated his problems with admirable power and clearness. I cannot in this place enter into the elementary explanations which would be necessary to illustrate this for more than a score of readers.

452. In respect to individuals, both non-relative and pairs, he has added some fundamental propositions to those which had been published. But he is very much mistaken in supposing that I have expressed contrary views. He simply mistakes my meaning.

453. In regard to algebraical signs, I cannot accept any of Professor Schröder's proposals except this one. While it would be a serious hindrance to the promulgation of the new doctrine to insist on new types being cut, and while I, therefore, think my own course in using the dagger as the sign of relative addition must be continued, yet I have always given that sign in its cursive form a scorpion-tail curve to the left; and it would be finical to insist on one form of curve rather than another. In almost all other cases, in my judgment, Professor Schröder's signs can never be generally received, because they are at war with a principle, the general character of which is such that Professor Schröder would be the last of all men to wish to violate it, a principle which the biologists have been led to adopt in regard to their systematic nomenclature. It is that priority must be respected, or all will fall into chaos. I will not enter further into this matter in this article. †1

454. Of what use does this new logical doctrine promise to be? The first service it may be expected to render is that of correcting a considerable number of hasty assumptions about logic which have been allowed to affect philosophy. In the next place, if Kant has shown that metaphysical conceptions spring from formal logic, this great generalisation upon formal logic must lead to a new apprehension of the metaphysical conceptions which shall render them more adequate to the needs of science. In short, "exact" logic will prove a stepping-stone to "exact" metaphysics. In the next place, it must immensely widen our logical notions. For example, a class consisting of a lot of things jumbled higgledy-piggledy must now be seen to be but a degenerate form of the more general idea of a system. Generalisation, which has hitherto meant passing to a larger class, must mean taking in the conception of the whole system of which we see but a fragment, etc., etc. In the next place, it is already evident to those who know what has already been made out, that that speculative rhetoric, or objective logic, mentioned at the beginning of this article, is destined to grow into a colossal doctrine which may be expected to lead to most important philosophical conclusions. Finally, the calculus of the new logic, which is applicable to everything, will certainly be applied to settle certain logical questions of extreme difficulty relating to the foundations of mathematics. Whether or not it can lead to any method of discovering methods in mathematics it is difficult to say. Such a thing is conceivable.

455. It is now more than thirty years since my first published contribution to "exact" logic. †2 Among other serious studies, this has received a part of my attention ever since. I have contemplated it in all sorts of perspectives and have often reviewed my reasons for believing in its importance. My confidence that the key of philosophy is here, is stronger than ever after reading Schröder's last volume. One thing which helps to make me feel that we are developing a living science, and not a dead doctrine, is the healthy mental independence it fosters, as evidenced, for example, in the divergence between Professor Schröder's opinions and mine. There is no bovine nor ovine gregariousness here. But Professor Schröder and I have a common method which we shall ultimately succeed in applying to our differences, and we shall settle them to our common satisfaction; and when that method is pouring in upon us new and incontrovertible positively valuable results, it will be as nothing to either of us to confess that where he had not yet been able to apply that method he has fallen into error.



Paper 16: The Logic of Relatives †1

§1. Three Grades of Clearness †2

456. The third volume of Professor Schröder's Exact Logic, †P1 which volume bears separately the title I have chosen for this paper, is exciting some interest even in this country. There are in America a few inquirers into logic, sincere and diligent, who are not of the genus that buries its head in the sand — men who devote their thoughts to the study with a view to learning something that they do not yet know, and not for the sake of upholding orthodoxy, or any other foregone conclusion. For them this article is written as a kind of popular exposition of the work that is now being done in the field of logic. To them I desire to convey some idea of what the new logic is, how two "algebras," that is, systems of diagrammatical representation by means of letters and other characters, more or less analogous to those of the algebra of arithmetic, have been invented for the study of the logic of relatives, and how Schröder uses one of these (with some aid from the other and from other notations) to solve some interesting problems of reasoning. I also wish to illustrate one other of several important uses to which the new logic may be put. To this end I must first clearly show what a relation is.

457. Now there are three grades of clearness in our apprehensions of the meanings of words. The first consists in the connexion of the word with familiar experience. In that sense, we all have a clear idea of what reality is and what force is — even those who talk so glibly of mental force being correlated with the physical forces. The second grade consists in the abstract definition, depending upon an analysis of just what it is that makes the word applicable. An example of defective apprehension in this grade is Professor Tait's holding (in an appendix to the reprint of his Britannica article, Mechanics) that energy is "objective" (meaning it is a substance), because it is permanent, or "persistent." For independence of time does not of itself suffice to make a substance; it is also requisite that the aggregant parts should always preserve their identity, which is not the case in the transformations of energy. The third grade of clearness consists in such a representation of the idea that fruitful reasoning can be made to turn upon it, and that it can be applied to the resolution of difficult practical problems.

§2. Of the Term Relation in Its First Grade of Clearness

458. An essential part of speech, the Preposition, exists for the purpose of expressing relations. Essential it is, in that no language can exist without prepositions, either as separate words placed before or after their objects, as case-declensions, as syntactical arrangements of words, or some equivalent forms. Such words as "brother," "slayer," "at the time," "alongside," "not," "characteristic property" are relational words, or relatives, in this sense, that each of them becomes a general name when another general name is affixed to it as object. In the Indo-European languages, in Greek, for example, the so-called genitive case (an inapt phrase like most of the terminology of grammar) is, very roughly speaking, the form most proper to the attached name. By such attachments, we get such names as "brother of Napoleon," "slayer of giants," "{epi 'Ellissaiou}, at the time of Elias," "{para allélön}, alongside of each other," "not guilty," "a characteristic property of gallium." Not is a relative because it means "other than"; scarcely, though a relational word of highly complex meaning, is not a relative. It has, however, to be treated in the logic of relatives. Other relatives do not become general names until two or more names have been thus affixed. Thus, "giver to the city" is just such a relative as the preceding; for "giver to the city of a statue of himself" is a complete general name (that is, there might be several such humble admirers of themselves, though there be but one, as yet); but "giver" requires two names to be attached to it, before it becomes a complete name. The dative case is a somewhat usual form for the second object. The archaic, instrumental, and locative cases were serviceable for third and fourth objects.

459. Our European languages are peculiar in their marked differentiation of common nouns from verbs. Proper nouns must exist in all languages; and so must such "pronouns," or indicative words, as this, that, something, anything. But it is probably true that in the great majority of the tongues of men, distinctive common nouns either do not exist or are exceptional formations. In their meaning as they stand in sentences, and in many comparatively widely-studied languages, common nouns are akin to participles, as being mere inflexions of verbs. If a language has a verb meaning "is a man," a noun "man" becomes a superfluity. For all men are mortals is perfectly expressed by "Anything either is-a-man not or is-a-mortal." Some man is a miser is expressed by "Something both is-a-man and is-a-miser." The best treatment of the logic of relatives, as I contend, will dispense altogether with class names and only use such verbs. A verb requiring an object or objects to complete the sense may be called a complete relative.

A verb by itself signifies a mere dream, an imagination unattached to any particular occasion. It calls up in the mind an icon. A relative is just that, an icon, or image, without attachments to experience, without "a local habitation and a name," but with indications of the need of such attachments.

460. An indexical word, such as a proper noun or demonstrative or selective pronoun, has force to draw the attention of the listener to some hecceity common to the experience of speaker and listener. By a hecceity, I mean, some element of existence which, not merely by the likeness between its different apparitions, but by an inward force of identity, manifesting itself in the continuity of its apparition throughout time and in space, is distinct from everything else, and is thus fit (as it can in no other way be) to receive a proper name or to be indicated as this or that. Contrast this with the signification of the verb, which is sometimes in my thought, sometimes in yours, and which has no other identity than the agreement between its several manifestations. That is what we call an abstraction or idea. The nominalists say it is a mere name. Strike out the "mere," and this opinion is approximately true. The realists say it is real. Substitute for "is," may be, that is, is provided experience and reason shall, as their final upshot, uphold the truth of the particular predicate, and the natural existence of the law it expresses, and this is likewise true. It is certainly a great mistake to look upon an idea, merely because it has not the mode of existence of a hecceity, as a lifeless thing.

461. The proposition, or sentence, signifies that an eternal fitness, or truth, a permanent conditional force, or law, attaches certain hecceities to certain parts of an idea. Thus, take the idea of "buying by — of — from — in exchange for —." This has four places where hecceities, denoted by indexical words, may be attached. The proposition "A buys B from C at the price D," signifies an eternal, irrefragable, conditional force gradually compelling those attachments in the opinions of inquiring minds.

462. Whether or not there be in the reality any definite separation between the hecceity-element and the idea-element is a question of metaphysics, not of logic. But it is certain that in the expression of a fact we have a considerable range of choice as to how much we will denote by the indexical and how much signify by iconic words. Thus, we have stated "all men are mortal" in such a form that there is but one index. But we may also state it thus: "Taking anything, either it possesses not humanity or it possesses mortality." Here "humanity" and "mortality" are really proper names, or purely denotative signs, of familiar ideas. Accordingly, as here stated, there are three indices. Mathematical reasoning largely depends on this treatment of ideas as things †1; for it aids in the iconic representation of the whole fact. Yet for some purposes it is disadvantageous. These truths will find illustration in §13 below.

463. Any portion of a proposition expressing ideas but requiring something to be attached to it in order to complete the sense, is in a general way relational. But it is only a relative in case the attachment of indexical signs will suffice to make it a proposition, or, at least, a complete general name. Such a word as exceedingly or previously is relational, but is not a relative, because significant words require to be added to it to make complete sense.

§3. Of Relation in the Second Grade of Clearness

464. Is relation anything more than a connexion between two things? For example, can we not state that A gives B to C without using any other relational phrase than that one thing is connected with another? Let us try. We have the general idea of giving. Connected with it are the general ideas of giver, gift, and "donée." We have also a particular transaction connected with no general idea except through that of giving. We have a first party connected with this transaction and also with the general idea of giver. We have a second party connected with that transaction, and also with the general idea of "donée." We have a subject connected with that transaction and also with the general idea of gift. A is the only hecceity directly connected with the first party; C is the only hecceity directly connected with the second party, B is the only hecceity directly connected with the subject. Does not this long statement amount to this, that A gives B to C?

In order to have a distinct conception of Relation, it is necessary not merely to answer this question but to comprehend the reason of the answer. I shall answer it in the negative. For, in the first place, if relation were nothing but connexion of two things, all things would be connected. For certainly, if we say that A is unconnected with B, that non-connexion is a relation between A and B. Besides, it is evident that any two things whatever make a pair. Everything, then, is equally related to everything else, if mere connexion be all there is in relation. But that which is equally and necessarily true of everything is no positive fact, at all. This would reduce relation, considered as simple connexion between two things, to nothing, unless we take refuge in saying that relation in general is indeed nothing, but that modes of relation are something. If, however, these different modes of relation are different modes of connexion, relation ceases to be simple bare connexion. Going back, however, to the example of the last paragraph, it will be pointed out that the peculiarity of the mode of connexion of A with the transaction consists in A's being in connexion with an element connected with the transaction, which element is connected with the peculiar general idea of a giver. It will, therefore, be said, by those who attempt to defend an affirmative answer to our question, that the peculiarity of a mode of connexion consists in this, that that connexion is indirect and takes place through something which is connected with a peculiar general idea. But I say that is no answer at all; for if all things are equally connected, nothing can be more connected with one idea than with another. This is unanswerable. Still, the affirmative side may modify their position somewhat. They may say, we grant that it is necessary to recognise that relation is something more than connexion; it is positive connexion. Granting that all things are connected, still all are not positively connected. The various modes of relationship are, then, explained as above. But to this I reply: you propose to make the peculiarity of the connexion of A with the transaction depend (no matter by what machinery) upon that connexion having a positive connexion with the idea of a giver. But "positive connexion" is not enough; the relation of the general idea is quite peculiar. In order that it may be characterised, it must, on your principles, be made indirect, taking place through something which is itself connected with a general idea. But this last connexion is again more than a mere general positive connexion. The same device must be resorted to, and so on ad infinitum. In short, you are guilty of a circulus in definiendo. You make the relation of any two things consist in their connexion being connected with a general idea. But that last connexion is, on your own principles, itself a relation, and you are thus defining relation by relation; and if for the second occurrence you substitute the definition, you have to repeat the substitution ad infinitum.

The affirmative position has consequently again to be modified. But, instead of further tracing possible tergiversations, let us directly establish one or two positive positions. In the first place, I say that every relationship concerns some definite number of correlates. Some relations have such properties that this fact is concealed. Thus, any number of men may be brothers. Still, brotherhood is a relation between pairs. If A, B, and C are all brothers, this is merely the consequence of the three relations, A is brother of B, B is brother of C, C is brother of A. Try to construct a relation which shall exist either between two or between three things such as "— is either a brother or betrayer of — to —." You can only make sense of it by somehow interpreting the dual relation as a triple one. We may express this as saying that every relation has a definite number of blanks to be filled by indices, or otherwise. In the case of the majority of relatives, these blanks are qualitatively different from one another. These qualities are thereby communicated to the connexions.

465. In a complete proposition there are no blanks. It may be called a medad, or medadic relative, from {médamos}, none, and {-ada} the accusative ending of such words as {monas, dyas, trias, tetras} etc. †P1 A non-relative name with a substantive verb, as "— is a man," or "man that is —," or "—'s manhood" has one blank; it is a monad, or monadic relative. An ordinary relative with an active verb as "— is a lover of —" or "the loving by — of —" has two blanks; it is a dyad, or dyadic relative. A higher relative similarly treated has a plurality of blanks. It may be called a polyad. The rank of a relative among these may be called its adinity, that is, the peculiar quality of the number it embodies.

466. A relative, then, may be defined as the equivalent of a word or phrase which, either as it is (when I term it a complete relative), or else when the verb "is" is attached to it (and if it wants such attachment, I term it a nominal relative), becomes a sentence with some number of proper names left blank. A relationship, or fundamentum relationis, is a fact relative to a number of objects, considered apart from those objects, as if, after the statement of the fact, the designations of those objects had been erased. A relation is a relationship considered as something that may be said to be true of one of the objects, the others being separated from the relationship yet kept in view. Thus, for each relationship there are as many relations as there are blanks. For example, corresponding to the relationship which consists in one thing loving another there are two relations, that of loving and that of being loved by. There is a nominal relative for each of these relations, as "lover of —," and "loved by —." These nominal relatives belonging to one relationship, are in their relation to one another termed correlatives. In the case of a dyad, the two correlatives, and the corresponding relations are said, each to be the converse of the other. The objects whose designations fill the blanks of a complete relative are called the correlates. The correlate to which a nominal relative is attributed is called the relate.

467. In the statement of a relationship, the designations of the correlates ought to be considered as so many logical subjects and the relative itself as the predicate. The entire set of logical subjects may also be considered as a collective subject, of which the statement of the relationship is predicate.

§4. Of Relation in the Third Grade of Clearness †1

468. Mr. A. B. Kempe has published in the Philosophical Transactions †2 a profound and masterly "Memoir on the Theory of Mathematical Form," which treats of the representation of relationships by "Graphs," which is Clifford's name for a diagram, consisting of spots and lines, in imitation of the chemical diagrams showing the constitution of compounds. Mr. Kempe seems to consider a relationship to be nothing but a complex of bare connexions of pairs of objects, the opinion refuted in the last section. Accordingly, while I have learned much from the study of his memoir, I am obliged to modify what I have found there so much that it will not be convenient to cite it; because long explanations of the relation of my views to his would become necessary if I did so.

469. A chemical atom is quite like a relative in having a definite number of loose ends or "unsaturated bonds," corresponding to the blanks of the relative. †3 In a chemical molecule, each loose end of one atom is joined to a loose end, which it is assumed must belong to some other atom, although in the vapor of mercury, in argon, etc., two loose ends of the same atom would seem to be joined; and why pronounce such hermaphroditism impossible? Thus the chemical molecule is a medad, like a complete proposition. Regarding proper names and other indices, after an "is" has been attached to them, as monads, they, together with other monads, correspond to the two series of chemical elements, H, Li, Na, K, Rb, Cs, etc., and Fl, Cl, Br, I. The dyadic relatives correspond to the two series, Mg, Ca, Sr, Ba, etc., and O, S, Se, Te, etc. The triadic relatives correspond to the two series B, Al, Zn, In, Tl, etc., and N, P, As, Sb, Bi, etc. Tetradic relatives are, as we shall see, a superfluity; they correspond to the series C, Si, Ti, Sn, Ta, etc. The proposition "John gives John to John" corresponds in

inline image inline image
Figure 1 Figure 2

its constitution, as Figs. 1 and 2 show, precisely to ammonia.

470. But beyond this point the analogy ceases to be striking. In fact, the analogy with the ruling theory of chemical compounds quite breaks down. Yet I cannot resist the temptation to pursue it. After all, any analogy, however fanciful, which serves to focus attention upon matters which might otherwise escape observation is valuable. A chemical compound might be expected to be quite as much like a proposition as like an algebraical invariant; and the brooding upon chemical graphs has hatched out an important theory in invariants. †1 Fifty years ago, when I was first studying chemistry, the theory was that every compound consisted of two oppositely electrified atoms or radicles; and in like manner every compound radicle consisted of two opposite atoms or radicles. The argument to this effect was that chemical attraction is evidently between things unlike one another and evidently has a saturation point; and further that we observe that it is the elements the most extremely unlike which attract one another. Julius Lothar Meyer's curve having for its ordinates the atomic volumes of the elements and for its abscissas their atomic weights tends to support the opinion that elements strongly to attract one another must have opposite characters †1; for we see that it is the elements on the steepest downward slopes of that curve which have the strongest attractions for the elements on the steepest upward inclines. But when chemists became convinced of the doctrine of valency, that is, that every element has a fixed number of loose ends, and when they consequently began to write graphs for compounds, it seems to have been assumed that this necessitated an abandonment of the position that atoms and radicles combine by opposition of characters, which had further been weakened by the refutation of some mistaken arguments in its favor. But if chemistry is of no aid to logic, logic here comes in to enlighten chemistry. For in logic, the medad must always be composed of one part having a negative, or antecedental, character, and another part of a positive, or consequental, character; and if either of these parts is compound its constituents are similarly related to one another. Yet this does not, at all, interfere with the doctrine that each relative has a definite number of blanks or loose ends. We shall find that, in logic, the negative character is a character of reversion in this sense, that if the negative part of a medad is compound, its negative part has, on the whole, a positive character. We shall also find, that if the negative part of a medad is compound, the bond joining its positive and negative parts has its character reversed, just as those relatives themselves have. †2

471. Several propositions are in this last paragraph stated about logical medads which now must be shown to be true. In the first place, although it be granted that every relative has a definite number of blanks, or loose ends, yet it would seem, at first sight, that there is no need of each of these joining no more than one other. For instance, taking the triad

inline image

Figure 3

"— kills — to gratify —," why may not the three loose ends all join in one node and then be connected with the loose end of the monad "John is —" as in Figure 3 making the proposition "John it is that kills what is John to gratify what is John"? The answer is, that a little exercise of generalising power will show that such a four-way node is really a tetradic relative, which may be expressed in words thus, "— is identical with — and with — and with —"; so that the medad is

inline image

Figure 4

really equivalent to that of Figure 4, which corresponds to prussic acid as shown in Figure 5.

inline image

Figure 5

Thus, it becomes plain that every node of bonds is equivalent to a relative; and the doctrine of valency is established for us in logic.

472. We have next to inquire into the proposition that in every combination of relatives there is a negative and a positive constituent. This is a corollary from the general logical doctrine of the illative character of the copula, a doctrine precisely opposed to the opinion of the quantification of the predicate. A satisfactory discussion of this fundamental question would require a whole article. I will only say in outline that it can be positively demonstrated in several ways that a proposition of the form "man = rational animal," is a compound of propositions each of a form which may be stated thus: "Every man (if there be any) is a rational animal" or "Men are exclusively (if anything) rational animals." Moreover, it must be acknowledged that the illative relation (that expressed by "therefore") is the most important of logical relations, the be-all and the end-all of the rest. It can be demonstrated that formal logic needs no other elementary logical relation than this; but that with a symbol for this and symbols of relatives, including monads, and with a mode of representing the attachments of them, all syllogistic may be developed, far more perfectly than any advocate of the quantified predicate ever developed it, and in short in a way which leaves nothing to be desired. This in fact will be virtually shown in the present paper. It can further be shown that no other copula will of itself suffice for all purposes. Consequently, the copula of equality ought to be regarded as merely derivative.

473. Now, in studying the logic of relatives we must sedulously avoid the error of regarding it as a highly specialised doctrine. It is, on the contrary, nothing but formal logic generalised to the very tip-top. In accordance with this view, or rather with this theorem (for it is susceptible of positive demonstration), we must regard the relative copula, which is the bond between two blanks of relatives, as only a generalisation of the ordinary copula, and thus of the "ergo." When we say that from the proposition A the proposition B necessarily follows, we say that "the truth of A in every way in which it can exist at all is the truth of B," or otherwise stated "A is true only in so far as B is true." This is the very same relation which we express when we say that "every man is mortal," or "men are exclusively mortals." For this is the same as to say, "Take anything whatever, M; then, if M is a man, it follows necessarily that M is mortal." This mode of junction is essentially the same as that between the relatives in the compound relative "lover, in every way in which it may be a lover at all, of a servant," or, otherwise expressed, "lover (if at all) exclusively of servants." For to say that "Tom is a lover (if at all) only of servants of Dick," is the same as to say "Take anything whatever, M; then, if M is loved by Tom, M is a servant of Dick," or "everything there may be that is loved by Tom is a servant of Dick." †1

474. Now it is to be observed that the illative relation is not simply convertible; that is to say, that "from A necessarily follows B" does not necessarily imply that "from B necessarily follows A." Among the vagaries of some German logicians of some of the inexact schools, the convertibility of illation (like almost every other imaginable absurdity) has been maintained; but all the other inexact schools deny it, and exact logic condemns it, at once. Consequently, the copula of inclusion, which is but the ergo freed from the accident of asserting the truth of its antecedent, is equally inconvertible. For though "men include only mortals," it does not follow that "mortals include only men," but, on the contrary, what follows is "mortals include all men." Consequently, again, the fundamental relative copula is inconvertible. That is, because "Tom loves (if anybody) only a servant (or servants) of Dick," it does not follow that "Dick is served (if at all) only by somebody loved by Tom," but, on the contrary, what follows is "Dick is master of every person (there may be) who is loved by Tom." †1 We thus see, clearly, first, that, as the fundamental relative copula, we must take that particular mode of junction; secondly, that that mode is at bottom the mode of junction of the ergo, and so joins a relative of antecedental character to a relative of consequental character; and, thirdly, that that copula is inconvertible, so that the two kinds of constituents are of opposite characters. There are, no doubt, convertible modes of junction of relatives, as in "lover of a servant" †P1; but it will be shown below that these are complex and indirect in their constitution.

475. It remains to be shown that the antecedent part of a medad has a negative, or reversed, character, and how this, in case it be compound, affects both its relatives and their bonds. But since this matter is best studied in examples, I will first explain how I propose to draw the logical graphs.

It is necessary to use, as the sign of the relative copula, some symbol which shall distinguish the antecedent from the consequent; and since, if the antecedent is compound (owing to the very character which I am about to demonstrate, namely, its reversing the characters of the relatives and the bonds it contains), it is very important to know just how much is included in that antecedent, while it is a matter of comparative indifference how much is included in the consequent (though it is simply everything not in the antecedent), and since further (for the same reason) it is important to know how many antecedents, each after the first a part of another, contain a given relative or copula, I find it best to make the line which joins antecedent and consequent encircle the whole of the former. Letters of the alphabet may be used as abbreviations of complete relatives; and the proper number of bonds may be attached to each. If one of these is encircled, that circle must have a bond corresponding to each bond of the encircled letter. Chemists sometimes write above atoms Roman numerals to indicate their adinities; but I do not think this necessary. Figure 7 shows, in a complete medad, my sign of the

inline image inline image
Figure 7 Figure 8

relative copula. Here, h is the monad "— is a man," and d is the monad "— is mortal." The antecedent is completely enclosed, and the meaning is "Anything whatever, if it be a man, is mortal." If the circle encloses a dyadic or polyadic relative, it must, of course, have a tail for every bond of that relative. Thus, in Figure 8, l is the dyad "— loves —," and it is important to remark that the bond to the left is the lover and that to the right is the loved. Monads are the only relatives for which we need not be attentive to the positions of attachment of the bonds. In this figure, w is the monad "— is wise," and v is the monad "— is virtuous." The l and v are enclosed in a large common circle. Had this not been done, the medad could not be read (as far as any rules yet given show), because it would not consist of antecedent and consequent. As it is, we begin the reading of the medad at the bond connecting antecedent and consequent. Every bond of a logical graph denotes a hecceity; and every unencircled bond (as this one is) stands for any hecceity the reader may choose from the universe. This medad evidently refers to the universe of men. Hence the interpretation begins: "Let M be any man you please." We proceed along this bond in the direction of the antecedent, and on entering the circle of the antecedent we say: "If M be." We then enter the inner circle. Now, entering a circle means a relation to every. Accordingly we add "whatever." Traversing l from left to right, we say "lover." (Had it been from right to left we should have read it "loved.") Leaving the circle is the mark of a relation "only to," which words we add. Coming to v we say "what is virtuous." Thus our antecedent reads: "Let M be any man you please. If M be whatever it may that is lover only to the virtuous." We now return to the consequent and read, "M is wise." Thus the whole means, "Whoever loves only the virtuous is wise."

As another example, take the graph of Figure 9, where l has

inline image

Figure 9

the same meaning as before and m is the dyad "— is mother of —." Suppose we start with the left hand bond. We begin with saying "Whatever." Since cutting this bond does not sever the medad, we proceed at once to read the whole as an unconditional statement and we add to our "whatever" "there is." We can now move round the ring of the medad either clockwise or counter-clockwise. Taking the last way, we come to l from the left hand and therefore add "is a lover." Moving on, we enter the circle round m; and entering a circle is a sign that we must say "of everything that." Since we pass through m backwards we do not read "is mother" but "is mothered" or "has for mother." Then, since we pass out of the circle we should have to add "only"; but coming back, as we do, to the starting point, we need only say "that same thing." Thus, the interpretation is "Whatever there is, is lover of everything that has for mother that same thing," or "Every woman loves everything of which she is mother." Starting at the same point and going round the other way, the reading would be "Everybody is mother (if at all) only of what is loved by herself." Starting on the right and proceeding clockwise, "Everything is loved by every mother of itself." Proceeding counter-clockwise, "Everything has for mothers only lovers of itself."

476. Triple relatives afford no particular difficulty. Thus, in Figure 10, w and v have the same significations as before; r is the monad, "— is a reward," and g is the triad "— gives | to —." It can be read either

inline image

Figure 10

"Whatever is wise gives every reward to every virtuous person," or "Every virtuous person has every reward given to him by everybody that is wise," or "Every reward is given by everybody who is wise to every virtuous person."

477. A few more examples will be instructive. Figure 11, where A is the proper name "Alexander" means "Alexander loves only the virtuous," i.e., "Take anybody you please; then, if he be Alexander and if he loves anybody, this latter is virtuous."

inline image inline image inline image
Figure 11 Figure 12 Figure 13

If you attempt, in reading this medad, to start to the right of l, you fall into difficulty, because your antecedent does not then consist of an antecedent and consequent, but of two circles joined by a bond, a combination to be considered below. But Figure 12 may be read with equal ease on whichever side of l you begin, whether as "whoever is wise loves everybody that is virtuous," or "whoever is virtuous is loved by everybody that is wise." If in Figure 13 -b- be the dyad "— is a benefactor of —," the medad reads, "Alexander stands only to virtuous persons in the relation of loving only their benefactors."

Figure 14, where -s- is the dyad "— is a servant of —" may be read, according to the above principles, in the several ways following:

"Whoever stands to any person in the relation of lover to none but his servants benefits him."

"Every person stands only to a person benefited by him in the relation of a lover only of a servant of that person."

inline image inline image
Figure 14 Figure 15

"Every person, M, is benefactor of everybody who stands to M in the relation of being served by everybody loved by him."

"Every person, N, is benefited by everybody who stands to N in the relation of loving only servants of him."

"Every person, N, stands only to a benefactor of N in the relation of being served by everybody loved by him."

"Take any two persons, M and N. If, then, N is served by every lover of M, N is benefited by M."

Figure 15 represents a medad which means, "Every servant of any person, is a benefactor of whomever may be loved by that person." Equivalent statements easily read off from the graphs are as follows:

"Anybody, M, no matter who, is servant (if at all) only of somebody who loves (if at all) only persons benefited by M."

"Anybody, no matter who, stands to every master of him in the relation of benefactor of whatever person may be loved by him."

"Anybody, no matter who, stands to whoever loves him in the relation of being benefited by whatever servant he may have."

"Anybody, N, is loved (if at all) only by a person who is served (if at all) only by benefactors of N."

"Anybody, no matter who, loves (if at all) only persons benefited by all servants of his."

"Anybody, no matter who, is served (if at all) only by benefactors of everybody loved by him."

478. I will now give an example containing triadic relatives, but no monads. Let p be "— prevents — from communicating with —," the second blank being represented by a bond from the right of p and the third by a bond from below p. Let β mean "— would betray — to —," the arrangement of bonds being the same as with p. Then, Figure 16 means that "whoever loves only persons who prevent every servant

inline image

Figure 16

of any person, A, from communicating with any person, B, would betray B to A." I will only notice one equivalent statement, viz.: "Take any three persons, A, B, C, no matter who. Then, either C betrays B to A, or else two persons, M and N, can be found, such that M does not prevent N from communicating with B, although M is loved by C and N is a servant of A."

479. This last interpretation is an example of the method which is, by far, the plainest and most unmistakable of any in complicated cases. The rule for producing it is as follows:

1. Assign a letter of the alphabet to denote the hecceity represented by each bond. †P1

2. Begin by saying: "Take any things you please, namely," and name the letters representing bonds not encircled; then add, "Then suitably select objects, namely," and name the letters representing bonds each once encircled; then add, "Then take any things you please, namely," and name the letters representing bonds each twice encircled. Proceed in this way until all the letters representing bonds have been named, no letter being named until all those encircled fewer times have been named; and each hecceity corresponding to a letter encircled odd times is to be suitably chosen according to the intent of the assertor of the medad proposition, while each hecceity corresponding to a bond encircled even times is to be taken as the interpreter or the opponent of the proposition pleases. †1

3. Declare that you are about to make statements concerning certain propositions, to which, for the sake of convenience, you will assign numbers in advance of enunciating them or stating their relations to one another. These numbers are to be formed in the following way. There is to be a number for each letter of the medad (that is for those which form spots of the graph, not for the letters assigned by clause 1 of this rule to the bonds), and also a number for each circle round more than one letter; and the first figure of that number is to be a 1 or a 2, according as the letter or the circle is in the principal antecedent or the principal consequent; the second figure is to be 1 or 2, according as the letter or the circle belongs to the antecedent or the consequent of the principal antecedent or consequent, and so on.

Declare that one or other of those propositions whose numbers contain no 1 before the last figure is true. Declare that each of those propositions whose numbers contain an odd number of 1's before the last figure consists in the assertion that some one or another of the propositions whose numbers commence with its number is true. For example, 11 consists in the assertion that either 111 or 1121 or 1122 is true, supposing that these are the only propositions whose numbers commence with 11. Declare that each of those propositions whose numbers contain an even number of 1's (or none) before the last figure consists in the assertion that every one of the propositions whose numbers commence with its number is true. Thus, 12 consists in the assertion that 121, 1221, 1222 are all true, provided those are the only propositions whose numbers commence with 12. The process described in this clause will be abridged except in excessively complicated cases.

4. Finally, you are to enunciate all those numbered propositions which correspond to single letters. Namely, each proposition whose number contains an even number of 1's, will consist in affirming the relative of the spot-letter to which that number corresponds after filling each blank with that bond-letter which by clause 1 of this rule was assigned to the bond at that blank. But if the number of the proposition contains an odd number of 1's, the relative, with its blanks filled in the same way, is to be denied.

480. In order to illustrate this rule, I will restate the meanings of the medads of Figures 7-16, in all the formality of the rule; although such formality is uncalled for and awkward, except in far more complicated cases.

Figure 7. Let A be anything you please. There are two propositions, 1 and 2, one of which is true. Proposition 1 is, that A is not a man. Proposition 2 is, that A is mortal. More simply, Whatever A may be, either A is not a man or A is mortal.

Figure 8. Let A be anybody you please. Then, I will find a person, B, so that either proposition 1 or proposition 2 shall be true. Proposition 1 asserts that both propositions 11 and 12 are true. Proposition 11 is that A loves B. Proposition 12 is that B is not virtuous. Proposition 2 is that A is wise. More simply, Take anybody, A, you please. Then, either A is wise, or else a person, B, can be found such that B is not virtuous and A loves B.

Figure 9. Let A and B be any persons you please. Then, either proposition 1 or proposition 2 is true. Proposition 1 is that A is not a mother of B. Proposition 2 is that A loves B. More simply, whatever two persons A and B may be, either A is not a mother of B or A loves B.

Figure 10. Let A, B, C be any three things you please. Then, one of the propositions numbered, 1, 21, 221, 222 is true. Proposition 1 is that A is not wise. Proposition 21 is that B is not a reward. Proposition 221 is that C is not virtuous. Proposition 222 is that A gives B to C. More simply, take any three things, A, B, C, you please. Then, either A is not wise, or B is not a reward, or C is not virtuous, or A gives B to C.

Figure 11. Take any two persons, A and B, you please. Then, one of the propositions 1, 21, 22 is true. 1 is that A is not Alexander. 21 is that A does not love B. Proposition 3 is that B is virtuous.

Figure 12. Take any two persons, A and B. Then, one of the propositions 1, 21, 22 is true. 1 is that A is not wise. 21 is that B is not virtuous. 22 is that A loves B.

Figure 13. Take any two persons, A and C. Then a person, B can be found such that one of the propositions 1, 21, 22 is true. Proposition 21 asserts that both 211 and 212 are true. Proposition 1 that A is not Alexander. Proposition 211 is that A loves B. Proposition 212 is that B does not benefit C. Proposition 22 is that C is virtuous. More simply, taking any two persons, A and C, either A is not Alexander, or C is virtuous, or there is some person, B, who is loved by A without benefiting C.

Figure 14. Take any two persons, A and B, and I will then select a person C. Either proposition 1 or proposition 2 is true. Proposition 1 is that both 11 and 12 are true. Proposition 11 is that A loves C. Proposition 12 is that C is not a servant of B. Proposition 2 is that A benefits B. More simply, of any two persons, A and B, either A benefits the other, B, or else there is a person, C, who is loved by A but is not a servant of B.

Figure 15. Take any three persons, A, B, C. Then one of the propositions 1, 21, 22 is true. 1 is that A is not a servant of B; 21 is that B is not a lover of C; 22 is that A benefits C.

Figure 16. Take any three persons, A, B, C. Then I can so select D and E, that one of the propositions 1 or 2 is true. 1 is that 11 and 121 and 122 are all true. 11 is that A loves D, 121 is that E is a servant of C, 122 is that D does not prevent E from communicating with B. 2 is that A betrays B to C.

I have preferred to give these examples rather than fill my pages with a dry abstract demonstration of the correctness of the rule. If the reader requires such a proof, he can easily construct it. This rule makes evident the reversing effect of the encirclements, not only upon the "quality" of the relatives as affirmative or negative, but also upon the selection of the hecceities as performable by advocate or opponent of the proposition, as well as upon the conjunctions of the propositions as disjunctive or conjunctive, or (to avoid this absurd grammatical terminology) as alternative or simultaneous.

481. It is a curious example of the degree to which the thoughts of logicians have been tied down to the accidents of the particular language they happened to write (mostly Latin), that while they hold it for an axiom that two nots annul one another, it was left for me to say as late as 1867 †P1 that some in formal logic ought to be understood, and could be understood, so that some-some should mean any. I suppose that were ordinary speech of any authority as to the forms of logic, in the overwhelming majority of human tongues two negatives intensify one another. And it is plain that if "not" be conceived as less than anything, what is less than that is a fortiori not. On the other hand, although some is conceived in our languages as more than none, so that two "somes" intensify one another, yet what it ought to signify for the purposes of syllogistic is that, instead of the selection of the instance being left — as it is, when we say "any man is not good" — to the opponent of the proposition, when we say "some man is not good," this selection is transferred to the opponent's opponent, that is to the defender of the proposition. Repeat the some, and the selection goes to the opponent's opponent's opponent, that is, to the opponent again, and it becomes equivalent to any. In more formal statement, to say "Every man is mortal," or "Any man is mortal," is to say, "A man, as suitable as any to prove the proposition false, is mortal," while "Some man is mortal" is equivalent to "A man, as suitable as any to prove the proposition not false, is mortal." "Some-some man is mortal" is accordingly "A man, as suitable as any to prove the proposition not not-false, is mortal."

482. In like manner, encircled 2N + 1 times, a disjunctive conjunction of propositions becomes a copulative conjunction. Here, the case is altogether similar. Encircled even times, the statement is that some one (or more) of the propositions is true; encircled odd times, the statement is that any one of the propositions is true. The negative of "lover of every servant" is "non-lover of some servant." The negative of "lover every way (that it is a lover) of a servant" is "lover some way of a non-servant."

The general nature of a relative and of a medad has now been made clear. At any rate, it will become so, if the reader carefully goes through with the explanations. We have not, however, as yet shown how every kind of proposition can be graphically expressed, nor under what conditions a medad is necessarily true. For that purpose it will be necessary to study certain special logical relatives.

§5. Triads, the Primitive Relatives

483. That out of triads all polyads can be constructed is made plain by Figure 17.

inline image

Figure 17

484. Figure 18 shows that from two triads a dyad can be made. Figure 19 shows that from one triad a monad can be made. Figure 20 shows that from any even number of triads

inline image inline image inline image
Figure 18 Figure 19 Figure 20

a medad can be made. In general, the union of a μ-ad and a ν-ad gives a (μ + ν - 2λ)-ad, where λ is the number of bonds of union. This formula shows that artiads, or even-ads, can produce only artiads. But any perissid, or odd-ad (except a monad), can by repetition produce a relative of any adinity.

485. Since the principal object of a notation for relatives is not to produce a handy calculus for the solution of special logical problems, but to help the study of logical principles, the study of logical graphs from that point of view must be postponed to a future occasion. For present purposes that notation is best which carries analysis the furthest, and presents the smallest number of unanalyzed forms. It will be best, then, to use single letters for relatives of some one definite and odd number of blanks. We naturally choose three as the smallest number which will answer the purpose.

486. We shall, therefore, substitute for such a dyad as "— is lover of —" some such triad as "— is coexistent with | and a lover of —." If, then, we make -w- to signify "— is coexistent with | and with —," that which we have hitherto written as in Figure 12 will be written as in Figure 21. But having once recognised that such a mode of writing is possible,

inline image

Figure 21

we can continue to use our former methods, provided we now consider them as abbreviations.

487. The logical doctrine of this section, must, we may remark, find its application in metaphysics, if we are to accept the Kantian principle that metaphysical conceptions mirror those of formal logic.

§6. Relatives of Second Intention

488. The general method of graphical representation of propositions has now been given in all its essential elements, except, of course, that we have not, as yet, studied any truths concerning special relatives; for to do so would seem, at first, to be "extralogical." Logic in this stage of its development may be called paradisaical logic, because it represents the state of Man's cognition before the Fall. For although, with this apparatus, it is easy to write propositions necessarily true, it is absolutely impossible to write any which is necessarily false, or, in any way which that stage of logic affords, to find out that anything is false. The mind has not as yet eaten of the fruit of the Tree of Knowledge of Truth and Falsity. Probably it will not be doubted that every child in its mental development necessarily passes through a stage in which he has some ideas, but yet has never recognised that an idea may be erroneous; and a stage that every child necessarily passes through must have been formerly passed through by the race in its adult development. It may be doubted whether many of the lower animals have any clear and steady conception of falsehood; for their instincts work so unerringly that there is little to force it upon their attention. Yet plainly without a knowledge of falsehood no development of discursive reason can take place.

489. This paradisaical logic appears in the study of nonrelative formal logic. But there no possible avenue appears by which the knowledge of falsehood could be brought into this Garden of Eden except by the arbitrary and inexplicable introduction of the Serpent in the guise of a proposition necessarily false. The logic of relatives, affords such an avenue, and that, the very avenue by which in actual development, this stage of logic supervenes. It is the avenue of experience and logical reflexion.

490. By logical reflexion, I mean the observation of thoughts in their expressions. Aquinas remarked that this sort of reflexion is requisite to furnish us with those ideas which, from lack of contrast, ordinary external experience fails to bring into prominence. He called such ideas second intentions. It is by means of relatives of second intention that the general method of logical representation is to find completion.

491. Let inline image signify that "— is inline image †1 Then Figure 22 means

inline image inline image inline image
Figure 22 Figure 23 Figure 24

that taking any two things whatever, either the one is neither itself nor the other (putting it out of the question as an absurdity), or the other is a non-giver of something to that thing. That is, nothing gives all things, each to itself. Thus, the existence of any general description of thing can be denied.

inline image

Figure 25

Either medad of Figure 23 means no wise men are virtuous. Figure 24 is equivalent to Figure 7. Figure 25 means "each wise man is a lover of something virtuous." Thus we see that this mode of junction — lover of some virtuous — which seems so simple — is really complex. Figure 26 means "some one

inline image

Figure 26

thing is loved by all wise men." Figure 27 means that every man is either wise or virtuous. Figure 28 means that every man is both wise and virtuous.

inline image inline image
Figure 27 Figure 28

These explanations need not be carried further to show that we have here a perfectly efficient and highly analytical method of representing relations.

§7. The Algebra of Dyadic Relatives

492. Although the primitive relatives are triadic, yet they may be represented with but little violence by means of dyadic relatives, provided we allow several attachments to one blank. For instance, A gives B to C, may be represented by saying A is the first party in the transaction D, B is subject of D, C is second party of D, D is a giving by the first party of the subject to the second party. Triadic relatives cannot conveniently be represented on one line of writing. These considerations led me to invent the algebra of dyadic relatives as a tolerably convenient substitute in many cases for the graphical method of representation. In place of the one "operation," or mode of conjunction of graphical method, there are in this algebra four operations.

493. For the purpose of this algebra, I entirely discard the idea that every compound relative consists of an antecedent and a consequent part. I consider the circle round the antecedent as a mere sign of negation, for which in the algebra I substitute an obelus over that antecedent. The line between antecedent and consequent, I treat as a sign of an "operation" by itself. It signifies that anything whatever being taken as correlate of the first written member — antecedent or consequent — and as first relate of the second written member, either the one or the other is to be accepted. Thus in place of the relative of Figure 29 signifying that "taking anything whatever, M, either — is not a lover of M, or M is a benefactor of —," that is "— is a lover only of a benefactor of —," I write

inline image b.

Or if it happens to be read the other way, putting a short mark over any letters to signify that relate and correlate are interchanged, I write the same thing

inline image $

This operation, which may, at need, be denoted by a dagger in print, to which I give a scorpion-tail curve in its cursive form, I call relative addition. †1

494. The relative "— stands to everything which is a benefactor of — in the relation of servant of every lover of his," shows,

inline image inline image
Figure 29 Figure 30

as written in Figure 30, an unencircled bond between s and l. The junction of the l and the b may therefore be regarded as direct. Stating the relative so as to make this direct junction prominent, it is "— is servant of everything that is a lover of a benefactor of —." In the algebra, as far as already explained, "lover of a benefactor" would be written

~( inline image )

that is, not a non-lover of every benefactor, or not a lover only of non-benefactors. This mode of junction, I call, in the algebra, the operation of relative multiplication, and write it

lb.

We have, then, the purely formal, or meaningless, equation

lb = ~( inline image ).

And in like manner, as a consequence of this,

l inline image b = ~().

That is to say, "To say that A is a lover of everything but benefactors of B," or "A is a non-lover only of benefactors of B," is the same as to say that A is not a non-lover of a non-benefactor of B.

495. To express in the algebra the relative of Figure 31.

inline image

Figure 31

or "— is both a lover and a benefactor of —," I write

l · b,

calling this "the operation of non-relative multiplication." To express "— is either a lover or a benefactor of —," which might be written

~(l · b),

I write

l inline image b,

calling this the operation of non-relative addition, or more accurately, of aggregation. These last two operations belong to the Boolian algebra of non-relative logic. They are De Morgan's operations of composition and aggregation. Boole himself did not use the last, but in place of it an operation more properly termed addition which gives no interpretable result when the aggregants have any common aggregant. Mr. Venn †1 still holds out for Boole's operation, and there are weighty considerations in its favor. In my opinion, the decision between the two operations should depend upon whether the quantified predicate is rejected (when aggregation should be used), or accepted (when Boole's strict addition should be used).

496. The use of these four operations necessitates continual resort to parentheses, brackets, and braces to show how far the different compound relatives extend. It also becomes desirable to have a "copula of inclusion," or the sign of "is exclusively (if anything)." For this purpose I have since 1870 †1 employed the sign ⤙ (intended for an improved ≦). It is easily made in the composing room from a dash followed by <, and in its cursive form is struck off in two rapid strokes, thus inline image. Its meaning is exemplified in the formula

w inline image v

"anybody who is wise (if any there be) is exclusively found among the virtuous." We also require in this algebra the signs of relatives of second intention

0, "— is inconsistent with —," ♈ "— is coexistent with —,"
"— is other than —," | "— is identical with."

497. The algebra has a moderate amount of power in skilful hands; but its great defect is the vast multitude of purely formal propositions which it brings along. The most significant of these are

s(l inline image b) inline image sl inline image b

and

(l inline image b)s inline image l inline image bs

That is, whatever is a servant of something which is a lover of everything but benefactors is a servant-of-a-lover to everything but benefactors, etc.

498. Professor Schröder attaches, as it seems to me, too high a value to this algebra. That which is in his eyes the greatest recommendation of it is to me scarcely a merit, namely that it enables us to express in the outward guise of an equation propositions whose real meaning is much simpler than that of an equation.

§8. General Algebra of Logic

499. Besides the algebra just described, I have invented another which seems to me much more valuable. It expresses with the utmost facility everything which can be expressed by a graph, and frequently much more clearly than the unabridged graphs described above. The method of using it in the solution of special problems has also been fully developed by me.

500. In this algebra every proposition consists of two parts, its quantifiers and its Boolian. The Boolian consists of a number of relatives united by a non-relative multiplication and aggregation. No relative operations are required (though they can be introduced if desired). Each elementary relative is represented by a letter on the line of writing with subjacent indices to denote the hecceities which fill its blanks. An obelus is drawn over such a relative to deny it.

501. To the left of the Boolian are written the quantifiers. Each of these is a Π or a Σ with one of the indices written subjacent to it, to signify that in the Boolian every object in the universe is to be imaged substituted successively for that index and the non-relative product (if the quantifier is Π) or the aggregate (if the quantifier is Σ) of the results taken. The order of the quantifiers is, of course, material. Thus

ΠiΣjlij = (l11 inline image l12 inline image l13 etc.) · (l21 inline image l22 inline image l23 inline image etc.) · etc.

will mean anything loves something. But

ΣjΠilij = l11 · l21 · l31 · etc. inline image l12 · l22 · l32 · etc. inline image l23 · l23 · l33 · etc. inline image etc.

will mean something is loved by all things.

502. This algebra, which has but two operations, and those easily manageable, is, in my opinion, the most convenient apparatus for the study of difficult logical problems, although the graphical method is capable of such modification as to render it substantially as convenient on the average. Nor would I refuse to avail myself of the algebra of dyadic relatives in the simpler cases in which it is easily handled.

§9. Method of Calculating with the General Algebra

503. My rules for working this algebra, the fruit of long experience with applying it to a great variety of genuine inquiries, have never been published. †1 Nor can I here do more than state such as the beginner will be likely to require.

504. A number of premisses being given, it is required to know the most important conclusions of a certain description which can be drawn from them. The first step will be to express the premisses by means of the general algebra, taking care to use entirely different letters as indices in the different premisses.

505. These premisses are then to be copulated (or, in Whewell's phrase, colligated), i.e., non-relatively multiplied together, by multiplying their Boolians and writing before the product all the quantifiers. The relative order of the quantifiers of each premiss must (in general) be undisturbed; but the relative order of quantifiers of different premisses is arbitrary. The student ought to place Σ's as far to the left and Π's as far to the right as possible. Different arrangements of the quantifiers will lead to different conclusions from the premisses. It sometimes happens that each of several arrangements leads to a conclusion which could not easily be reached from any other arrangement.

506. The premisses, being so copulated, become one copulated premiss. This copulated premiss is next to be logically multiplied into itself any number of times, the indices being different in all the different factors. For there will be certain conclusions which I call conclusions of the first order, which can be drawn from the copulated premiss without such involution, certain others, which I call inferences of the second order, which can be drawn from its square, etc. But after involution has been carried to a certain point, higher powers will only lead to inferences of subsidiary importance. The student will get a just idea of this matter by considering the rise and decline of interest in the theorems of any mathematical theory, such as geometry or the theory of numbers, as the fundamental hypotheses are applied more and more times in the demonstrations. The number of factors in the copulated premiss, which embraces all the hypotheses that either theory assumes, is not great. Yet from this premiss many thousand conclusions have already been drawn in the case of geometry and hundreds in the case of the theory of numbers. New conclusions are now coming in faster than ever before. From the nature of logic they can never be exhausted. But as time goes on the conclusions become more special and less important. It is true that mathematics, as a whole, does not become more special nor its late discoveries less important, because there is a growth of the hypotheses. Up to a certain degree, the importance of the conclusions increases with their "order." Thus, in geometry, there is nothing worth mention of the first order, and hardly of the second. But there is a great falling off in the importance of conclusions in the theories mentioned long before the fiftieth order has been reached.

507. This involution having been performed, the next step will be the identification (occasionally the diversification) of certain indices. The rule is, that any index quantified with a Π can be transmitted, throughout the Boolian, into any other index whose quantifier stands to the left of its own, which now becomes useless, since it refers to nothing in the Boolian. For example, in

ΣiΠjlij

which in the Algebra of Dyadic Relatives would be written ♈(linline image0), we can identify j with i and write

Σilii

which in the other algebra becomes ♈(l·|)♈.

508. That done, the Boolian is to be manipulated according to any of the methods of non-relative Boolian algebra, and the conclusion is read off.

509. But it is only in the simplest cases that the above operations suffice. Relatives of second intention will often have to be introduced; and their peculiar properties must be attended to. Those of 0 and ♈ are covered by the rules of nonrelative Boolian algebra; but it is not so with | and . We have, for example, to observe that

Πixi inline image yi = ΠiΠjxi inline image ij inline image yi †1

Σixi·yi = ΣiΣjxi - |ji·yj.

Exceedingly important are the relatives signifying "— is a quality of —" and "— is a relation of — to —." It may be said that mathematical reasoning (which is the only deductive reasoning, if not absolutely, at least eminently) almost entirely turns on the consideration of abstractions as if they were objects. The protest of nominalism against such hypostatisation, although, if it knew how to formulate itself, it would be justified as against much of the empty disputation of the medieval Dunces, yet, as it was and is formulated, is simply a protest against the only kind of thinking that has ever advanced human culture. Nobody will work long with the logic of relatives — unless he restricts the problems of his studies very much — without seeing that this is true.

§10. Schröder's Coception of Logical Problems

510. Of my own labors in the logic of relatives since my last publication in 1884, †1 I intend to give a slight hint in §13. But I desire to give some idea of a part of the contents of Schröder's last volume. In doing so, I shall adhere to my own notation; for I cannot accept Professor Schröder's proposed innovations. I shall give my reasons in detail for this dissent in the Bulletin of the American Mathematical Society. †2 I will here only indicate their general nature. I have no objection whatever to the creation of a new system of signs ab ovo, if anybody can propose such a system sufficiently recommending itself. But that Professor Schröder does not attempt. He wishes his notation to have the support of existing habits and conventions, while proposing a measure of reform in the present usage. For that he must obtain general consent. Now it seems to me quite certain that no such general agreement can be obtained without the strictest deference to the principle of priority. Without that, new notations can only lead to confusion thrice confounded. The experience of biologists in regard to the nomenclature of their genera and other groups shows that this is so. I believe that their experience shows that the only way to secure uniformity in regard to conventions of this sort, is to accept for each operation and relative the sign definitively recommended by the person who introduced that operation or relative into the Boolian algebra, unless there are the most substantial reasons for dissatisfaction with the meaning of the sign. Objections of lesser magnitude may justify slight modifications of signs; as I modify Jevons's ·|· to inline image, by uniting the two dots by a connecting line, and as I so far yield to Schröder's objections to using ∞ for the sign of whatever is, as to resort to the similarly shaped sign of Aries ♈ (especially as a notation of some power is obtained by using all the signs of the Zodiac in the same sense, as I shall show elsewhere). In my opinion, Professor Schröder alleges no sufficient reason for a single one of his innovations; and I further consider them as positively objectionable.

511. The volume consists of thirty-one long sections filling six hundred and fifty pages. I can, therefore, not attempt to do more than to exemplify its contents by specimens of the work selected as particularly interesting. Professor Schröder chiefly occupies himself with what he calls "solution-problems," in which it is required to deduce from a given proposition an equation of which one member consists in a certain relative determined in advance, while the other member shall not contain that relative. He rightly remarks that such problems often involve problems of elimination.

512. While I am not at all disposed to deny that the so-called "solution-problems," consisting in the ascertainment of the general forms of relatives which satisfy given conditions, are often of considerable importance, I cannot admit that the interest of logical study centres in them. I hold that it is usually much more to the purpose to express in the simplest way what a given premiss discloses in regard to the constitution of a relative, whether that simplest expression is of the nature of an equation or not. Thus, one of Schröder's problems is, "Given xa, required x," — for instance, knowing that an opossum is a marsupial, give a description of the opossum. †1 The so-called solution is inline image(x = a · u), †2 or opossums embrace precisely what is common to marsupials and to some other class. In my judgment xa might with great propriety be called the solution of inline image(x = a · u), †2. When the information contained in a proposition is not of the nature of an equation, why should we, by circumlocutions, insist upon expressing it in the form of an equation?

513. Professor Schröder attaches great importance to the generality of solutions. In my opinion, this is a mistake. It is not merely that he insists that solutions shall be complete, as for example when we require every root of a numerical equation, but further that they shall all be embraced under one algebraical expression. Upon that he insists and with that he is satisfied. Whether or not the "solution" is such as to exhibit anything of the real constitution of the relative which forms the first member of the equation he does not seem to care; at least, there is no apparent consideration of the question of how such a result can be secured.

514. Pure mathematics always selects for the subjects of its studies manifolds of perfect homogeneity; and thence it comes that for the problems which first present themselves general solutions are possible, which notwithstanding their generality, guide us at once to all the particular solutions. But even in pure mathematics the class of problems which are capable of solutions at once general and useful is an exceedingly limited one. All others have to be treated by subdivision of cases. That is what meets us everywhere in higher algebra. As for general solutions, they are for the most part trivial — like the well-known and obvious test for a prime number that the continued product of all lesser numbers increased by 1 shall be divisible by that number. Only in those cases in which a general solution points the way to the particular solutions is it valuable; for it is only the particular solutions which picture to the mind the solution of a problem; and a form of words which fails to produce a definite picture in the mind is meaningless. †1

515. Professor Schröder endeavors to give the most general formula of a logical problem. It is in dealing with such very general and fundamental matters that the exact logician is most in danger of violating his own principles of exactitude. To seek a formula for all logical problems is to ask what it is, in general terms, that men inquire. To answer that question, my own logical proceeding would be to note that it asks what the essence of a question, in general, is. Now a question is a rational contrivance or device, and in order to understand any rational contrivance, experience shows that the best way is to begin by considering what circumstances of need prompted the contrivance, and then upon what general principle its action is designed to fill that need. Applying this general experience to the case before us, we remark that every question is prompted by some need — that is, by some unsatisfactory condition of things, and that the object of asking the question is to fill that need by bringing reason to bear upon it and to do this by a hypnotically suggestive indication of that to which the mind has to apply itself. I do not know that I have ever, before this minute, considered the question what is the most general formulation of a problem in general; for I do not find much virtue in general formulæ. Nor do I think my answer to this question affords any particularly precious suggestion. But its ordinary character makes it all the better an illustration of the manner — or one of the manners — in which an exact logician may attack, off-hand, a suddenly sprung question. A question, I say, is an indication suggestive (in the hypnotic sense) of what has to be thought about in order to satisfy some more or less pressing want. Ideas like those of this statement, and not talk about φ x, and "roots," and the like, must, in my opinion, form the staple of a logical analysis and useful description of a problem, in general. I am none the less a mathematical logician for that. If of two students of the theory of numbers one should insist upon considering numbers as expressed in a system of notation like the Arabic (though using now one number as base of the numeration, and now another), while the other student should maintain that all that was foreign to the theory of numbers, which ought not to consider upon what system the numbers with which it deals are expressed, those two students would, to my apprehension, occupy positions analogous to that of Schröder and mine in regard to this matter of the formulation of the problems of logic; and supposing the student who wished to consider the forms of expression of numbers were to accuse the other of being wanting in the spirit of an arithmetician, that charge would be unjust in quite the same way in which it would be unjust to charge me with deficiency in the mathematical spirit on account of my regarding the conceptions of "values," and "roots," and all that as very special ideas, which can only lumber up the field of consciousness with such hindrances as it is the very end and aim of that diagrammatic method of thinking that characterises the mathematician to get rid of.

516. But different questions are so very unlike that the only way to get much idea of the nature of a problem is to consider the different cases separately. There are in the first place questions about needs and their fulfillment which are not directly affected by the asking of the questions. A very good example is a chess problem. You have only to experiment in the imagination just as you would do on the board if it were permitted to touch the men, and if your experiments are intelligently conducted and are carried far enough, the solution required must be discovered. In other cases, the need to which the question relates is nothing but the intellectual need of having that question answered. It may happen that questions of this kind can likewise be answered by imaginary experimentation; but the more usual case requires real experimentation. The need is of one or other of two kinds. In the one class of cases we experience on several occasions to which our own deliberate action gave a common character, an excitation of one and the same novel idea or sensation, and the need is that a large number of propositions having the same novel consequent but different antecedents, should be replaced by one proposition which brings in the novel element, so that the others shall appear as mere consequences of everyday facts with a single novel one. We may express this intellectual need in a brief phrase as the need of synthetising a multitude of subjects. It is the need of generalisation. In another class of cases, we find in some new thing, or new situation, a great number of characters, the same as would naturally present themselves as consequences of a hypothetical state of things, and the need is that the large number of novel propositions with one subject or antecedent should be replaced by a single novel proposition, namely that the new thing or new occasion belongs to the hypothetical class, from which all those other novelties shall follow as mere consequences of matters of course. This intellectual need, briefly stated, is the need of synthetising a multitude of predicates. It is the need of theory. Every problem, then, is either a problem of consequences, a problem of generalisation, or a problem of theory. †1 This statement illustrates how special solutions are the only ones which directly mean anything or embody any knowledge; and general solutions are only useful when they happen to suggest what the special solutions will be.

517. Professor Schröder entertains very different ideas upon these matters. The general problem, according to him, †1 is, "Given the proposition Fx = 0, required the 'value' of x0," that is, an expression not containing x which can be equated to x. This "value" must be the "general root," that is, it must, under one general description, cover every possible object which fulfils a given condition. This, by the way, is the simplest explanation of what Schröder means by a 'solution-problem"; it is the problem to find that form of relative which necessarily fulfils a given condition and in which every relative that fulfils that condition can be expressed. Schröder shows that the solution of such a problem can be put into the form (inline imagex = fu), which means that a suitable logical function (f) of any relative, u, no matter what, will satisfy the condition Fx = 0; and that nothing which is not equivalent to such a function will satisfy that condition. He further shows, what is very significant, that the solution may be required to satisfy the "adventitious condition" fx = x. †2 This fact about the adventitious condition is all that prevents me from rating the value of the whole discussion as far from high.

518. Professor Schröder next produces what he calls "the rigorous solution" of the general question. This promises something very fine — the rigorously correct resolution of everything that ever could (but for this knowledge) puzzle the human mind. It is true that it supposes that a particular relative has been found which shall satisfy the condition Fx = 0. But that is seldom difficult to find. Either 0, or ♈, or some other trivial solution commonly offers itself. Supposing, then, that a be this particular solution, that is, that Fa = 0, the "rigorous solution" is

x = fu = a · ♈(Fu)♈ inline image u ·(0 inline image ~(Fu)0). †3

That is, it is such a function of u that when u satisfies the condition Fu = 0, fu = u; but when u does not satisfy this condition fu = a. Now Fa = 0.

519. Since Professor Schröder carries his algebraicity so very far, and talks of "roots," "values," "solutions," etc., when, even in my opinion, with my bias towards algebra, such phrases are out of place, let us see how this "rigorous solution" would stand the climate of numerical algebra. What should we say of a man who professed to give rigorous general solutions of algebraic equations of every degree (a problem included, of course, under Professor Schröder's general problem)? Take the equation: x5 + Ax4 + Bx3 + Cx2 + Dx + E = 0. Multiplying by x - a we get

x6 + (A-a)x5 + (B-aA)x4 + (C-aB)x3 + (D-aC)x2 + (E-aD) x-aE = 0

The roots of this equation are precisely the same as those of the proposed quintic together with the additional root x = a. Hence, if we solve the sextic we thereby solve the quintic. Now our Schröderian solver would say, "There is a certain function, fu, every value of which, no matter what be the value of the variable, is a root of the sextic. And this function is formed by a direct operation. Namely, for all values of u which satisfy the equation

u6 + (A-a)u5 + (B-aA)u4 + (C-aB)u3 + (D-aC)u2 + (E-aD) u-aE = 0

fu = u, while for all other values, fu = a. Then, x = fu is the expression of every root of the sextic and of nothing else. It is safe to say that Professor Schröder would pronounce a pretender to algebraical power who should talk in that fashion to be a proper subject for surveillance if not for confinement in an asylum. Yet he would only be applying Professor Schröder's "rigorous solution," neither more nor less. It is true that Schröder considers this solution as somewhat unsatisfactory; but he fails to state any principle according to which it should be so. Nor does he hold it too unsatisfactory to be frequently resorted to in the course of the volume. The invention of this solution exhibits in a high degree that very effective ingenuity which the solution itself so utterly lacks, owing to its resting on no correct conception of the nature of problems in general and of their solutions and of the meaning of a proposition.

§11. Professor Schröder's Pentagrammatical Notation

520. Professor Schröder's greatest success in the logic of relatives, is due precisely to his having, in regard to certain questions, proceeded by the separation of cases, quite abandoning the glittering generalities of the algebra of dyadic relatives. As his greatest success, I reckon his solutions of "inverse row and column problems" in §16, resting upon an investigation in §15 of the relations of various compound relatives which end in 0, ♈, |, and . The investigations of §15 might perfectly well have been carried through without any other instrument than the algebra of dyadic relatives. This course would have had certain advantages, such as that of exhibiting the principles on which the formulæ rest. But directness of proof would not have been of the number of those advantages; this is on the contrary decidedly with the notation invented and used by Professor Schröder. This notation may be called pentagrammatic, since it denotes a relative by a row of five characters. Imagine a list to be made of all the objects in the universe. Second, imagine a switchboard, consisting of a horizontal strip of brass for each object (these strips being fastened on a wall at a little distance one over another according to the order of the objects in the list) together with a vertical strip of brass for each object (these strips being fastened a little forward of the others, and being arranged in the same order), with holes at all the intersections, so that when a brass plug is inserted in any hole, the object corresponding to the horizontal brass strip can act in some way upon the object corresponding to the vertical brass strip. In order then, by means of this switchboard, to get an analogue of any dyadic relative, a lover of —," we insert plugs so that A and B, being any two objects, A can act on B, if and only if A is a lover of B. Now in Professor Schröder's pentagrammatic notation, the first of the five characters denoting any logical function of a primitive relative, a, refers to those horizontal strips, all whose holes are plugged in the representation of a (or, as we may say for short, "in a"), the second refers to those horizontal strips, each of which has in a every hole plugged but one. This one, not necessarily the same for all such strips, may be denoted by A. The third character refers to those horizontal strips which in a have several holes plugged, and several empty. The full holes (different, it may be, in the different horizontal strips) may be denoted by β. The fourth character refers to those horizontal strips which in a have, each of them, but one hole plugged, generally a different hole in each. This one plugged hole may be denoted by Γ. The fifth character will refer to those rows each of which in a has all its holes empty. Then, a will be denoted by ♈ĀβΓ0; and ā by 0Aβ̅Γ̅♈; †1 for in ā, all the holes must be filled that are void in a, and vice versa. Consequently ā = 0Ā♈♈♈. This shall be shown as soon as we have first examined the pentagrammatic symbol for a. This symbol divides a into four aggregants, viz.:

a = (ainline image0)inline imagea·[(ainline image|)·āinline imagea · a · (ā · ā)inline imagea · (āinline image|)

In order to prove, by the algebra itself that this equation holds, we remark that a = a · binline imagea · , whatever b may be. For b, substitute (ainline image0) . Then, ainline image0 inline image a inline image ; but ainline image = a. Hence, a · b = ainline image0. a · = a · ā♈ = a · ā(| inline image ) = a · (ā|inline imageā ). But ā| = ā, and a · ā = 0. Hence a · = a · ā . Thus a = ainline image0 inline image a · ā . Now, in ā = ā · c inline image ā · , substitute for c, ainline image |. This gives ā = (ainline image |) · ā inline image ā · ā; and thus, a = ainline image0 inline image a · (ainline image |) · ā inline image a · (ā · ā) . Finally, a = a · a inline image a · (ā inline image |). But a · (ā inline image |) = a · (ā inline image |) · (ā · ā) inline image a · (ā inline image |) · {[(ainline image |) inline image a] inline image |}.

And

a · (ā inline image |) · {[(ainline image |) ainline image |}
= a · {ā · [(ainline image |) inline image a] inline image |} (by distribution)
= a · [ā · (ainline image |) inline image | (since ā · a = 0)
= a · (ā inline image |) · (ainline image | inline image |) (by distribution)
= a · (ā inline image |) · (ainline image0) (if more than 2 things exist)
= a · (ā inline image |) · (ainline image | · ) (since 0 = 1 · )
= a · (ā inline image |) · (ainline image |) · (ainline image ) (by distribution)
= a · (ā inline image |) · (ainline image |) (since ainline image = a)
= a · (ā · ainline image |) (by distribution)
= a · (0 inline image |) (since ā · a = 0)
= a · 0 (if more than 1 object exists)
= 0.

So that a · (ā inline image |) = a · (ā inline image |) · (ā · ā) and thus

a = ainline image0 inline image a · [(ainline image |) · āTinline imagea · a (ā ·ā)Tinline imagea · (ā inline image |) This is the meaning of the symbol ♈ĀbG0.

521. We, now, at length, return, as promised to the examination of ā . First, ainline image0 inline image ~(ā ) inline image0. For ~(ā ) = ainline image | and ainline image | inline image0 = ainline image (| inline image0) = ainline image0. Hence the first character in the pentagrammatic symbol for ā must be 0. Second a · [(a inline image |) · ā] inline image ā · [(āinline image|) · ~(ā )] . For it is plain that a · [(ainline image |) · ā] inline image [(ainline image |) · ā] inline image ā . Also ā inline image āinline image ā( inline image |) inline image āinline image|. Hence [(ainline image |) · ā] inline image [(ainline image|) · (āinline image|)] . But ainline image| = ~(ā ). Hence, a · [(ainline image|) · ā inline image ā · [(āinline image|) · ~(ā )] . Hence, the second character in the pentagrammatic sign for ā , is the same as that of a. Thirdly a · a · (ā · ā) inline image āinline image0. For ā inline image ā| inline image ā( inline image|) inline image (āinline image|). Hence (ā · ā ) inline image [(āinline image|) · (āinline image )] inline image (āinline image| · ) inline image(āinline image0) inline image āinline image0 inline image āinline image0. Consequently, the third character of the pentagrammatic symbol of ā must be ♈. Fourthly, a · (āinline image|) inline image āinline image0. For we have just seen that ā inline image āinline image|. Hence āinline image| inline image āinline image|inline image|. But |inline image| = 0 if there is more than one object in the universe. Hence ā inline image | inline image āinline image0. Consequently, the fourth character of the pentagrammatic formula for ā is ♈. Finally, āinline image0 inline image āinline image0. For āinline image0 inline image āinline image0inline image0 inline image āinline image| · inline image0 inline image (āinline image|) · (āinline image )inline image0 inline image āinline image|inline image0 inline image āinline image0. Hence the fifth character of the pentagram of ā is ♈. In fine, that pentagram is 0Ā ♈ ♈ ♈. Professor Schröder obtains this result more directly by means of a special calculus of the pentagrammatic notation. In that way, he obtains, in §15, a vast number of formulæ, which in §16 are applied in the first place with great success to the solution of such problems as this: Required a form of relation in which everything stands to something but nothing to everything. The author finds instantaneously that every relative signifying such a relation must be reducible to the form ū· u inline image | · (u inline image0 inline image ū inline image0). †1 In fact, the first term of this expression ū· u, for which ū· u ♈ might as well be written, embraces all the relatives in question. For let ū be any such relative. Then, u = ū· u. The second term is added, curiously enough, merely to exclude other relations. For if u is such a relative that something is u to everything or to nothing, then that something would be in the relation ū· u to nothing. To give it a correlate the second term is added; and since all the relatives are already included, it matters not what that correlate be, so long as the second term does not exclude any of the required relatives which are included under the first term. Let v be any relative of the kind required, then v · (uinline image0 inline image ūinline image0) will answer for the second term. If we had no letter expressing a relation known to be of the required kind, the problem would be impossible. Fortunately, both | and are of that kind. Of course, the negative of such a relative is itself such a relative; so that

(uinline image0 ūinline image0) · (v inline image u· ū ♈)

would be an equivalent form, equally with

(uinline image0 inline image ūinline image0) · v inline image u . ♈ ū ♈.

522. §16 concludes with some examples of eliminations of great apparent complexity. †1 In the first of these we have given x = (ūinline image|)♈ inline image u; and it is required to eliminate u. We have, however, instantly u inline image x

(ū inline image |)♈ inline image x

Whence, immediately,

( inline image |)♈ inline image x,

or

inline image (x · x )♈.

The next example, the most complicated, requires u to be eliminated from the equation

x = ūinline image0inline image(uinline image|)♈ · ūΤinline image(uinline image|) · ūinline image(ūinline image|) · u inline image(u · ū inline image0) · ū,

He performs the elimination by means of the pentagrammatic notation very easily as follows: Putting u = ♈ĀβG0

ūinline image0 = 0000♈
(uinline image|)♈ · ū = 0Ā000
(uinline image|) · ū = 0A000
(ūinline image|) · u = 000Γ0
(u · ~u inline image0) ·u = 00β̅00
sum 0♈β̅Γ♈

Thus, x is of the form ♈β̅Γ0, which has been found in former problems to imply xinline image| inline image x.

Without the pentagrammatical notation this elimination would prove troublesome, although with that as a guide it could easily be obtained by the algebra alone.

§12. Professor Schröder's Iconic Solution of x inline image φx.

523. Another valuable result obtained by Professor Schröder is the solutions of the problem

x inline image φx.

Namely, he shows that

x = fu

where

fu = u· φu

(Of course, by contraposition, this gives for the solution of φx inline image x, x = fu where fu = u inline image φu.) The correctness of this solution will appear upon a moment's reflexion; and nearly all the useful solutions in the volume are cases under this.

524. It happens very frequently that the iteration of the functional operation is unnecessary, because it has no effect.

Suppose, for example, that we desire the general form of a "transitive" relative, that is, such a one, x, that

xx inline image x.

In this case, since | inline image l inline image $ whatever l may be, we have

x inline image x| inline image x(x inline image $) inline image xx inline image $ inline image x inline image $,

or

x inline image x inline image $

If, then,

fu = u · (u inline image $ū)

we have

x = fu.

Here,

fu inline image u;

so that

fu inline image fu.

Also,

f2u = fu · (fu inline image $~(fu) = u · (u inline image $ū) · [u · (u inline image $ū) inline image ($ū inline image u $ū)]
= u · (u inline image $ū) · [uf(│inline imageu)$ū] · [u inline image $ū inline imageinline imageu)$ū]

Now

fu = u · (u inline image $ū) = u · (u inline image $ū) · (u inline image $ū) · (u inline image $ū) = u · (u inline image $ū) · (u inline image |$ū) · (u inline image |$ū)
image
u · (u inline image $ū) · [uinline image(|inline imageu)$ū] · [uinline image($ū inline image u)$ū] inline image
image
u · (u inline image $ū) · [uinline image(|inline imageu)$ū] · (u inline image $ū inline image u$ū)
image
u · (u inline image $ū) · [uinline image(|inline imageu)$ū] · [u inline image $ū inline image (|inline imageu)$ū] inline image f2u

Thus fu = f¥u; and

x = inline imageu · (u inline image $ū)

This is a truly iconic result; that is, it shows us what the constitution of a transitive relative really is. It shows us that transitiveness always depends upon inclusion; for to say that A is l inline image $ of B is to say that the things loved by B are included among those loved by A. †1 The factor u inline image $ū is transitive by itself; for

(uinline image$ū)(uinline image$ū)inline imageuinline image$ūuinline image$ūinline imageuinline imageinline image$ūinline imageuinline image$ū.

The effect of the other factor, u, of the form for the general transitive is merely in certain cases to exclude universal identity, and thus to extend the class of relatives represented by u inline image $ū so as to include those of which it is not true that | inline image x. Here we have an instance of restriction having the effect of extension, that is, restriction of special relatives extends the class of relatives represented. This does not take place in all cases, but only where certain relatives can be represented in more than one way.

525. Indicating, for a moment, the copula by a dash, the typical and fundamental syllogism is

A - B B - C
∴ A - C.

That is to say, the principle of this syllogism enters into every syllogism. But to say that this is a valid syllogism is merely to say that the copula expresses a transitive relation. Hence, when we now find that transitiveness always depends upon inclusion, the initial analysis by which the copula of inclusion was taken as the general one is fully confirmed. For the chief end of formal logic is the representation of the syllogism.

§13. Introduction to the Logic of Quantity †2

526. The great importance of the idea of quantity in demonstrative reasoning seems to me not yet sufficiently explained. It appears, however, to be connected with the circumstance that the relations of being greater than and of being at least as great as are transitive relations. Still, a satisfactory evolutionary logic of mathematics remains a desideratum. I intend to take up that problem in a future paper. †1 Meantime the development of projective geometry and of geometrical topics has shown that there are at least two large mathematical theories of continuity into which the idea of continuous quantity, in the usual sense of that word, does not enter at all. For projective geometry Schubert †2 has developed an algebraical calculus which has a most remarkable affinity to the Boolian algebra of logic. It is, however, imperfect, in that it only gives imaginary points, rays, and planes, without deciding whether they are real or not. This defect cannot be remedied until topology — or, as I prefer to call it, mathematical topics — has been further developed and its logic accurately analysed. †3 To do this ought to be one of the first tasks of exact logicians. But before that can be accomplished, a perfectly satisfactory logical account of the conception of continuity is required. This involves the definition of a certain kind of infinity; and in order to make that quite clear, it is requisite to begin by developing the logical doctrine of infinite multitude. This doctrine still remains, after the works of Cantor, Dedekind, and others, in an inchoate condition. For example, such a question remains unanswered as the following: Is it, or is it not, logically possible for two collections to be so multitudinous that neither can be put into a one-to-one correspondence with a part or the whole of the other? To resolve this problem demands, not a mere application of logic, but a further development of the conception of logical possibility.

527. †4 I formerly defined the possible as that which in a given state of information (real or feigned) we do not know not to be true. †5 But this definition today seems to me only a twisted phrase which, by means of two negatives, conceals an anacoluthon. We know in advance of experience that certain things are not true, because we see they are impossible. Thus, if a chemist tests the contents of a hundred bottles for fluorine, and finds it present in the majority, and if another chemist tests them for oxygen and finds it in the majority, and if each of them reports his result to me, it will be useless for them to come to me together and say that they know infallibly that fluorine and oxygen cannot be present in the same bottle; for I see that such infallibility is impossible. I know it is not true, because I satisfy myself that there is no room for it even in that ideal world of which the real world is but a fragment. I need no sensible experimentation, because ideal experimentation establishes a much broader answer to the question than sensible experimentation could give. It has come about through the agencies of development that man is endowed with intelligence of such a nature that he can by ideal experiments ascertain that in a certain universe of logical possibility certain combinations occur while others do not occur. Of those which occur in the ideal world some do and some do not occur in the real world; but all that occur in the real world occur also in the ideal world. †P1 For the real world is the world of sensible experience, and it is a part of the process of sensible experience to locate its facts in the world of ideas. This is what I mean by saying that the sensible world is but a fragment of the ideal world. In respect to the ideal world we are virtually omniscient; that is to say, there is nothing but lack of time, of perseverance, and of activity of mind to prevent our making the requisite experiments to ascertain positively whether a given combination occurs or not. Thus, every proposition about the ideal world can be ascertained to be either true or false. A description of thing which occurs in that world is possible, in the substantive logical sense. Very many writers assert that everything is logically possible which involves no contradiction. Let us call that sort of logical possibility, essential, or formal, logical possibility. It is not the only logical possibility; for in this sense, two propositions contradictory of one another may both be severally possible, although their combination is not possible. †P2 But in the substantive sense, the contradictory of a possible proposition is impossible, because we are virtually omniscient in regard to the ideal world. For example, there is no contradiction in supposing that only four, or any other number, of independent atoms exist. But it is made clear to us by ideal experimentation, that five atoms are to be found in the ideal world. Whether all five are to be found in the sensible world or not, to say that there are only four in the ideal world is a proposition absolutely to be rejected, notwithstanding its involving no contradiction.

528. It would be a great mistake to suppose that ideal experimentation can be performed without danger of error; but by the exercise of care and industry this danger may be reduced indefinitely. In sensible experimentation, no care can always avoid error. The results of induction from sensible experimentation are to afford some ratio of frequency with which a given consequence follows given conditions in the existing order of experience. In induction from ideal experimentation, no particular order of experience is forced upon us; and consequently no such numerical ratio is deducible. We are confined to a dichotomy: the result either is that some description of thing occurs or that it does not occur. For example, we cannot say that one number in every three is divisible by three and one in every five is divisible by five. This is, indeed, so if we choose to arrange the numbers in the order of counting; but if we arrange them with reference to their prime factors, just as many are divisible by one prime as by another. I mean, for instance, when they are arranged [in blocks] as follows:

1,2,4,8,etc. 5,10,20,40,etc. 7,14,28,56,etc. 35,70,etc.
3,6,12,24,etc. 15,30,60,120,etc. 21,42,84,168,etc. 105,210,etc.
9,18,36,72,etc. 45,90,180,360,etc. etc. etc. †P1
27,54,108,216,etc. 135,270,540,1080,etc.
etc. etc.

529. Thus, dichotomy rules the ideal world. Plato, therefore, for whom that world alone was real, showed that insight into concepts but dimly apprehended that has always characterised philosophers of the first order, in holding dichotomy to be the only truthful mode of division. Lofty moral sense consists in regarding, not indeed the, but yet an ideal world as in some sense the only real one; and hence it is that stern moralists are always inclined to dual distinctions. †1

530. Ideal experimentation has one or other of two forms of results. It either proves that Σimi, a particular proposition true of the ideal world, and going on, finds Σjj also true; that is, that m and are both possible, or it succeeds in its induction and shows the universal proposition Πii to be true of the ideal world; that is that is necessary and m impossible.

531. Every result of an ideal induction clothes itself, in our modes of thinking, in the dress of a contradiction. It is an anacoluthon to say that a proposition is impossible because it is self-contradictory. It rather is thought so as to appear self-contradictory because the ideal induction has shown it to be impossible. But the result is that in the absence of any interfering contradiction every particular proposition is possible in the substantive logical sense, and its contradictory universal proposition is impossible. But where contradiction interferes this is reversed.

532. In former publications †2 I have given the appellation of universal or particular to a proposition according as its first quantifier is Π or Σ. But the study of substantive logical possibility has led me to substitute the appellations negative and affirmative in this sense, and to call a proposition universal or particular according as its last quantifier is Π or Σ †3. For letting l be any relative, one or other of the two propositions

ΠiΣjlij ΣiΠjij

and one or other of the two propositions

ΠjΣiij ΣjΠilij

are true, while the other one of each pair is false. Now, in the absence of any peculiar property of the special relative l, the two similar forms ΣiΠiij †4 and ΣjΠilij must be equally possible in the substantive logical sense. But these two propositions cannot both be true. Hence, both must be false in the ideal world, in the absence of any constraining contradiction. Accordingly, these ought to be regarded as universal propositions, and their contradictions, ΠiΣjlij and ΠjΣiji, as particular propositions. †1

533. There are two opposite points of view, each having its logical value, from one of which, of two quantifiers of the same proposition, the preceding is more important than the following, while from the other point of view the reverse is the case. Accordingly, we may say that an affirmative proposition is particular in a secondary way, and that a particular proposition is affirmative in a secondary way.

534. If an index is not quantified at all, the proposition is, with reference to that index, singular. To ascertain whether or not such a proposition is true of the ideal world, it must be shown to depend upon some universal or particular proposition.

535. If some of the quantifiers refer not to hecceities, having in themselves no general characters except the logical characters of identity, diversity, etc., but refer to characters, whether non-relative or relative, these alone are to be considered in determining the "quantity" of an ideal proposition as universal or particular. For anything whatever is true of some character, unless that proposition be downright absurd; while nothing is true of all characters except what is formally necessary. †2 Consider, for example, a dyadic relation. This is nothing but an aggregation of pairs. Now any two hecceities may in either order form a pair; and any aggregate whatever of such pairs will form some dyadic relation. Hence, we may totally disregard the manner in which the hecceities are connected in determining the possibility of a hypothesis about some dyadic relation.

536. Characters have themselves characters, such as importance, obviousness, complexity, and the like. If some of the quantified indices denote such characters of characters, they will, in reference to a purely ideal world be paramount in determining the quantity of the proposition as universal or particular.

537. All quantitative comparison depends upon a correspondence. A correspondence is a relation which every subject †P1 of one collection bears to a subject of another collection, to which no other is in the same relation. That is to say, the relative "corresponds to" has

inline imageu · (| inline image ū)

not merely as its form, but as its definition. This relative is transitive; for its relative product into itself is

[inline imageu · (|inline imageū)] [inline imagev · (|inline image)]inline image inline image inline imageuv · (|inline imageū)(}inline image)

inline image inline image inline imageuv · (|inline imageūinline image)inline image inline image inline imageuv · (|inline image~(uv)inline image inline imagew · (|inline image).

But it is to be observed that if the P's, the Q's, and the R's are three collections, it does not follow because every P corresponds to an R, and every Q corresponds to an R that every object of the aggregate collection Pinline imageQ corresponds to an R. The dictum de omni in external appearance fails here. For P may be u·(|inline imageū)R and Q may be [v·(|inline image)]R; but the aggregate of these is not [(uinline imagev) · (|inline image~(uinline imagev))]R, which equals [(uinline imagev)·(|inline imageū)·(|inline image)]R. The aggregate of the two first is {(uinline imagev)·[v·(|inline image)inline image|inline imageū·[u·(|inline imageū)inline image|inline image]}R, which is obviously too broad to be necessarily included under the other expression. Correspondence is, therefore, not a relation between the subjects of one collection and those of another, but between the collections themselves. Let qαi/emph> mean that i is a subject of the collection, α, and let rβjk mean that j stands in the relation β to k. Then, to say that the collection P corresponds to the collection Q, or, as it is sometimes expressed, that "for every subject of Q there is a subject of P," is to make the assertion expressed by

ΣβΠiΣjΠkPi inline image rβij·(|ikinline imageβkj)·qQj. †1

In the algebra of dual relatives this may be written

ΣβP inline image inline image[rβ · (|inline imageβ)]Q.

The transitivity is evident; for

ΣβΣγinline image[rβ·(|inline imageβ)]{inline image[rγ·(|inline imageγ)]}
inline image ΣβΣγinline image[rβ·(|inline imageβ)]{inline image[rγ·(|inline imageγ)]}
inline image ΣβΣγinline image[rβ·(|inline imageβ)]{ inline image[rγ·(|inline imageγ)]}
inline image ΣβΣγinline image[rβ·(|inline imageβ)][rγ·(|inline imageγ)]
inline image ΣβΣγinline image[rβrγ·(|inline imageβinline image~rγ)]
inline image Σδinline image[rδ·(|inline imageδ)]. †P1

538. Not only is the relative of correspondence transitive but it also possesses what may be called antithetic transitivity. Namely, if c be the relative, not only is cc inline image c but also c inline image c inline image c. To demonstrate this very important proposition is, however, far from easy. The quantifiers of the assertion that for every subject of one character there is a subject of another are ΣβΠiΣjΠk. Hence, the proposition is particular †2 and will be true in the ideal world, except in case a positive contradiction is involved.

539. Let us see how such contradiction can arise. The assertion that for every subject of P there is a subject of Q is

ΣβΠiΣjΠkPi inline image rβij·(|ikinline imageβkj)·qQj.

This cannot vanish if the first aggregant term does not vanish, that is, if ΠiqPi or there is no subject of P. It cannot vanish if everything is a subject of Q. For in that case, the last factor of the latter aggregant disappears, and substituting | for rβ the second aggregant becomes ♈. The expression cannot vanish if every subject of P is a subject of Q. For when | is substituted for rβ, we get

ΠiPi inline image qQi.

If P has but a single individual subject and Q has a subject, for every P there is a Q. For in this case we have only to take for β the relation of the subject of P to any one of the subjects of Q. But if P has more than one subject, and Q has but one, the expression above vanishes. For let 1 and 2 be the two subjects of P. Substituting 1 for i, we get

Πkrβ1j·(|1k inline image βkj) · qQj

Substituting 2 for i we get

Πkrβ2j·(|2k inline image βkj) · qQj

Multiplying these

ΠkΠkrβ1j·rβ2j·(|1k inline image βkj)·(|2k' inline image βk'j)· qQi

Substituting 2 for k and 1 for k', this gives

rβ1j·rβ2j·β2j·β1j·qQi

which involves two contradictions.

540. It is to be remarked that although if every subject of P is a subject of Q, then for every subject of P there is a subject of Q, yet it does not follow that if the subjects of P are a part only of the subjects of Q, that there is then not a subject of P for every subject of Q. For example, numbering 2, 4, 6, etc., as the first, second, third, etc., of the even numbers, there is an even number for every whole number, although the even numbers form but a part of the whole numbers.

541. It is now requisite, in order to prove that c inline image c inline image c, to draw three propositions from the doctrine of substantive logical possibility. The first is that given any relation, there is a possible relation which differs from the given relation only in excluding any of the pairs we may choose to exclude. Suppose, for instance, that for every subject of P there is a subject of Q, that is that

ΣβP inline image [rβ · (| inline image β)]Q.

The factor (| inline image β) here has the effect of allowing each correlate but one relate. Each relate is, however, allowed any number of correlates. If we exclude all but one of these, the one retained being, if possible, a subject of Q, we have a possible relation, β', such that

Σβ'P inline image [rβ' · (| inline image β') · (β'inline image|)]Q

542. The second proposition of substantive logical possibility is that whatever is true of some of a class is true of the whole of some class. That is, if we accept a proposition of the form Σiai · bi, we can write

ΣγΠi{gi}inline imageāibi,

though this will generally fail positively to assert, in itself, what is implied, that the collection γ excludes whatever is a but not b, and includes something in common with a. There are, however, cases in which this implication is easily made plain.

Applying these two principles to the relation of correspondence, we get a new statement of the assertion that for every P there is a Q. Namely, if we write aai to signify that i is a relate of the relative ra to some correlate, that is if aai = (i inline image ra♈), if we write baj to signify that inline image is a correlate of the relative ra, to some relate, that is if baj = (j inline image $ra♈), and if we write pca to signify that ra is an aggregate of the relative rc, that is, if pca = (ra inline image rc), then the proposition that for every subject of P there is a subject of Q may be put in the form,

ΣcΣγΠxΠyΣδΣεΠaΣiΣjΠβΠuΠv

[ca inline image aai · qPi · baj · qQj · qγi · (āau inline image |iu) · (a v inline image |jv) · (cβ inline image |aβ inline image āβi · bβj)] · (Px inline image aδx · pcδ) · (Qy inline image γy inline image bεy · pcε)

This states that there is a collection of pairs, c, any single pair of which, a, has for its sole first subject a subject of P, and for its sole second subject a subject of Q which is at the same time a subject of a collection, j, and that no two pairs of the collection, c, have the same first subject or the same second subject, and that every subject of P is a first subject of some pair of this collection, c, and every subject of Q which is at the same time a subject of γ is a second subject of some pair of the same collection, c.

543. The third proposition of the doctrine of substantive logical possibility of which we have need is that all hecceities are alike in respect to their capacity for entering into possible pairs. Consequently, all the objects of any collection whatever may be severally and distinctly paired with all the objects of a collection which shall either be wholly contained in, or else shall entirely contain, any other collection whatever. Consequently,

ΠPΠQΣcΣδΠxΣδΠyΣδΠaΣiΣjΠuΠvΠβΠmΠn

[ca inline image aai · qPi · baj · qδj · (āau inline image |iu) · (a v inline image |vj) · (cβ inline image |aβ inline image āβi · bβj)] · (Px inline image aδx · pcδ) · (δy inline image bεy · pcε) · (δm inline image qQm inline image Qn inline image qδn).

544. Although the above three propositions belong to a system of doctrine not universally recognised, yet I believe their truth is unquestionable. Suppose, now, that it is not true that for every subject of P there is a subject of Q. Then, in the last formula, Πmδm inline image qQm inline image 0. This leaves for the last factor ΠnQn inline image qδn, and then the formula expresses that for every subject of Q there is a subject of P. In other words, we have demonstrated the important proposition that two collections cannot be disparate in respect to correspondence, but that for every subject of the one there must be a subject of the other.

545. The theorem c inline image c%c is now established; for since of any two collections one corresponds to the other, we have ♈ inline image c inline image or (non-relatively multiplying by $) $ inline image c. Hence, c inline image |c inline image ($ inline image c)c inline image $ inline image cc inline image c inline image cc; and, by the transitive principle cc inline image c, we finally obtain c inline image c inline image c.

546. Thus is established the conception of multitude. Namely, if for every subject of P there is a subject of Q, while there is not for every subject of Q a subject of P, the multitude of Q is said to be greater that that of P. But if for every subject of each collection there is a subject of the other, the multitudes of the two collections are said to be equal the one to the other. We may create a scale of objects, one for every group of equal collections. Calling these objects arithms, the first arithm will belong to 0 considered as a collection, the second to individuals, etc. Calling a collection the counting of which can be completed an enumerable collection, the multitude of any enumerable collection equals that of the arithms that precede its arithm. Calling a collection whose multitude equals that of all the arithms of enumerable collection a denumerable collection (because its subjects can all be distinguished by ordinal numbers, though the counting of it cannot be completed), the arithms preceding the arithm of denumerable collections form a denumerable collection. More multitudinous collections are greater than the collections of arithms which precede their arithm.

547. Let there be a denumerable collection, say the cardinal numbers; and let there be two houses. Let there be a collection of children, each of whom wishes to have those numbers placed in some way into those houses, no two children wishing for the same distribution, but every distribution being wished for by some child. Then, as Dr. George Cantor has proved, †1 the collection of children is greater in multitude than the collection of numbers. Let a collection equal in multitude to that collection of children be called an abnumeral collection of the first dignity. The real numbers (surd and rational) constitute such a collection.

548. I now ask, suppose that for every way of placing the subjects of one collection in two houses, there is a way of placing the subjects of another collection in two houses, does it follow that for every subject of the former collection there is a subject of the latter? In order to answer this, I first ask whether the multitude of possible ways of placing the subjects of a collection in two houses can equal the multitude of those subjects. If so, let there be such a multitude of children. Then, each having but one wish, they can among them wish for every possible distribution of themselves among two houses. Then, however they may actually be distributed, some child will be perfectly contented. But ask each child which house he wishes himself to be in, and put every child in the house where he does not want to be. Then, no child would be content. Consequently, it is absurd to suppose that any collection can equal in multitude the possible ways of distributing its subjects in two houses.

549. Accordingly, the multitude of ways of placing a collection of objects abnumeral of the first dignity into two houses is still greater in multitude than that multitude, and may be called abnumeral of the second dignity. There will be a denumerable succession of such dignities. But there cannot be any multitude of an infinite dignity; for if there were, the multitude of ways of distributing it into two houses would be no greater than itself. †P1

550. We thus not only answer the question proposed, and show that of two unequal multitudes the multitude of ways of distributing the greater is the greater; but we obtain the entire scale of collectional quantity, which we find to consist of two equal parts (that is two parts whose multitudes of grades are equal), the one finite, the other infinite. Corresponding to the multitude of 0 on the finite scale is the abnumeral of 0 dignity, which is the denumerable, on the infinite scale, etc. †1

551. So much of the general logical doctrine of quantity has been here given, in order to illustrate the power of the logic of relatives in enabling us to treat with unerring confidence the most difficult conceptions, before which mathematicians have heretofore shrunk appalled.

552. I had been desirous of examining Professor Schröder's developments concerning individuals and individual pairs; but owing to the length this paper has already reached, I must remit that to some future occasion. †2



Paper 17: The Logic of Mathematics in Relation to Education †1

§1. Of Mathematics in General

553. In order to understand what number is, it is necessary first to acquaint ourselves with the nature of the business of mathematics in which number is employed.

554. I wish I knew with certainty the precise origin of the definition of mathematics as the science of quantity. It certainly cannot be Greek, because the Greeks were advanced in projective geometry, whose problems are such as these: whether or not four points obtained in a given way lie in one plane; whether or not four planes have a point in common; whether or not two rays (or unlimited straight lines) intersect, and the like — problems which have nothing to do with quantity, as such. Aristotle †2 names, as the subjects of mathematical study, quantity and continuity. But though he never gives a formal definition of mathematics, he makes quite clear, in more than a dozen places, his view that mathematics ought not to be defined by the things which it studies but by its peculiar mode and degree of abstractness. Precisely what he conceives this to be it would require me to go too far into the technicalities of his philosophy to explain; and I do not suppose anybody would today regard the details of his opinion as important for my purpose. Geometry, arithmetic, astronomy, and music were, in the Roman schools of the fifth century †P1 and earlier, recognized as the four branches of mathematics. And we find Boëthius (A.D. 500) defining them as the arts which relate, not to quantity, but to quantities, or quanta. What this would seem to imply is, that mathematics is the foundation of the minutely exact sciences; but really it is not worth our while, for the present purpose, to ascertain what the schoolmasters of that degenerate age conceived mathematics to be.

555. In modern times projective geometry was, until the middle of this century, almost forgotten, the extraordinary book of Desargues †P1 having been completely lost until, in 1845, Chasles came across a MS. copy of it; and, especially before imaginaries became very prominent, the definition of mathematics as the science of quantity suited well enough such mathematics as existed in the seventeenth and eighteenth centuries.

556. Kant, in the Critique of Pure Reason (Methodology, chapter I, section 1), distinctly rejects the definition of mathematics as the science of quantity. What really distinguishes mathematics, according to him, is not the subject of which it treats, but its method, which consists in studying constructions, or diagrams. That such is its method is unquestionably correct; for, even in algebra, the great purpose which the symbolism subserves is to bring a skeleton representation of the relations concerned in the problem before the mind's eye in a schematic shape, which can be studied much as a geometrical figure is studied.

557. But Rowan Hamilton and De Morgan, having a superficial acquaintance with Kant, were just enough influenced by the Critique to be led, when they found reason for rejecting the definition as the science of quantity, to conclude that mathematics was the science of pure time and pure space. Notwithstanding the profound deference which every mathematician must pay to Hamilton's opinions and my own admiration for De Morgan, I must say that it is rare to meet with a careful definition of a science so extremely objectionable as this. If Hamilton and De Morgan had attentively read what Kant himself has to say about number, in the first chapter of the Analytic of principles and elsewhere, they would have seen that it has no more to do with time and space than has every conception. Hamilton's intention probably was, by means of this definition, to throw a slur upon the introduction of imaginaries into geometry, as a false science; but what De Morgan, who was a student of multiple algebra, and whose own formal logic is plainly mathematical, could have had in view, it is hard to comprehend, unless he wished to oppose Boole's theory of logic. Not only do mathematicians study hypotheses which, both in truth and according to the Kantian epistemology, no otherwise relate to time and space than do all hypotheses whatsoever, but we now all clearly see, since the non-Euclidean geometry has become familiar to us, that there is a real science of space and a real science of time, and that these sciences are positive and experiential — branches of physics, and so not mathematical except in the sense in which thermotics and electricity are mathematical; that is, as calling in the aid of mathematics. But the gravest objection of all to the definition is that it altogether ignores the veritable characteristics of this science, as they were pointed out by Aristotle and by Kant.

558. Of late decades philosophical mathematicians have come to a pretty just understanding of the nature of their own pursuit. I do not know that anybody struck the true note before Benjamin Peirce, who, in 1870, †P1 declared mathematics to be "the science which draws necessary conclusions," adding that it must be defined "subjectively" and not "objectively." A view substantially in accord with his, though needlessly complicated, is given in the article "Mathematics," in the ninth edition of the Encyclopædia Britannica. The author, Professor George Chrystal, holds that the essence of mathematics lies in its making pure hypotheses, and in the character of the hypotheses which it makes. What the mathematicians mean by a "hypothesis" is a proposition imagined to be strictly true of an ideal state of things. In this sense, it is only about hypotheses that necessary reasoning has any application; for, in regard to the real world, we have no right to presume that any given intelligible proposition is true in absolute strictness. On the other hand, probable reasoning deals with the ordinary course of experience; now, nothing like a course of experience exists for ideal hypotheses. Hence to say that mathematics busies itself in drawing necessary conclusions, and to say that it busies itself with hypotheses, are two statements which the logician perceives come to the same thing.

559. A simple way of arriving at a true conception of the mathematician's business is to consider what service it is which he is called in to render in the course of any scientific or other inquiry. Mathematics has always been more or less a trade. An engineer, or a business company (say, an insurance company), or a buyer (say, of land), or a physicist, finds it suits his purpose to ascertain what the necessary consequences of possible facts would be; but the facts are so complicated that he cannot deal with them in his usual way. He calls upon a mathematician and states the question. Now the mathematician does not conceive it to be any part of his duty to verify the facts stated. He accepts them absolutely without question. He does not in the least care whether they are correct or not. He finds, however, in almost every case that the statement has one inconvenience, and in many cases that it has a second. The first inconvenience is that, though the statement may not at first sound very complicated, yet, when it is accurately analyzed, it is found to imply so intricate a condition of things that it far surpasses the power of the mathematician to say with exactitude what its consequence would be. At the same time, it frequently happens that the facts, as stated, are insufficient to answer the question that is put. Accordingly, the first business of the mathematician, often a most difficult task, is to frame another simpler but quite fictitious problem (supplemented, perhaps, by some supposition), which shall be within his powers, while at the same time it is sufficiently like the problem set before him to answer, well or ill, as a substitute for it. †P1 This substituted problem differs also from that which was first set before the mathematician in another respect: namely, that it is highly abstract. All features that have no bearing upon the relations of the premisses to the conclusion are effaced and obliterated. The skeletonization or diagrammatization of the problem serves more purposes than one; but its principal purpose is to strip the significant relations of all disguise. Only one kind of concrete clothing is permitted — namely, such as, whether from habit or from the constitution of the mind, has become so familiar that it decidedly aids in tracing the consequences of the hypothesis. Thus, the mathematician does two very different things: namely, he first frames a pure hypothesis stripped of all features which do not concern the drawing of consequences from it, and this he does without inquiring or caring whether it agrees with the actual facts or not; and, secondly, he proceeds to draw necessary consequences from that hypothesis.

560. Kant is entirely right in saying that, in drawing those consequences, the mathematician uses what, in geometry, is called a "construction," or in general a diagram, or visual array of characters or lines. Such a construction is formed according to a precept furnished by the hypothesis. Being formed, the construction is submitted to the scrutiny of observation, and new relations are discovered among its parts, not stated in the precept by which it was formed, and are found, by a little mental experimentation, to be such that they will always be present in such a construction. Thus, the necessary reasoning of mathematics is performed by means of observation and experiment, and its necessary character is due simply to the circumstance that the subject of this observation and experiment is a diagram of our own creation, the conditions of whose being we know all about.

But Kant, owing to the slight development which formal logic had received in his time, and especially owing to his total ignorance of the logic of relatives, which throws a brilliant light upon the whole of logic, fell into error in supposing that mathematical and philosophical necessary reasoning are distinguished by the circumstance that the former uses constructions. This is not true. All necessary reasoning whatsoever proceeds by constructions; and the only difference between mathematical and philosophical necessary deductions is that the latter are so excessively simple that the construction attracts no attention and is overlooked. The construction exists in the simplest syllogism in Barbara. Why do the logicians like to state a syllogism by writing the major premiss on one line and the minor below it, with letters substituted for the subject and predicates? It is merely because the reasoner has to notice that relation between the parts of those premisses which such a diagram brings into prominence. If the reasoner makes use of syllogistic in drawing his conclusion, he has such a diagram or construction in his mind's eye, and observes the result of eliminating the middle term. If, however, he trusts to his unaided reason, he still uses some kind of a diagram which is familiar to him personally. The true difference between the necessary logic of philosophy and mathematics is merely one of degree. It is that, in mathematics, the reasoning is frightfully intricate, while the elementary conceptions are of the last degree of familiarity; in contrast to philosophy, where the reasonings are as simple as they can be, while the elementary conceptions are abstruse and hard to get clearly apprehended. But there is another much deeper line of demarcation between the two sciences. It is that mathematics studies nothing but pure hypotheses, and is the only science which never inquires what the actual facts are; while philosophy, although it uses no microscopes or other apparatus of special observation, is really an experimental science, resting on that experience which is common to us all; so that its principal reasonings are not mathematically necessary at all, but are only necessary in the sense that all the world knows beyond all doubt those truths of experience upon which philosophy is founded. This is why the mathematician holds the reasoning of the metaphysician in supreme contempt, while he himself, when he ventures into philosophy, is apt to reason fantastically and not solidly, because he does not recognize that he is upon ground where elaborate deduction is of no more avail than it is in chemistry or biology.

561. I have thus set forth what I believe to be the prevalent opinion of philosophical mathematicians concerning the nature of their science. It will be found to be significant for the question of number. But were I to drop this branch of the subject without saying one word more, my criticism of the old definition, "mathematics is the science of quantity," would not be quite just. It must be admitted that quantity is useful in almost every branch of mathematics. Jevons wrote a book entitled Pure logic, the science of quality, which expounded, with a certain modification, the logical algebra of Boole. But it is a mistake to regard that algebra as one in which there is no system of quantity. As Boole rightly holds, there is a quadratic equation which is fundamental in it. The meaning of that equation may be expressed as follows: Every proposition has one or other of two values, being either true (which gives it one value) or false (which gives it the other). So stated, we see that the algebra of Boole is nothing but the algebra of that system of quantities which has but two values — the simplest conceivable system of quantity. The widow of the great Boole has lately written a little book †P1 in which she points out that, in solving a mathematical problem, we usually introduce some part or element into the construction which, when it has served our purpose, is removed. Of that nature is a scale of quantity, together with the apparatus by which it is transported unchanged from one part of the diagram to another, for the purpose of comparing those two parts. Something of this general description seems to be indispensable in mathematics. Take, for example, the Theorem of Pappus concerning ten rays in a plane. The demonstration of it which is now usual, that of von Staudt, introduces a third dimension; and the utility of that arises from the fact that a ray, or unlimited straight line, being the intersection of two planes, these planes show us exactly where the ray runs, while, as long as we confine ourselves to the consideration of a single plane, we have no easy method of describing precisely what the course of the ray is. Now this is not precisely a system of quantity; but it is closely analogous to such a system, and that it serves precisely the same purpose will appear when we remember that that same theorem can easily (though not so easily) be demonstrated by means of the barycentric calculus. Although, then, it is not true that all mathematics is a science of quantity, yet it is true that all mathematics makes use of a scaffolding altogether analogous to a system of quantity; and quantity itself has more or less utility in every branch of mathematics which has as yet developed into any large theory.

562. I have only to add that the hypotheses of mathematics may be divided into those general hypotheses which are adhered to throughout a whole branch of mathematics, and the particular hypotheses which are peculiar to different special problems. †1

§2. Of Pure Number

562A. The system of pure number is the general hypothesis of arithmetic — at any rate, of scientific arithmetic (or, the theory of numbers) — for whether it is best to say anything about it in vulgar arithmetic (or, the art of computing with the Arabic figures) †P1 is a question of educational theory to be considered after studying Counting and Dating.

562B. Preparatory to showing what this system is, it will be well to describe a still more general hypothesis — that of a sequence. A sequence is a multitude of objects connected with a relation, which we may call the "relation of sequence," or R, in a manner I proceed to define. It will be convenient to use this following locution: I shall say, A is R to B, meaning that A stands to B in the relation R, and with the same meaning I shall also say, B is R'd by A. Then the sequence is defined by these two precepts:

Precept I. If A is any object of the sequence whatever and B is any object of the sequence whatever, either A is not R to B or else B is not R to A.

Precept II. If A is any object of the sequence whatever, B is any such object, and C is any such object, then, so far as Precept I permits, either A is R to B, or B is R to C, or C is R to A.

Certain corollaries are easily deduced from these precepts. First, if, in applying Precept I, we choose for B the same object represented by A, that precept leads to the conclusion that no object of the sequence is R itself. This, then, shows the limitation which Precept I imposes upon Precept II. Namely, in the latter, A, B, and C cannot, all three, represent the same object. Secondly, if in Precept II we take C to be the same as B, that precept shows that of any two different objects of the sequence, A and B, either A is R to B or B is R to A. For by the first corollary, B cannot be R to C; that is, to itself. Thirdly, if A, B, and C represent objects of the sequence (the same or different), then if A is R'd by B, it follows that either B is R to C or C is R to A. For if A is R'd by B, then by Precept I, A is not R to B, and then, by Precept II, either B is R to C or C is R to A, or else A, B, and C are identical. But the last cannot be the case by the first corollary, if A is R'd by B. Fourthly, whatever objects of the sequence A, B, and C may be, if A is R'd by B and B is R'd by C, then A is R'd by C. For, by Precept I, if B is R'd by C, C is not R to B. †1 Hence, by the third corollary, if A is R'd by B and B is R'd by C, C is R to A, that is, A is R'd by C. Fifthly, whatever objects of the sequence A, B, and C may be, either A is not R'd by B or B is not R'd by C, or C is not R'd by A. For, by the fourth corollary, either A is not R'd by B, or B is not R'd by C, or A is R'd by C. But in the last case, by Precept I, C is not R'd by A. †2

This defines a sequence in general. The description of the relation R agrees with that of a relation which is familiar to us, namely, that of following in sequence. Of course, it equally agrees with the relation of preceding. Substituting "following" for R, Precept I becomes, that whatever members of the sequence A and B may be; they do not both follow one another. Precept II becomes that, whatever members of the sequence A, B, and C may be, provided they are not all three identical, either A follows B, or B follows C, or C follows A. The first corollary becomes, that no member of the sequence follows itself. The second becomes, that of any two different members of the sequence one follows the other. The third becomes, that whatever members A, B, and C, may be, if A is followed by B, either B follows C or C follows A. The fourth becomes, that whatever members A, B, and C may be, if A is followed by B and B by C, then A is followed by C. The fifth becomes, that whatever members A, B, and C may be, either A is not followed by B, or B is not followed by C, or C is not followed by A.

A sequence may nor may not have an absolute end; that is, a member which is followed by no member; and it may or may not have an absolute beginning; that is, a member which follows no member. These distinctions are not mathematically important in most cases.

562C. A highly important distinction, however, arises in a way which I proceed to describe. We know from the fourth corollary that, taking any member of the sequence M, and any member N, which does not follow M, then whatever member X, we may select, either X does not follow M or is not followed by N. †1 Now M may be such that it is followed by a member N, of which this same thing is true; namely, that there is no member of the sequence which at once follows M and is followed by N. The member N may, in this case, be said hardly to follow M, since it has this property which belongs generally to members which do not follow M. We usually say that N follows next after M. †P1 Now a sequence may either be such that no member has another that hardly, or next, follows it, or such that some members do and some do not, or that all have such members hardly following them.

562D. If a sequence be such that some member has a member that hardly follows it, the sequence is such that some member has a member that hardly precedes it. For, if N follows hard after M, then M hardly precedes N. But, nevertheless, a sequence may be such that every member has a member hardly following it without being such that every member has a member hardly preceding it. An accurate logic will show that this is quite admissible. Suppose, for example, the sequence is composed of all the whole numbers, together with whole numbers each added to a vulgar fraction having 1 for its numerator. And suppose these to follow one another in the reverse order of their values, thus:

4, 3 1/2, 3 1/3, 3 1/4, 3 1/5, . . . 3, 2 1/2, 2 1/3, 2 1/4, 2 1/5, . . . 2, 1 1/2, 1 1/3, 1 1/4, 1 1/5

. . . 1, 1/2, 1/3, 1/4, 1/5 . . .

Then every member, as p + 1/q has p + 1/(q+1) following hard after it. But the whole numbers have no members which they follow hard after.

562E. If a sequence be such that there is some member of it, not its absolute end, which has no member hardly following it, then a part of that sequence forms a sequence with the same, R (or, relation of sequence), which sequence has no absolute beginning. Namely, if M be a member of the total sequence such that some members are R'd by M, but such that taking any member P, either P is not R to M, or else there is a member X, such that P is R to X and X is R to M, in that case the partial sequence composed of all the members of the total sequence that are R'd by M, contains no member that is not R to another member of the same partial sequence. We might express the same thing by saying that if a member of a sequence not its absolute beginning follows hard upon no member, then all the members it follows form a sequence with no absolute end. But the converse of this proposition is not true. That is to say, it is not true that a sequence, every member of which has another following hard after it and itself following hard upon another, cannot be cut up into sequences having neither beginning nor end. For instance, take all numerals of the forms p + 1/q where p and q are whole numbers; and let them be arranged in the order of their magnitudes, thus:

1/5, 1/4, 1/3, 1/2, 2/3, 3/4, 4/5 . . . 1 1/5, 1 1/4, 1 1/3, 1 1/2, 1 2/3, 1 3/4, . . . 2 1/5, 2 1/4, 2 1/3, 2 2/3, 2 3/4, 2 4/5.

Here it is plain that every member has an assignable one next after and another next before it; and yet the sequence can be broken up into partial sequences, each without beginning or end.

562F. Let us call a sequence of which every member follows hard after a second, and is itself followed hard after by a third, a sparse sequence; and let us call a sparse sequence which cannot be broken into parts which want either beginning or end (except so far as they may retain the infinity of the total sequence), a simple sparse sequence.

Then the system of pure number may be defined as a simple sparse sequence, having an absolute beginning, called zero, but no absolute end. Of course, zero might be dropped from the system, or zero and unity might both be dropped, and so on ad infinitum. But slightly simpler definitions of addition and multiplication are obtained by retaining the zero.

562G. Arithmetic begins with the following fundamental theorem:
Whatever character belongs to zero, and belongs to the number following hard after any number to which it belongs, belongs to all numbers.

Demonstration. Let X be a character which zero possesses. And suppose that, whatever number N may be, either N does not possess X, or else N1, the number that follows hard after N, possesses X. Then I say that every number possesses X.

For, since zero possesses X, and zero is followed by every other number (by the second corollary above, since zero, as absolute beginning, follows no number) there are numbers (or, at least, a number) which are followed by whatever numbers there may be which do not possess X. Let us consider, then, the sequence composed of those numbers which are smaller than whatever numbers there may be which do not possess X; these numbers being taken in their order, that is, with the same R as the total sequence. (Even a single number may be regarded as a sequence of which the absolute beginnings and end are identical.) This sequence may either be the total sequence of numbers, in which case it will have no end, or it may be a partial sequence. In the latter case, every number of this sequence will be followed by every number which does not belong to it, since the latter is not followed by some number not possessing X which does follow every number of this partial sequence. Hence, since the sequence of pure number is a simple sparse sequence, it follows that the partial sequence (if it be partial) must have an absolute end. Let N be this absolute end. Then N is followed by whatever number there may be which does not possess X. But by the first corollary, no number is followed by itself. Hence N possesses X. Hence, by hypothesis, N1, the number that follows hard after N, possesses X. Hence, by the definition of "following hard after," N1 is followed by every number except itself, and follows N; that is, by whatever number there may be which does not possess X. But this is contrary to the hypothesis that N was the absolute end of the partial series; that is, was not followed by any number followed by whatever number does not possess X. Hence, the supposition that the series is a partial one is absurd. Hence, every number is followed by whatever number there may be that does not possess X. But, by the first corollary, no number is followed by itself. Hence, every number possesses X. Q.E.D.

562H. By means of this theorem all the elementary propositions of arithmetic as to the associativeness and distributiveness of addition and multiplication, as well as the fundamental principles of subtraction and division of whole numbers are easily deduced from the following definitions of addition and multiplication. Of course subtraction and division are to be defined as inverse operations: Definition of addition. The sum of two numbers, M and N, is a number depending upon M and N according to the following precepts.

Precept III. The Sum of 0 and 0 is 0.

Precept IV. The sum of any number, M, and of the number N1, next following any number, N, is the number next following the sum of M and N.

562I. Definition of multiplication. The product of two numbers, M and N, is a number depending upon M and N according to the following precepts:
Precept V. The product of 0 and 0 is 0.

Precept VI. The product of any number, M, and of the number, N1, next following any number, N, is the number which is the sum of M and the product of M and N. This gives a sufficient idea of pure number. But I will remark that, when this method is applied to division, it leads so easily into the theory of numbers that it is difficult to restrain the pupil from that line of thought. But I need not say that I should not teach the doctrine of pure number to young children. At present, I am merely collecting the conceptions of the subject. I shall pass next to the application of number to the counting of collections. †1



Paper 18: Infinitesimals †1

563. Will you kindly accord me space for a few remarks about Infinity and Continuity which I seem called upon to make by several notes to Professor Royce's Supplementary Essay in his strong work The World and the Individual? I must confess that I am hardly prepared to discuss the subject as I ought to be, since I have never had an opportunity sufficiently to examine the two small books by Dedekind, †2 nor two memoirs by Cantor, †3 that have appeared since those contained in the second volume of the Acta Mathematica. I cannot even refer to Schröder's Logic.

564. (1) There has been some question whether Dedekind's definition of an infinite collection or that which results from negativing my definition of a finite collection is the best. It seems to me that two definitions of the same conception, not subject to any conditions, as a figure in space, for example, is subject to geometrical conditions, must be substantially the same. I pointed out (Am. Journ. Math. IV. 86, †4 but whether I first made the suggestion or not I do not know) that a finite collection differs from an infinite collection in nothing else than that the syllogism of transposed quality [quantity] is applicable to it (and by the consequences of this logical property). For that reason, the character of being finite seemed to me a positive extra determination which an infinite collection does not possess. Dr. Dedekind defines an infinite collection as one of which every echter Theil is similar to the whole collection. †5 It obviously would not do to say a part, simply, for every collection, even if it be infinite, is composed of individuals; and these individuals are parts of it, differing from the whole in being indivisible. Now I do not believe that it is possible to define an echter Theil without substantially coming to my definition. But, however that may be, Dedekind's definition is not of the kind of which I was in search. I sought to define a finite collection in logical terms. But a "part," in its mathematical, or collective, sense, is not a logical term, and itself requires definition.

565. (2) Professor Royce remarks that my opinion that differentials may quite logically be considered as true infinitesimals, if we like, is shared by no mathematician "outside of Italy." †1 As a logician, I am more comforted by corroboration in the clear mental atmosphere of Italy than I could be by any seconding from a tobacco-clouded and bemused land (if any such there be) where no philosophical eccentricity misses its champion, but where sane logic has not found favor. Meantime, I beg leave briefly to submit certain reasons for my opinion.

566. In the first place, I proved in January, 1897, in an article in the Monist (VII 215), †2 that the multitude of possible collections of members of any given collection whatever is greater than the multitude of the latter collection itself. That demonstration is so simple, that, with your permission, I will here repeat it. If there be any collection as great as the multitude of possible collections of its members, let the members of one such collection be called the A's. Then, by Cantor's definition of the relation of multitude, there must be some possible relation, r, such that every possible collection of A's is r to some A, while no two possible collections of A's are r to the same A. But now I will define a certain possible collection of A's, which I will call the collection of B's, as follows: Whatever A there may be that is not included in any collection of A's that is r to it, shall be included in the collection of B's, and whatever A there may be that is included in a collection of A's that is r to it, shall not be included in the collection of B's. If there is any A to which no collection of A's stands in the relation r, I do not care whether it is included among the B's or not. Now I say the collection of B's is not in the relation r to any A. For every A is either an A to which no collection of A's stands in the relation r, or it is included in a collection of A's that is r to it, or it is excluded from every collection of A's that is r to it. Now the collection of B's, being a collection of A's, is not r to any A to which no collection of A's is r; and it is not r to any A that is included in a collection of A's that is r to it, since only one collection of A's is r to the same A, so that were that the case the A in question would be a B, contrary to the definition which makes the collection of B's exclude every A included in a collection that is r to it; and finally, the collection of B's is not r to any A not included in any collection of A's that is r to it, since by definition every such A is a B, so that, if the collection of B's were r to that A, that A would be included in a collection of A's that was r to it. It is thus absurd to say that the collection of B's is r to any A; and thus there is always a possible collection of A's not r to any A; in other words, the multitude of possible collections of A's is greater than the multitude of the A's themselves. That is, every multitude is less than a multitude; or, there is no maximum multitude.

567. In the second place I postulate that it is an admissible hypothesis that there may be a something, which we will call a line, having the following properties: first, points may be determined in a certain relation to it, which relation we will designate as that of "lying on" that line; second, four different points being so determined, each of them is separated from one of the others by the remaining two; third, any three points, A, B, C, being taken on the line, any multitude whatever of points can be determined upon it so that every one of them is separated from A by B and C.

568. In the third place, the possible points so determinable on that line cannot be distinguished from one another by being put into one-to-one correspondence with any system of "assignable quantities." For such assignable quantities form a collection whose multitude is exceeded by that of another collection, namely, the collection of all possible collections of those "assignable quantities." But points are, by our postulate, determinable on the line in excess of that or of any other multitude. Now, those who say that two different points on a line must be at a finite distance from one another, virtually assert that the points are distinguishable by corresponding (in a one-to-one correspondence) to different individuals of a system of "assignable quantities." This system is a collection of individual quantities of very moderate multitude, being no more than the multitude of all possible collections of integral numbers. For by those "assignable quantities" are meant those toward which the values of fractions can indefinitely approximate. According to my postulate, which involves no contradiction, a line may be so conceived that its points are not so distinguishable and consequently can be at infinitesimal distances.

Since, according to this conception, any multitude of points whatever are determinable on the line (not, of course, by us, but of their own nature), and since there is no maximum multitude, it follows that the points cannot be regarded as constituent parts of the line, existing on it by virtue of the line's existence. For if they were so, they would form a collection; and there would be a multitude greater than that of the points determinable on a line. We must, therefore, conceive that there are only so many points on the line as have been marked, or otherwise determined, upon it. Those do form a collection; but ever a greater collection remains determinable upon the line. All the determinable points cannot form a collection, since, by the postulate, if they did, the multitude of that collection would not be less than another multitude. The explanation of their not forming a collection is that all the determinable points are not individuals, distinct, each from all the rest. For individuals can only be distinct from one another in three ways: First, by acts of reaction, immediate or mediate, upon one another; second, by having per se different qualities; and third, by being in one-to-one correspondence to individuals that are distinct from one another in one of the first two ways. Now the points on a line not yet actually determined are mere potentialities, and, as such, cannot react upon one another actually; and, per se, they are all exactly alike; and they cannot be in one-to-one correspondence to any collection, since the multitude of that collection would require to be a maximum multitude. Consequently, all the possible points are not distinct from one another; although any possible multitude of points, once determined, become so distinct by the act of determination. It may be asked, "If the totality of the points determinable on a line does not constitute a collection, what shall we call it?" The answer is plain: the possibility of determining more than any given multitude of points, or, in other words, the fact that there is room for any multitude at every part of the line, makes it continuous. Every point actually marked upon it breaks its continuity, in one sense.

569. Not only is this view admissible without any violation of logic, but I find — though I cannot ask the space to explain this here — that it forms a basis for the differential calculus preferable, perhaps, at any rate, quite as clear, as the doctrine of limits. But this is not all. The subject of topical geometry has remained in a backward state because, as I apprehend, nobody has found a way of reasoning about it with demonstrative rigor. But the above conception of a line leads to a definition of continuity very similar to that of Kant. Although Kant confuses continuity with infinite divisibility, yet it is noticeable that he always defines a continuum as that of which every part (not every echter Theil) has itself parts. This is a very different thing from infinite divisibility, since it implies that the continuum is not composed of points, as, for example, the system of rational fractions, though infinitely divisible, is composed of the individual fractions. If we define a continuum as that every part of which can be divided into any multitude of parts whatsoever — or if we replace this by an equivalent definition in purely logical terms — we find it lends itself at once to mathematical demonstrations, and enables us to work with ease in topical geometry.

570. (3) Professor Royce wants to know †1 how I could, in a passage which he cites, attribute to Cantor the above opinion about infinitesimals. My intention in that passage was simply to acknowledge myself, in a general way, to be no more than a follower of Cantor in regard to infinity, not to make him responsible for any particular opinion of my own. However, Cantor proposed, if I remember rightly, so far to modify the kinetical theory of gases as to make the multitude of ordinary atoms equal to that of the integral numbers, and that of the atoms of ether equal to the multitude of possible collections of such numbers. †1 Now, since it is essential to that theory that encounters shall take place, and that promiscuously, it would seem to follow that each atom has, in the random distribution, certain next neighbors, so that if there are an infinite multitude in a finite space, the infinitesimals must be actual real distances, and not the mere mathematical conceptions, like √-1, which is all that I contend for.



Paper 19: Nomenclature and Divisions of Dyadic Relations †1

§1. NomenclatureE

571. A dyadic relation is a character whose being consists in the logical possibility of a definite fact concerning an ordered pair, or dyad, of subjects; the first of these being termed the relate, the second the correlate; and the relation is said to subsist between the relate and correlate when the fact in whose possibility its being consists actually has place between these objects. The relation, by itself, is, therefore, an ens rationis and mere logical possibility; but its subsistence is of the nature of a fact. When the quality of the fact concerning two objects is considered, without reference to any distinction between these subjects other than that which this fact establishes, and therefore regardless of which of them is relate, which correlate, its possibility is termed by the author a relationship. (It is a useful distinction, but cannot be translated into every language.)

572. The broadest division of dyadic relations is into those which can only subsist between two subjects of different categories of being (as between an existing individual and a quality) and those which can subsist between two subjects of the same category. A relation of the former kind may advantageously be termed a reference; a relation of the latter kind, a dyadic relation proper.

573. A dyadic relation proper is either such as can only have place between two subjects of different universes of discourse (as the membership of a natural person in a corporation), or is such as can subsist between two objects of the same universe. A relation of the former description may be termed a referential relation; a relation of the latter description, a rerelation. †P1

574. A rerelation may either be such as can only subsist between characters or between laws (such as the relation of "essentially depending upon"), or it may be such as can subsist between two existent individual objects. In the former case, it may be termed a modal relation (not a good term), in the latter case an existential relation. The author's writings on the logic of relations †P2 were substantially restricted to existential relations; and the same restriction will be continued in the body of what here follows. A note at the end of this section will treat of modal relations.

575. The number of different species of existential relations for which technical designations are required is so great that it will be best to adopt names for them which shall, by their form, furnish technical definitions of them, in imitation of the nomenclature of chemistry. The following rules will here be used. Any name (for which in this statement of the rules of word-formation we may put x), having been adopted for all relations of a given description, the preposition extra (or ex, or e) will be prefixed to that name ("extra x") in order to form a name descriptive of any relation to which the primitive name does not apply; the preposition contra will be prefixed (forming "contra-x,") to make a name applicable only to such relations as consist precisely in the non-subsistence of corresponding relations to which the primitive name does apply; the preposition juxta will be prefixed so as to bear the sense of contra-extra, or (what is the same) extra-contra; the preposition red (or re) will be prefixed to form a name applicable to a relation if, and only if, the correlate of it stands to its relate in a relation to which the primitive name applies, so that, in other words, a "red-x" is a relation the converse of an x; the preposition com (or con, or co,) will be prefixed to form the general name of any relation which consists in its relate and correlate alike standing in one relation of the primitive kind to one and the same individual correlate; the preposition ultra will be prefixed to form a name applicable only to a relation which subsists between any given relate and correlate only in case the former stands in a relation of the primitive kind to some individual to which the latter does not stand in that same relation; the preposition trans will be used so as to be equivalent to contra-red-ultra, or (what is the same) recontrultra, so that A will be in a relation "trans-x" to B, if, and only if, there is an x-relation in which A stands to whatever individual there may be to which B stands in that very same relation; and the preposition super to form the name of a relation which is, at once, ultra and trans, in respect to the very same relation of the primitive kind. Any of these prepositions may be prefixed, in the same sense, occasionally (and where no misunderstanding could result) not only to names of classes of relations and their cognates, but also to relative terms. But it is chiefly the prepositions com, ultra, trans and super, that will be so used. For example, taking the relative term "loves," there will be little occasion to use the first four of the following expressions, especially, the first and third, which become almost meaningless, while the last four will often be convenient.

1. A extra-loves B; that is, stands in some other relation, whether loving besides or not;

2. A contra-loves B; that is, does not love B;

3. A juxta-loves B; that is, stands in some other relation than that of not loving, whether loving or not;

4. A reloves B; that is, is loved by;

5. A coloves B; that is, loves something loved by;

6. A ultraloves B; that is, loves something not loved by;

7. A transloves B; that is, loves whatever may be loved by;

8. A superloves B; that is, loves whatever may be loved by and something else. †1

576. By a seed (granum) of an existential relation is to be understood an existing individual which not only stands in that relation to some correlate, but to which also some relate stands in that relation. By a spike of a relation is to be understood any collection of seeds of it of which it is both true that every one of them stands in that relation to some one of them; and it is also true that to every one seed of the spike some seed of the spike stands in that same relation. †2 Thus, two spikes of the same relation may have common seeds, or one may even be a part of another. A simple spike is a spike not containing any other spike as a part of it.

577. There are four general kinds of consideration on which divisions of existential dyadic relations may be based. The connections between the four systems of division so resulting have not been sufficiently studied to be treated here.

§2. First System of Divisions †3

578. An existential dyadic relation may be termed a lation, to express its possibly subsisting between two existing individuals, and in opposition to an extralation, which is a fictive lation, looked upon as a lation, but which cannot (or, at least, does not) subsist between two existing individuals. In short, it is incompossibility.

579. A lation is either a contralation which does not necessarily subsist between the members of every dyad, or else it is the juxtalation, or coëxistence, which subsists between every such dyad.

580. Schröder, who always conscientiously follows the writer's terminology, except where he sees good and sufficient reasons for departing from it, †P1 thinks that, in place of "coëxistence," the term "compossibility" should be used. This is a nice question. It is to be kept in view that an existential dyadic relation is to be regarded as a brute fact existing between two existing things, whatever may be thought about it. In saying this, we enunciate no metaphysical nor epistemological proposition, but simply say how the matter must be understood. Such a relation has no being at all unless the two things exist. It has not even that mode of being which consists in conceivability. That is, you cannot conceive that A strikes B, for example, while one of them is non-existent. But when we apply a general word to a class of such dyads, we are thinking of more than ever can exist. If the logician says "A strikes B," meaning (as is usual in logic, except when time is expressly under consideration) did strike, is striking, or will strike B, he asserts only that that event either has happened or will, at some time in the endless future, have happened. But it is impossible, "in the nature of things," as we say, that is, is logically impossible, that all that ever will have happened should at any time actually have happened. Thus, the assertion transcends actual existence. This is because "strike" is a general word, and as such, relates to what is, or is not, possible; and because of the axiom that it is impossible that all that is possible should actually exist. Thus, in absolute strictness of language, "striking" is not an existential relation, but only signifies a class of existential relations. But this having, once for all, been well-understood, it becomes permissible to speak of a general relation as an existential relation, meaning a general class of existential relations. We may, thus, even distinguish between what is essentially true of an existential relation and what is accidentally, but universally, true of it, although what in absolute strictness is an existential relation has no other mode of being than existence, and consequently has no essence. It is not possible, if it does not exist. Still, we have to bear in mind that though we allow ourselves to speak of general relations as existential, yet what we mean are classes of existential relations in the strictest sense; and therefore they have no possibility distinct from existence. Accordingly, coexistence is the right word, and compossibility is only a modal relation. It is quite another thing with the extralation, since this is not an existential relation, so that "incompossibility" seems the more appropriate word; but if I were to countenance any modification of my nomenclature for this pair of relations, it would be that of the substitution of "non-coexistence" for "incompossibility"; and yet, how could things not coexist? Non-coexistence is nonsense, while incompossibility is an intelligible modal relation borrowed to fill a blank in the scheme of divisions of lations.

581. A contralation (really, the only kind of lation there is) is to be termed a perlation, if, and only if, there is some individual that stands in this relation to every individual of the universe. A contralation is to be termed an extraperlation, if, and only if, there is nothing which stands in this relation to everything.

A contralation is to be termed a contraperlation, if, and only if, there is something which does not stand in this relation to anything. A contralation is to be termed a juxtalation, †1 if, and only if, everything stands to something or other in this relation.

A contralation is to be termed a reperlation, if, and only if, there is something to which everything stands in this relation. An extrareperlation is any contralation in which there is nothing to which something does not stand. †2

A contrareperlation is a contralation such that there is something to which nothing stands in it. If everything has something or other in a given relation to it, that relation (and none other than such) is a juxtareperlation.

Every perlation is necessarily a juxtareperlation; although not every juxtareperlation is a perlation; and similarly, every reperlation is a juxtaperlation. †1 So, every contraperlation is an extrareperlation; and every contrareperlation is an extraperlation. †2 The converse of none of these propositions is true. Consequently, the eight classes of contralations named divide all contralations into nine classes, as shown in this table:


Figure 1 †1

It is a little puzzling to find that a relation may be at once in a class and in the contra-class. Thus, taking the citizens of a town as a universe of discourse, if there is one of them who is helpful to himself and to all the others, helpfulness becomes for that universe, a perlation, and unhelpfulness a contraperlation. But if there is also in that town a citizen who is helpful to nobody, not even to himself, then helpfulness and unhelpfulness become alike contraperlative perlations. If, however, the first supposition remaining, there is no entirely unhelpful citizen, helpfulness will be a juxtaperlative perlation; and if there is a citizen who is helped by everybody, helpfulness and helpedness will be reperlative perlations.

582. The preposition pene may be prefixed to any one of these terms to show that in speaking of relations to everything or to something we mean everything else or something else, whether this limits or extends the class of relations. The dotted lines of the diagram show the effect of this prefix in slightly enlarging the border classes at the expense of the interior classes. If the character of a perlation is not due merely to the non-existence of anything that is not a correlate of the individual or individuals that make it a perlation, but is essential to the relation itself, regardless of any other general condition, then whatever is in this relation to anything must be in this relation to everything. A perlation of which this is true may be termed an essential perlation. In a similar sense we may speak of an essential reperlation. Thus "loving something coexisting with" is an essential perlation, "coexisting with something that loves" is an essential reperlation. These are said to be formed from the primitive relation of loving. An essentially reperlative perlation must be the juxtalation unless it be the extralation, and so, no lation at all. Thus, either nothing is in the relation of "coexisting with a creator of something coexisting with" to anything, or else everything is in that relation to everything. Thus, an essentially reperlative perlation is not essentially either coexistence or non-coexistence, but is either an accidental juxtalation or an accidental extralation. In the algebra of dyadic relations such an expression is said to be enveloped. An essential perlation is essentially a contraperlation, and an essential reperlation is essentially a contrareperlation.[?] But a contressentiperlation, which consists in the absence of an essential perlation, is itself an essential perlation. Thus, "loving nobody that coexists with," is equivalent to "non-loving somebody coexisting with," although its relation to its primitive is different.

§3. Second System of Divisions †1

583. An existential relation may be termed a suilation, if, and only if, every individual of the universe stands in that relation to itself. An existential relation may be termed an ambilation, if, and only if, every individual of the universe stands in that relation to every other.

584. Every suilation is a juxtasuilation and every ambilation a juxtambilation. Consequently, there will result nine classes of relations as shown in this table:

Figure 2. †1

585. The following terms were proposed by the author in 1870, †1 and since they have been generally accepted by writers on the subject, he is more bound to adhere to them than anybody else, although he does not now think they were very judiciously chosen. Namely, a juxtasuilation is termed a self-relation, a contrasuilation an alio-relation; a contrambilation is called a concurrency, a juxtambilation an opponency.

586. There is but one ambilative suilation. It is the juxtalation, or coëxistence. There is but one contrambilative suilation: it is the relation of individual identity, called numerical identity by the logicians. But the adjective seems needless. There is but one ambilative [contra] suilation: it is the relation of individual otherness, or negation. There is properly no contrambilative contrasuilation: it would be the absurd relation of incompossibility. These four relations are to be termed the Four Cardinal Dyadic Relations of Second Intention. It will be enough to call them the cardinilations, or cardinal relations.

587. Any peneperlation or penereperlation is a juxtambilation; any perlation or reperlation is, in addition, a juxtasuilation. Any penecontraperlation or penecontrareperlation is an extrambilation: any contraperlation or contrareperlation is, in addition, an extrasuilation. Every ambilation is a penereperlative penereperlation †2: every contrambilation is a penecontrareperlative penecontraperlation. Every suilation is a juxtareperlative juxtaperlation: every contrasuilation is an extrareperlative extraperlation. †3

§4. Third System of Divisions †4

588. This system of divisions which depends upon the identity or otherness of relates and correlates not necessarily (as in the Second System) relates and correlates of one another, is multiform and irregular, a condition, it can hardly be doubted, incident to its not having been sufficiently studied.

589. A relation having seeds may be termed a granilation. The term at present in use is a repeating relation, an extragranilation being called a non-repeating relation. Every juxtasuilation is a granilation, and every juxtagranilation is a suilation.

Extragranilations are of subsidiary importance (yet not of very little importance), since an essential extragranilation closely approximates to the nature of a reference.

590. A granilation may contain a spike, †1 or it may not. In the former case, it may be termed a spicalation, in the latter an extraspicalation. Extressentispicalations, or granilations not necessarily containing spikes, comprise the most important class of existential relations in logic. This is the class of transitive relations (the term was introduced by De Morgan). †2 A relation is transitive if, and only if, any individual object in that relation to a second which is in the same relation to a third is itself in that relation to this third. Thus, the relation of being as great as (in all such phrases "as" is, in logic, to be understood in the sense of "at least as") is essentially transitive. For if A is as great as B, and B is as great as C, then A is necessarily as great as C. The relation of "greater than" is also essentially transitive.

591. Imagine all the dyads (or ordered pairs) of individuals in the universe to be arrayed in a matrix (Cayley's term, though the application of the conception to the logic of relations was first made by the author) as follows:

A:A A:B A:C A:D A:E A:F etc.
B:A B:B B:C B:D B:E B:F etc.
C:A C:B C:C C:D C:E C:F etc.
D:A D:B D:C D:D D:E D:F etc.
E:A E:B E:C E:D E:E E:F etc.
F:A F:B F:C F:D F:E F:F etc.
etc. etc. etc. etc. etc. etc. etc.

Figure 3

Let the horizontal rows be termed ranks and the vertical rows columns (Howard Staunton's terms?) Let the diagonal line containing the identical pairs A:A, B :B, etc. be termed the principal, dexter, or leading diagonal (all terms in common use.) Then a relation is transitive if, and only if, whatever pair of dyads be taken, of each of which the first individual is in that relation to the second, then in case the rank of either one of these dyads in the matrix crosses the column of the other on the dexter diagonal, the column of the former dyad crosses the rank of the latter in a dyad of which the first member is in the transitive relation to the second. Thus, suppose, to begin with: that the only individual in the relation whose transitivity is to be examined is B; and that it is in that relation to C alone. Then, it is an extragranilation; but it is, according to the definition, a transitive relation, since B does stand in that relation to whatever there may be that C is so related to so long as C is not so related to anything. But now suppose that, in addition, C is in the relation in question to F. Then, since the column of (B:C) meets the rank of (C:F) in (C:C) on the dexter diagonal, while the rank of (B:C) meets the column (C:F) in (B:F) the relation will not be transitive unless B stands in that relation to F. If, in addition, any other dyad in the same rank as (C:F) belongs to the relation, there will have to be a corresponding dyad of the relation in the rank of (B:C); and if there is any additional dyad in the column of (B:C), there will have to be a corresponding dyad of the relation in the column of (C:F). †1 Therefore, in examining the matrix of a relation in order to ascertain whether the latter is transitive or not, it will be sufficient to consider each pair in the dexter diagonal and examine whether for every dyad in its rank, there is a co-columnar dyad in the rank of every dyad in its (the dexter pair's) column, and conversely. Suppose, for example, we desire to represent a transitive relation in the seven-by-seven matrix of the first figure below with the condition that the squares with crosses shall be occupied by dyads of the relation.

Since the first, fourth, and fifth squares of the dexter diagonals (or say, dexter squares) have blank ranks, they need no consideration; and the same is true of the second and sixth dexter squares, which have blank columns. It is thus only necessary to consider the third and seventh dexter squares. inline image

Figure 4

The third will require dyads for the squares marked 3, after which, the seventh will require dyads in the squares marked 7; and, as it happens, these do not make any new dyads requisite, as they might have done. It will generally be necessary to reexamine all the dexter squares until it is found that no new dyads are required. †1

In this example, there is no need of filling any of the dexter squares; but any of them can be filled without affecting the transitivity of the relation; and in general no transitive extraspicalation is necessarily juxtasuilative; but it may be so in any way. That is, the modification of it by making it include any identical pairs will never destroy its transitivity.

592. On the other hand, a transitive spicalation must subsist between every possible dyad whose members are individuals of the spike. Thus, in order to render the relation whose matrix is shown in Figure 5 transitive, this relation having the two simple spikes A, B, and C, D, E, F, this matrix must be filled up as in Figure 6.

A:B
B:A C:D
D:E
E:F
F:C

Figure 5

A:A A:B
B:A B:B
C:C C:D C:E C:F
D:C D:D D:E D:F
E:C E:D E:E E:F
F:C F:D F:E F:F

Figure 6

Such a relation, whose matrix is entirely composed of full squares having their dexter diagonals on the dexter diagonal of the whole matrix, may be termed a simililation. (The term formerly proposed by the author, "copulative relation," †1 has not found much favor, and is too unsuitable.)

593. A transitive relation which is such that whatever individual stands in this relation to another stands in the same relation to some individual that stands in the same relation to that other is called an idempotent relation †2 (B. Peirce's Term, †3 which is generally received). An idempotent contrasuilation must subsist between individuals at least denumerable in multitude, and may be termed an endlessly divisible idempotence.

594. If a transitive relation is so connected with a relation, called a primitive of it, which may be a reference, that any individual stands in this transitive relation to an individual, if, and only if, the former individual stands in the primitive relation to whatever there may be to which the latter individual stands in that primitive relation, then, and only then, it is to be termed a translation of that primitive relation. †4 Thus, taking "loving" as the primitive relation, the relation of "loving whatever may be loved by" is the translation of loving. Every translation is transitive suilation. Moreover, given any transitive relation, every relate of it stands to its correlate in a translation of that same relation. This is easily seen if we consider that, for example to say that "every lover of a servant of a mistress is a tormentor of that mistress" is the same as to say that "every lover is a tormentor of every mistress whom his beloved may be servant of"; and the same will evidently be true in any analogous case. If, therefore (substituting "lover of" for "servant of" and for "tormentor of"), every lover of a lover of anybody is a lover of the last (which is as much as to say that loving is a transitive relation), it follows that every lover of anybody loves whatever may be loved by his beloved; that is, he stands in the translation of loving to that which he loves. But this does not imply that every transitive relation constitutes a translation. This is evidently untrue, since every translation is a suilation, but not every transitive relation has that character. Since, however, every translation is a transitive relation, it evidently follows that every transitive relation is precisely a compound of some relation and its translation; such as "at once loving and loving everything loved by." This was Schröder's most brilliant discovery, notwithstanding the simplicity of it, when once stated.

595. The ultralation †1 of any primitive relation consists in any relate of it standing to its correlate in relation of standing in the primitive relation to something to which its correlate does not stand in that primitive relation. Thus, "loving something not loved by" is the ultralation of loving. The translation of any primitive is the contraredultralation of the same primitive. Thus, the redultralation of loving is "not loving something loved by" and the contraredultralation is "loving whatever may be loved by."

596. Almost every step in necessary reasoning depends on two premisses which come to the reasoner's knowledge from different quarters. He "puts two and two together," or, in Whewell's †2 terminology, he colligates the facts. But the two premisses may be (in fact they are, although this, as a mere psychological fact does not concern us), colligated, or compounded (De Morgan's †3 term) into a copulative proposition (the traditional term, copulatum being used by Aulus Gellius †4 in the second century of our era, in this sense) from which as a single premiss the conclusion follows. This is to explain the frequent mention of the premiss of a necessary inference. The relation of a necessary conclusion to its premiss is an essential translation of "being true of." For that reason translations may, in all cases, give rise to necessary conclusions; and there is hardly any necessary inference that does not depend on a translation. Consequently, translations and ultralations are logically the most important of all dyadic relations.

597. The compound of the ultralation and translation of the same primitive is to be termed the superlation of that primitive. Thus, superloving is loving everything loved by and loving something not loved by. Essentially transitive relations (at least, all that are of importance) are either translations (and so, suilations) or are superlations (and so, contrasuilations).

598. A juxtasuilation is, strictly speaking, a spicalation having a spike consisting of a single seed. Such a spike may be called a unispike. A spike consisting of two individuals may be termed a bispike. A spike of which every individual stands in the relation, to which the spike belongs, to only one individual of the spike, and in the converse relation to only one individual of the spike, may be termed a cyclical spike. †1 Every simple spike, that is, every spike whose existence does not consist in the existence of other spikes, is a cyclical spike. It is true that a cyclical spike, and even a simple spike, may be composed of innumerable individuals; and in that case we are not compelled to regard it as running round into itself. Consider, for example, the relation "coming next after" in the both-ways endless series.

. . . 10-5, 10-4, 10-3, 10-2, 10-1, 100, 101, 102, 103, 104, 105, 10 . . .

This, according to the definition, is a relation having a single simple spike. But since on one side, the values run toward zero, and on the other side toward infinity, it cannot well be considered as returning into itself. Yet it is a nice question whether we may not properly conceive such a series, if endless, as reaching the limit, which by definition it just fails to reach. For positive endlessness as much transcends the finite series as does the limit, and reaches that limit in becoming infinite. †P1 The limiting term, however is identical with the next term. If that view of the matter be admissible (which is here left doubtful), there is no spike, unless the series returns into itself. The question will be fully considered in the lectures. †1

599. A spike is said wholly to consist of simple spikes only in case every dyad belonging to the relation forms a part of a simple spike belonging to the relation. Thus, the relation which subsists only between the following dyads, is composed wholly of simple spikes: (A:B), (B:C), (C:D), (D:A), (A:E), (E:F), (F:D). For every dyad either belongs to the simple spike (A:B), (B:C), (C:D), (D:A), or to the simple spike (D:A), (A:E), (E:F), (F:D), (D:A); and the circumstance that (D:A) belongs to both simple spikes does not conflict with this. But the spike (A:B), (A:D), (B:C), (C:A), (D:E), (E:F), (F:D) does not wholly consist of simple spikes, since (A:D) does not here belong to any simple spike.

600. A relation of which every correlate stands in this relation to every one of its relates, and consequently to every relate of which every one of its correlates stand in the same relation, is termed an equiparance †2 (the term has been used since Antonius Andreas about A.D. 1300 †3). An extrequiparance is termed a disquiparance †2 (the term has been in common use since Franciscus Mayro †4, about A. D. 1300). Equiparances and disquiparances are also termed convertible and inconvertible relations (De Morgan), †5 and quite commonly reciprocating and non-reciprocating relations. All the relates and correlates of an equiparance form a spike wholly consisting of spikes of two members and spikes of one member. (A spike of two members is either simple or is wholly composed of a simple spike of two members and, at most, two spikes of one member, every spike of one member being simple.)

§5. Fourth System of Divisions

601. This system depends upon the multitudes of relates and correlates. It has been treated with such elaboration by Mr. Kempe, whose "Memoir on the Theory of Mathematical Forms" may be found in the Philosophical Transactions for 1886, that only the most important points will here be noticed. Mr. Kempe's nomenclature is, in a few cases, unsuitable; but it is remarkable that the severest criticism has detected only one or two errors in the extensive work, and that that one, or may be two, attach to matters of minute detail toward the end of the memoir. It is one of the most solid treatises that have ever been written on the logic of relations — a work that goes toward the elevation of man.

602. In the first place, there are two cross dichotomies according as the relation has for its relates on the one hand, or for its correlates on the other, whatever there may be that possesses some general quality, or merely a singular object or singular objects. The names which the author proposed for these classes in 1870 †1 were rightly rejected by Schröder, who names them by reference to the matrix. †2 The ranks and files of a matrix have always in German been called Zeilen and Kolonnen respectively; and Schröder, accordingly names a relative term applicable to a single relate an "Einzeiler," and one applicable to a single correlate an "Einkolonner." A compartment of the matrix is called in German "ein Auge"; †P1 and Schröder calls a relative term applicable only to a single dyad, whether like A:A or like A:B, an Einauger. The author's original term an elementary relative is sufficiently good. But the terms one-rank relation, one-column relation, many-rank relation, many-column relations can be used to advantage.

A one-rank relation may either be accidentally such (as it happens that there is but one "day-bringing" luminary), or essentially such (as "creator of"), and so of course, it is with one-column relations.

603. A general relation, and indeed, any many-ranked many-column relation, may be limited in respect to the multitude of relates to each correlate, and conversely. The sixth chapter of the third volume of Schröder's Logik contains a valuable and entirely original investigation of these. He describes these as relations with empty ranks (Leerzeilen) and empty columns, with Singly occupied (Einbesetzten) ranks or columns, with multioccupied (Meerbesetzten) ranks or columns, with many-gapped (Meerlückige-) ranks or columns, singly gapped (Einlück-) ranks or columns, and with full (Voll-) ranks and columns.

These again may be accidental or essential. Thus, the relation of parentage has essentially two relates to each correlate, but an indefinite finite number of correlates to each relate.

604. Another pair of cross dichotomies depends upon whether the relates of each correlate, on the one hand, and the correlate of each relate, on the other, are, or are not, of a fixed number. Of particular consequence (meaning by "consequence" the leading to conclusions) are those relations for which both these numbers are fixed. Such a relation is termed a correspondence (a common term of mathematics). The number of relates to each correlate is usually prefixed to the designation, being followed by the number of correlates to each relate, with the proposition "to" intervening between the two numbers. Thus, we speak of a "two-to-three correspondence," meaning that each correlate has two relates and each relate three correlates.

605. There is an indefinite division of correspondences into the regular and the irregular. The regular are those in which there is some discernible rule as to what relates are joined to what correlates. Thus, the relations whose matrices are shown in Figures 7, 8 and 9 are all regular two-to-two correspondences; but their rules are different. It is certain, however, that no collection of individuals be it finite or infinite, can be arranged in succession without that succession's perfectly conforming to some kind of regularity. It is true that there is a difference between an accidental and an essential regularity. But the difference does not manifest itself in the existential

A:B A:F
B:C B:G
C:D C:H
D:A D:E
E:B E:F
F:C F:G
G:D G:H
H:A H:E

Figure 7

facts themselves. The problem of how an accidental regularity can be distinguished from an essential one is precisely the problem of inductive logic.

A:B A:E
B:C B:F
C:D C:G
D:E D:H
E:A E:F
F:B F:G
G:C G:H
H:A H:D

Figure 8

A:B A:C
B:C B:D
C:D C:E
D:E D:F
E:F E:G
F:G F:H
G:A G:H
H:A H:B

Figure 9

§6. Note on the Nomenclature and Divisions of Modal Dyadic Relations

606. A modal dyadic relation is either a relation between characters (including qualities and relations of individuals, of characters, and of concepts), or between symbols, or concepts.

607. Dyadic relations between characters mostly correspond to relations between the subjects of those characters or to relations between the symbols of them; and such need not be separately considered. There remain some relations between characters, especially between qualities, which do not seem to be derivative. Such are the relations of "being more intense than," of "being disparate to" (or in applicability to subjects of the same category, as multitude and intensity are disparate). But, so far as appears at present, no particular logical interest attaches to such relations, and they will here be passed by.

608. Dyadic relations between symbols, or concepts, are matters of logic, so far as they are not derived from relations between the objects and the characters to which the symbols refer. Noting that we are limiting ourselves to modal dyadic relations, it may probably be said that those of them that are truly and fundamentally dyadic arise from corresponding relations between propositions. To exemplify what is meant, the dyadic relations of logical breadth and depth, often called denotation and connotation, have played a great part in logical discussions, but these take their origin in the triadic relation between a sign, its object, and its interpretant sign; and furthermore, the distinction appears as a dichotomy owing to the limitation of the field of thought, which forgets that concepts grow, and that there is thus a third respect in which they may differ, depending on the state of knowledge, or amount of information. †1 To give a good and complete account of the dyadic relations of concepts would be impossible without taking into account the triadic relations which, for the most part, underlie them; and indeed almost a complete treatise upon the first †2 of the three divisions of logic would be required. †3



Paper 20: Notes on Symbolic Logic and MathematicsE

§1. Imaging †1

609. A term proposed to translate Abbildung in its logical use. In order to apprehend this meaning, it is indispensable to be acquainted with the history of the meanings of Abbildung. This word was used in 1845 by Gauss †2 for what is called in English a map-projection, which is an incorrect term, since many such modes of representation are not geometrical rectilinear projections at all; and of those which Gauss had in view, but a single one is so. In mathematics Abbildung is translated representation; but this word is preëmpted in logic. Since Bild is always translated image, imaging will answer very well for Abbildung. If a map of the entire globe was made on a sufficiently large scale, and out of doors, the map itself would be shown upon the map; and upon that image would be seen the map of the map; and so on, indefinitely. If the map were to cover the entire globe, it would be an image of nothing but itself, where each point would be imaged by some other point, itself imaged by a third, etc. But a map of the heavens does not show the map itself at all. A Mercator's projection shows the entire globe (except the poles) over and over again in endlessly recurring strips. †3 Many maps, if they were completed, would show two or more different places on the earth at each point of the map (or at any rate on a part of it), like one map drawn upon another. Such is obviously the case with any rectilinear projection of the entire sphere, excepting only the stereographic. These two peculiarities may coexist in the same map.

610. Any mathematical function of one variable may be regarded as an image of its variable according to some mode of imaging. For the real and imaginary quantities correspond, one to one and continuously, to the assignable points on a sphere. Although mathematics is by far the swiftest of the sciences in its generalisations, it was not until 1879 that Dedekind (in the third edition of his recension of Lejeune-Dirichlet's Zahlentheorie, §163, p. 470; but the writer has not examined the second edition) extended the conception to discrete systems in these words: "It very often happens in other sciences, as well as in mathematics, that there is a replacement of every element ω of a system Ω of elements or things by a corresponding element ω' [of a system Ω']. Such an act should be called a substitution. . . . But a still more convenient expression is found by regarding Ω' as the image of Ω, and ω' of ω, according to a certain mode of imaging." And he adds, in a footnote: "This power of the mind of comparing a thing ω, with a thing ω', or of relating ω to ω', or of considering ω' to correspond to ω, is one without which no thought would be possible." (We do not translate the main clause.) This is an early and significant acknowledgment that the so-called "logic of relatives" — then deemed beneath the notice of logicians -is an integral part of logic. This remark remained unnoticed until, in 1895, Schröder devoted the crowning chapter of his great work (Exakte Logik, iii. 553-649) to its development. Schröder says that, in the broadest sense, any relative whatever may be considered as an imaging — "nämlich als eine eventuell bald 'undeutige,' bald 'eindeutige,' bald 'mehrdeutige' Zuordnung." He presumably means that the logical universe is thus imaged in itself. However, in a narrower sense, he says, a mode of imaging is restricted to a relative which fulfills one or other of the two conditions of being never undeutig or being never mehrdeutig. That is, the relation must belong to one or other of two classes, the one embracing such that every object has an image, and the other such that no object has more than one image. Schröder's definitions (however interesting his developments) break all analogy with the important property of the imaging of continua noticed above. If this is to be regarded as essential, an imaging must be defined as a generic relation between an object-class and an image-class, which generic relation consists of specific relations, in each of which one individual, and no more, of the image-class stands to each individual of the object-class, and in each of which every individual of the image-class stands to one individual and no more of the object-class. This is substantially a return to Dedekind's definition, which makes an imaging a synonym for a substitution.

§2. Individual †1

611. (As a technical term of logic, individuum first appears in Boëthius, †2 in a translation from Victorinus, no doubt of {atomon}, a word used by Plato (Sophistes, 229 D) for an indivisible species, and by Aristotle, often in the same sense, but occasionally for an individual. Of course the physical and mathematical senses of the word were earlier. Aristotle's usual term for individuals is {ta kath' hekasta}, Latin singularia, English singulars.) Used in logic in two closely connected senses. (1) According to the more formal of these an individual is an object (or term) not only actually determinate in respect to having or wanting each general character and not both having and wanting any, but is necessitated by its mode of being to be so determinate. See Particular (in logic). †3

612. This definition does not prevent two distinct individuals from being precisely similar, since they may be distinguished by their hecceities (or determinations not of a generalizable nature); so that Leibnitz' principle of indiscernibles is not involved in this definition. Although the principles of contradiction and excluded middle may be regarded as together constituting the definition of the relation expressed by "not," yet they also imply that whatever exists consists of individuals. †4 This, however, does not seem to be an identical proposition or necessity of thought; for Kant's Law of Specification (Krit. d. reinen Vernunft, 1st ed., 656; 2d ed., 684; but it is requisite to read the whole section to understand his meaning), which has been widely accepted, treats logical quantity as a continuum in Kant's sense, i.e., that every part of which is composed of parts. Though this law is only regulative, it is supposed to be demanded by reason, and its wide acceptance as so demanded is a strong argument in favour of the conceivability of a world without individuals in the sense of the definition now considered. Besides, since it is not in the nature of concepts adequately to define individuals, it would seem that a world from which they were eliminated would only be the more intelligible. A new discussion of the matter, on a level with modern mathematical thought and with exact logic, is a desideratum. A highly important contribution is contained in Schröder's Logik, iii, Vorles. 10. What Scotus says (Quaest. in Met., VII 9, xiii and xv) is worth consideration.

613. (2) Another definition which avoids the above difficulties is that an individual is something which reacts. That is to say, it does react against some things, and is of such a nature that it might react, or have reacted, against my will.

This is the stoical definition of a reality; but since the Stoics were individualistic nominalists, this rather favours the satisfactoriness of the definition than otherwise. It may be objected that it is unintelligible; but in the sense in which this is true, it is a merit, since an individual is unintelligible in that sense. It is a brute fact that the moon exists, and all explanations suppose the existence of that same matter. That existence is unintelligible in the sense in which the definition is so. That is to say, a reaction may be experienced, but it cannot be conceived in its character of a reaction; for that element evaporates from every general idea. According to this definition, that which alone immediately presents itself as an individual is a reaction against the will. But everything whose identity consists in a continuity of reactions will be a single logical individual. Thus any portion of space, so far as it can be regarded as reacting, is for logic a single individual; its spatial extension is no objection. With this definition there is no difficulty about the truth that whatever exists is individual, since existence (not reality) and individuality are essentially the same thing; and whatever fulfills the present definition equally fulfills the former definition by virtue of the principles of contradiction and excluded middle, regarded as mere definitions of the relation expressed by "not." As for the principle of indiscernibles, if two individual things are exactly alike in all other respects, they must, according to this definition, differ in their spatial relations, since space is nothing but the intuitional presentation of the conditions of reaction, or of some of them. But there will be no logical hindrance to two things being exactly alike in all other respects; and if they are never so, that is a physical law, not a necessity of logic. This second definition, therefore, seems to be the preferable one. Cf. Particular (in logic). †1

§3. Involution †2

614. A term of Symbolic Logic borrowed from algebra, where it means the raising of a base to a power. In logic it has two different senses. (1) Relative involution: let lwm denote any lover of a well-wisher of a man. That is, any individual A is denoted by l w m, provided there are in existence individuals B and C (who may be identical with each other or with A), such that A loves B, while B wishes well to C, and C is a man. Further, let lwm denote any individual A, if, and only if, there is in existence an individual C, who is a man, and who is such that taking any individual B whatever, if B is a well-wisher of C, then A is a lover of B. The operation of combining l and w in this statement is termed "progressive involution." †3 Again, let lwm denote any individual A, if, and only if, there is in existence an individual B, who is loved by A, and who is such that taking any individual C whatever, if C is wished well by B, then C is a man. The operation of combining w and m in this statement is termed "regressive involution." †4 These designations were adopted because of the analogy of the general formulae to those of involution in the algebra of quantity.

These kinds of involution are not, at present, in use in symbolical logic; but they are, nevertheless, useful, especially in developing the conception of continuity. These two kinds of involution together constitute relative involution.

615. (2) Non-relative involution: consisting in the repeated introduction of the same premiss into a reasoning; as, for example, the half dozen simple premisses upon which the Theory of Numbers is based are introduced over and over again in the reasoning by which its myriad theorems are deduced. In exact logic the regular process of deduction begins by non-relatively multiplying together all the premisses to make one conjunctive premiss, from which whatever can be deduced by using those premisses as often as they are introduced as factors, can be deduced by processes of "immediate inference" from that single conjunctive premiss. But the general character of the conclusion is found to depend greatly upon the number of times the same factor is multiplied in. From this circumstance the importance and the name of non-relative involution arise.

§4. Logic (Exact) †1

616. The doctrine that the theory of validity and strength of reasoning ought to be made one of the "exact sciences," that is, that generalisations from ordinary experience ought, at an early point in its exposition, to be stated in a form from which by mathematical, or expository, reasoning, the rest of the theory can be strictly deduced; together with the attempt to carry this doctrine into practice.

617. This method was pursued, in the past, by Pascal (1623-62), Nicolas Bernoulli (1687-1759), Euler (1708-83), Ploucquet (1716-90), Lambert (1728-77), La Place (1749-1827), De Morgan (1806-71), Boole (1815-64), and many others; and a few men in different countries continue the study of the problems opened by the last two named logicians as well as those of the proper foundations of the doctrine and of its application to inductive reasoning. The results of this method, thus far, have comprised the development of the theory of probabilities, the logic of relatives, advances in the theory of inductive reasoning (as it is claimed), the syllogism of transposed quantity, the theory of the Fermatian inference, considerable steps towards an analysis of the logic of continuity and towards a method of reasoning in topical geometry, contributions towards several branches of mathematics by applications of "exact" logic, the logical graphs called after Euler and other systems for representing in intuitional form the relations of premisses to conclusions, and other things of the same general nature.

618. There are those, not merely outside the ranks of exact logic, but even within it, who seem to suppose that the aim is to produce a calculus, or semi-mechanical method, for performing all reasoning, or all deductive inquiry; but there is no reason to suppose that such a project, which is much more consonant with the ideas of the opponents of exact logic than with those of its serious students, can ever be realised. The real aim is to find an indisputable theory of reasoning by the aid of mathematics. The first step in the order of logic towards this end (though not necessarily the first in the order of inquiry) is to formulate with mathematical precision, definiteness, and simplicity, the general facts of experience which logic has to take into account.

619. The employment of algebra in the investigation of logic is open to the danger of degenerating into idle trifling of too rudimentary a character to be of mathematical interest, and too superficial to be of logical interest. It is further open to the danger that the rules of the symbols employed may be mistaken for first principles of logic. An algebra which brings along with it hundreds of purely formal theorems of no logical import whatever must be admitted, even by the inventor of it, to be extremely defective in that respect, however convenient it may be for certain purposes. On the other hand, it is indisputable that algebra has an advantage over speech in forcing us to reason explicitly and definitely, if at all. In that way it may afford very considerable aid to analysis. It has been employed with great advantage in the analysis of mathematical reasonings.

Algebraic reasoning involves intuition just as much as, though more insidiously than, does geometrical reasoning; and for the investigation of logic it is questionable whether the method of graphs is not superior. Graphs cannot, it is true, readily be applied to cases of great complexity; but for that very reason they are less liable to serve the purposes of the logical trifler. In the opinion of some exact logicians, they lead more directly to the ultimate analysis of logical problems than any algebra yet devised. See Logical Diagram (or Graph). †1

620. It is logical algebra, however, which has chiefly been pursued. De Morgan invented a system of symbols, which had the signal advantage of being entirely new and free from all associations, misleading or otherwise. Although he employed them for synthetical purposes almost exclusively, yet the great generality of some of the conceptions to which they led him is sufficient to show that they might have been applied with great advantage in analysis. Boole was led, no doubt from the consideration of the principles of the calculus of probabilities, to a wonderful application of ordinary algebra to the treatment of all deductive reasoning not turning upon any relations other than the logical relations between non-relative terms. By means of this simple calculus, he took some great steps towards the elucidation of probable reasoning; and had it not been that, in his pre-Darwinian day, the notion that certain subjects were profoundly mysterious, so that it was hopeless, if not impious, to seek to penetrate them, was still prevalent in Great Britain, his instrument and his intellectual force were adequate to carrying him further than he actually went. Most of the exact logicians of today are, from the nature of the case, followers of Boole. They have modified his algebra by disusing his addition, subtraction, and division, and by introducing a sign of logical aggregation. This was first done by Jevons; and he proposed ·|·, a sign of division turned up, to signify this operation. Inasmuch as this might easily be read as three signs, it would, perhaps, be better to join the two dots by a light curve, thus inline image. Some use the sign + for logical aggregation. The algebra of Boole has also been amplified so as to fit it for the logic of relatives. The system is, however, far from being perfect. See Relatives (logic of). †1

621. Certain terms of exact logic may be defined as follows:— †2
Copula is often defined as that which expresses the relation between the subject-term and the predicate-term of a proposition. But this is not sufficiently accurate for the purposes of exact logic. Passing over the objection that it applies only to categorical propositions, as if conditional and copulative propositions had no copula, contrary to logical tradition, it may be admitted that a copula often does fulfill the function mentioned; but it is only an accidental one, and its essential function is quite different. Thus, the proposition, "Some favoured patriarch is translated" is essentially the same as "A translated favoured patriarch is"; and "Every mother is a lover of that of which she is a mother" is the same as "A mother of something not loved by her is not." In the second and fourth forms, the copula connects no terms; but if it is dropped, we have a mere term instead of a proposition. Thus the essential office of the copula is to express a relation of a general term or terms to the universe. The universe must be well known and mutually known to be known and agreed to exist, in some sense, between speaker and hearer, between the mind as appealing to its own further consideration and the mind as so appealed to, or there can be no communication, or "common ground," at all. The universe is, thus, not a mere concept, but is the most real of experiences. Hence, to put a concept into relation to it, and into the relation of describing it, is to use a most peculiar sort of sign or thought; for such a relation must, if it subsist, exist quite otherwise than a relation between mere concepts. This, then, is what the copula essentially does. This it may do in three ways: first, by a vague reference to the universe collectively; second, by a reference to all the individuals existent in the universe distributively; third, by a vague reference to an individual of the universe selectively. "It is broad daylight," I exclaim, as I awake. My universe is the momentary experience as a whole. It is that which I connect as object of the composite photograph of daylight produced in my mind by all my similar experiences. Secondly, "Every woman loves something" is a description of every existing individual in the universe. Every such individual is said to be coexistent only with what, so far as it is a woman at all, is sure to be a lover of some existing individual. Thirdly, "Some favoured patriarch is translated" means that a certain description applies to a select individual. A hypothetical proposition, whether it be conditional (of which the alternative, or disjunctive, proposition is a mere species, or vice versa, as we choose to take it) or copulative, is either general or ut nunc. A general conditional is precisely equivalent to a universal categorical. "If you really want to be good, you can be," means "Whatever determinate state of things may be admissibly supposed in which you want to be good is a state of things in which you can be good." The universe is that of determinate states of things that are admissible hypothetically. It is true that some logicians appear to dispute this; but it is manifestly indisputable. Those logicians belong to two classes: those who think that logic ought to take account of the difference between one kind of universe and another (in which case, several other substantiæ of propositions must be admitted); and those who hold that logic should distinguish between propositions which are necessarily true or false together, but which regard the fact from different aspects. The exact logician holds it to be, in itself, a defect in a logical system of expression, to afford different ways of expressing the same state of facts; although this defect may be less important than a definite advantage gained by it. The copulative proposition is in a similar way equivalent to a particular categorical. Thus, to say "The man might not be able voluntarily to act otherwise than physical causes make him act, whether he try or not," is the same as to say that there is a state of things hypothetically admissible in which a man tries to act one way and voluntarily acts another way in consequence of physical causes. As to hypotheticals ut nunc, they refer to no range of possibility, but simply to what is true, vaguely taken collectively.

622. Although it is thus plain that the action of the copula in relating the subject-term to the predicate-term is a secondary one, it is nevertheless necessary to distinguish between copulas which establish different relations between these terms. Whatever the relation is, it must remain the same in all propositional forms, because its nature is not expressed in the proposition, but is a matter of established convention. With that proviso, the copula may imply any relation whatsoever. So understood, it is the abstract copula of De Morgan (Camb. Philos. Trans., x. 339). A transitive copula is one for which the mood Barbara is valid. Schröder has demonstrated the remarkable theorem that if we use IS in small capitals to represent any one such copula, of which "greater than" is an example, then there is some relative term r, such that the proposition "S IS P" is precisely equivalent to "S is r to P and is r to whatever P is r to." A copula of correlative inclusion is one for which both Barbara and the formula of identity hold good. Representing any one such copula by is in italics, there is a relative term r, such that the proposition "S is P" is precisely equivalent to "S is r to whatever P is r to." If the last proposition follows from the last but one, no matter what relative r may be, the copula is called the copula of inclusion, used by C. S. Peirce, Schröder, and others. De Morgan uses a copula defined as standing for any relation both transitive and convertible. The latter character consists in this, that whatever terms I and J may be, if we represent this copula by is in black-letter, then from "I is J" it follows that "J is I." From these two propositions, we conclude, by Barbara, that "I is I." Such copulas are, for example, "equal to," and "of the same colour as." For any such copula there will be some relative term r, such that the proposition "S is P" will be precisely equivalent to "S is r to everything, and only to everything, to which P is r." Such a copula may be called a copula of correlative identity. If the last proposition follows from the last but one, no matter what relative r may be, the copula is the copula of identity used by Thomson, Hamilton, Baynes, Jevons, and many others.

It has been demonstrated by Peirce that the copula of inclusion is logically simpler than that of identity. †1

623. Dialogism. A form of reasoning in which from a single premiss a disjunctive, or alternative, proposition is concluded introducing an additional term; opposed to a syllogism, in which from a copulative proposition a proposition is inferred from which a term is eliminated.

Syllogism.

All men are animals, and all animals are mortal;

∴ All men are mortal.

Dialogism.

Some men are not mortal;

∴ Either some men are not animals, or some animals are not mortal.

624. Dimension. An element or respect of extension of a logical universe of such a nature that the same term which is individual in one such element of extension is not so in another. Thus, we may consider different persons as individual in one respect, while they may be divisible in respect to time, and in respect to different admissible hypothetical states of things, etc. This is to be widely distinguished from different universes, as, for example, of things and of characters, where any given individual belonging to one cannot belong to another. The conception of a multidimensional logical universe is one of the fecund conceptions which exact logic owes to O. H. Mitchell. †1 Schröder, in his then second volume, where he is far below himself in many respects, pronounces this conception "untenable." But a doctrine which has, as a matter of fact, been held by Mitchell, Peirce, and others, on apparently cogent grounds, without meeting any attempt at refutation in about twenty years, may be regarded as being, for the present, at any rate, tenable enough to be held.

625. Dyadic relation. A fact relating to two individuals. Thus, the fact that A is similar to B, and the fact that A is a lover of B, and the fact that A and B are both men, are dyadic relations; while the fact that A gives B to C is a triadic relation. Every relation of one order of relativity may be regarded as a relative of another order of relativity if desired. Thus, man may be regarded as man coexistent with, and so as a relative expressing a dyadic relation, although for most purposes it will be regarded as a monad or non-relative term.

§5. Multitude (In Mathematics) †2

626. That relative character of a collection which makes it greater than some collections and less than others. A collection, say that of the A's, is greater than another, say that of the B's, if, and only if, it is impossible that there should be any relation r, such that every A stands in the relation r to a B to which no other A is in the relation r.

627. The precise analysis of the notion is due to G. Cantor, whose definition is, however, a little different in its mode of expression, since it is more abstract. He defines the character in these words: "By Mächtigkeit or cardinal number of a collection (Menge) M, we mean the universal concept, which by the help of our active faculty of thought results from the collection M by abstraction from the characters of the different members (Elemente) of that collection and from the order in which they are given (Gegebensein)." †1

628. A cardinal number, though confounded with multitude by Cantor, is in fact one of a series of vocables the prime purpose of which, quite unlike any other words, is to serve as an instrument in the performance of the experiment of counting; these numbers being pronounced in their order from the beginning, one as each member of the collection is disposed of in the operation of counting. If the operation comes to an end by the exhaustion of the collection, the last cardinal number pronounced is applied adjectivally to the collection, and expresses its multitude, by virtue of the theorem that a collection the counting of which comes to an end, always comes to an end with the pronunciation of the same cardinal number.

629. If the cardinal numbers are considered abstractedly from their use in counting, simply in themselves, as objects of mathematical reasoning, stripped of all accidents not pertinent to such study, they become indistinguishable from the similarly treated ordinal numbers, and are then usually called ordinal numbers by the mathematico-logicians. There is small objection to this; yet it is to be remarked that they are ordinal in different senses in grammar and in the logic of mathematics. For in grammar they are called ordinal as being adapted to express the ordinal places of other things in the series to which those things belong; while in the logic of mathematics the only relevant sense in which they are ordinal is as being defined by a serial order within their own system. The definition of this order is not difficult; but the syntax of ordinary language does not lend itself to the clear expression of such relations in the manner in which they ought to be expressed in order to bring out their logical character. It must, therefore, be here passed by. In fact, none of the doctrines of logic can be satisfactorily expressed under the limitations here imposed, however simple they may be. The doctrine of ordinal numbers is by Dedekind (Was sind und was sollen die Zahlen?) made to precede that of the cardinal numbers; and this is logically preferable, if hardly so imperative as Schröder considers it.

630. The doctrine of the so-called ordinal numbers is a doctrine of pure mathematics; the doctrine of cardinal numbers, or, rather, of multitude, is a doctrine of mathematics applied to logic. The smallest multitude is most conveniently considered to be zero; but this is a question of definition. A finite collection is one of which the syllogism of transposed quantity holds good. Of finite collections, it is true that the whole is greater than any part. It is singular that this is often taken as the type of an axiom, although it has from early times been a matter of familiar knowledge that it is not true of infinite collections. Every addition of one increases a finite multitude. An infinite collection cannot be separated into a lesser collection of parts all smaller than itself.

631. The multitude of all the different finite multitudes is the smallest infinite multitude. It is called the denumeral multitude. (Cantor uses a word equivalent to denumerable; but the other form has the advantage of being differentiated from words like enumerable, abnumerable, which denote classes of multitudes, not, like denumeral, a single multitude.) Following upon this is a denumeral series of multitudes called by C. S. Peirce the first, second, etc. abnumerable multitudes. Each is the multitude of possible collections formed from the members of a collection of the next preceding multitude. They seem to be the same multitudes that are denoted by Cantor as Alephs. The first of them is the multitude of different limits of possible convergent series of rational fractions, and therefore of all the quantities with which mathematical analysis can deal under the limitations of the doctrine of limits. (The imaginaries do not increase the multitude.) What comes after these is still a matter of dispute, and is perhaps of inferior interest. The transition to continuity is, however, a matter of supreme importance for the theory of scientific method; nor is it a very complicated matter; but it cannot be stated under the limitations of expression here imposed upon us.

§6. Postulate †1

632. (1) The earliest definition we have of postulate, which was a technical term of Greek geometers, is by Aristotle. †2 The passage has an appearance of incoherence; it is, however, plain that Aristotle makes a distinction between hypotheses and postulates which Euclid does not draw, and which is irrelevant. Omitting the distinction, the two have this in common — that they are propositions not necessarily true which are assumed as the bases of deductions.

If we turn to the first book of Euclid's Elements, we observe, in the first place, that he calls axioms by the name of common notions, a deliberate choice by him, for Aristotle, before his day, had called them axioms, though Aristotle usually calls them {ta koina}, nearly Euclid's name. These matters of common knowledge, according to Euclid's enumeration of them, are not specially geometrical, except that magnitudes superposable are equal (see the Cent. Dict., "Axiom"). On the other hand, the "postulates" of Euclid are all geometrical. They are as follows (according to the best MS. and all the evidence):—

(a) Between any two points a straight line can be drawn.

(b) Any terminated straight line can be prolonged at either end indefinitely.

(c) About any point in any plane as centre a circle may be described with any radius.

(d) All right angles are equal.

(e) If two straight lines in a plane are cut by a third, making the sum of the internal angles on one side less than two right angles, those two straight lines will meet if sufficiently produced.

(f) Two straight lines cannot enclose a space in a plane.

633. (2) Since Wolff it has been very common among Germans, and among English writers who follow them, to define a postulate as an indemonstrable practical proposition. That is to say, it is an indemonstrable particular proposition, asserting that some general description of an object exists (in the only sense in which pure geometrical forms can be said to exist), in contradistinction to axioms, which were supposed to be indemonstrable theoretical (i.e. universal) propositions, asserting that some general description of an object has no existence as a geometrical form.

It is certainly desirable to have two terms bearing these meanings; but it was an utter misunderstanding to suppose that such were the proper meanings either of the word axiom or of the word postulate. The manner in which this misunderstanding came about is somewhat instructive. An axiom was a perfectly indubitable statement about things, in contradistinction to a definition, which cannot be called in question. On the contrary, a postulate was an indemonstrable proposition, not indubitable. There was some question whether certain postulates might not be considered to be axiomatic. When that was done, all the remaining postulates were particular propositions; namely, the first three of Euclid's list. This view was aided by the illogical notion that definitions could be considered as among the foundations of geometrical truth. Some writers went so far as to say that definitions were, or ought to be, the sole foundation of geometry — an extreme nominalistic position. But if definitions are allowed to take such a position, one postulate, at most, suffices, without any axiom; and all the rest of geometry can be thrown into a single definition. Namely, it is only necessary to postulate, say, that a point is possible, and to define a point in such a way as to make it cover the whole of geometry. This was not seen; and the practice of throwing geometrical truth over into definitions so far prevailed as to aid in restricting postulates to particular propositions. That such assumptions of possibility had a markedly different logical function from assumptions of impossibility was sufficiently clear to Wolff and the earlier writers whom he followed to cause him to put forth his definitions of axiom and postulate; and they recommended themselves all the more, because the postulates had become so familiar that it was no longer recognized that they were open to doubt.

634. (3) Kant calls his principles of modality "postulates of empirical thought" in the sense of judgments which are objectively analytical but subjectively synthetical. In fact, the principles as stated by him are not synthetical in any sense whatever, but are mere definitions.

§7. Presupposition †1

635. Presupposition is either a conjecture or what is better called in English a Postulate. (q. v.)

As a philosophical term it translates the German Voraussetzung, and is presumably preferred to "postulate" by Germans and others imperfectly acquainted with the English language, because they suppose that postulate in English has the same meaning as Postulat in German, which is not true; for the English retains the old meaning, while the German has generally adopted the conception of Wolff. If postulate does not exactly translate German Voraussetzung, it comes, at any rate, quite as near to doing so as presupposition; a good translation would be "assumption."

§8. Relatives †1

636. If from any proposition having more than one subject (used to include "objects") we strike out the indices of the subjects, as in "— praises — to —," "— dat in matrimonium —," what remains and requires at least two insertions of subject-nouns to make a proposition is a "relative term," or "relative rhema," called briefly a "relative." The relative may be converted into a complete assertion by filling up the blanks with proper names or abstract nouns; this serves as a criterion.

But in such a relative there must be such an idea of the difference between the subjects to be applied that "dat in matrimonium" shall be different from "datur in matrimonium." In order to free ourselves from the accidents of speech, we might represent the sentence by the following diagram:

inline image

or, as follows:

dijk (Cinna = i, Cossutia = j, Caesar = k).

Then the relative will appear as

inline image

or as:

dijk.

But in either case, in order to explain what is meant, it will be necessary to explain how those three tails, or the three letters i, j, k, differ. The order shows which of three indices is given, which giver, which recipient.

637. Relatives may be more or less general like other terms, that is, one relative may be predicable of members of a set of which another is not, while the latter is predicable only of members of sets of which the former is predicable. By a set is meant an ordered system, so that ABC and BCA, though the same collection, are different sets. As any general term is predicable of any one of an aggregate of individuals, so a relative is predicable of any one of an aggregate of sets; and each such set may be regarded as an individual relative. By a system is meant an individual of which if anything is true, the truth of it consists in certain things being true of certain other individuals, called its members, regardless of the system. A system is either a sorite, heap, or mere collection, or it is a set. A sorite is a system of which, if anything is true, its truth consists of the truth of one predicate for any one of the members. A set is a system of which the truth of anything consists in the truth of different predicates. Of course the idea of relation is involved in the idea of a system. As it is very important for the understanding of relations that the conception of a system should be perfectly clear, let us consider the latter a moment in its simplest form, that of a sorite or mere collection. ABC is a sorite. Thus, it is true of it that it contains the three first letters of the alphabet, and the truth of that consists in A, B, and C being each one of the first three letters of the alphabet. It is true that it contains nothing but the first letters of the alphabet, because it is true of A, B, C severally that each is nothing but one of the first three letters of the alphabet. AB is a different sorite, because something is true of it which is not true of ABC. A may be regarded as a sorite provided we mean not A in its first intention and being, but a something whose being consists in A's being. The collection A is not the letter A, but it contains A and nothing else. If it be said that there is no such thing, the reply is that every collection, every system may be said to be an ens rationis. To this point we shall return. Even Nothing may be said to be a collection. For when we say that Nothing is less than 1, we do not mean that a self-subsisting individual is so, but that an ens rationis whose mode of being consists in the absence of everything is less than 1. The sorite ABC is other than ABΓ. But should I say that ABC contains two of the letters of Caesar's first name, and subsequently learn that that was a mistake, the real name being Gaius, that would not make ABC a different sorite.

638. That in the reality which corresponds to a proposition with a relative predicate is called the fundamentum relationis. A relationship is a system of such fundamenta. Relation is the relative character, conceived as belonging in different ways to the different relates, and (owing to the somewhat undue prominence given by familiar languages to one of these) especially to the relate which is denoted by the noun which is the subject nominative.

639. Relatives and relations are said to differ in their orders, according to the numbers of their relates. Dyadic or dual relations, or relatives of two relates, of which the second is called the correlate, differ somewhat widely from plural, or polyadic, relations. Triadic relations have all the principal characters of tetradic and higher relations. In fact, a compound of two triadic relatives may be a tetradic relative; as "praiser of to a maligner of — to —."

640. Relatives may be compounded in all the ways in which other terms can be compounded as well as in other ways closely related to those. Thus, A may be said to be at once a lover and a servant of B, and it may be said that there is something, X, such that A is a lover of X, while X is a servant of B; so that A is a lover of a servant of B. This mode of composition is called relative multiplication. So, not only may it be said that A is either a lover or a servant of B (not excluding both), but also that whatever X may be, either A is a lover of X or X is a servant of B; that is, A is a lover of everything there is besides servants of B. (This wording, by Schröder, slightly violates English idiom, but is valuable as showing the analogy to aggregation.) This mode of composition is called relative addition. So, again, it may not only be said that A is if a lover then a servant of B, but also that whatever X may be, if A is a lover of X, then X is a servant of B; that is, A is a lover only of servants of B. This is called relative regressive involution. Or it may be said that whatever X may be, A is a lover of X, if X is a servant of B, or A is a lover of whatever is a servant of B. This is called relative progressive involution. Polyadic relatives are capable of other modes of composition. Thus, it may be said that anything whatever, X, being taken, something Y exists, such that A praises X to Y while X maligns Y to B; that is, A praises everybody to somebody maligned by him to B. Or we can say that there is something, Y, such that, whatever X may be, A praises X to Y while X maligns Y to B; or, A praises everybody to somebody whom everybody maligns to B.

641. Deductive logic can really not be understood without the study of the logic of relatives, which corrects innumerable serious errors into which not merely logicians, but people who never opened a logic-book, fall from confining their attention to non-relative logic. One such error is that demonstrative reasoning is something altogether unlike observation. But the intricate forms of inference of relative logic call for such studied scrutiny of the representations of the facts, which representations are of an iconic kind, in that they represent relations in the fact by analogous relations in the representation, that we cannot fail to remark that it is by observation of diagrams that the reasoning proceeds in such cases. We successively simplify them and are always able to remark that such observation is required, and that it is even thus, and not otherwise, that the conclusion of a simple syllogism is seen to follow from its premisses. Again, non-relative logic has given logicians the idea that deductive inference was a following out of a rigid rule, so that machines have been constructed to draw conclusions. But this conception is not borne out by relative logic. People commonly talk of the conclusion from a pair of premisses, as if there were but one inference to be drawn. But relative logic shows that from any proposition whatever, without a second, an endless series of necessary consequences can be deduced; and it very frequently happens that a number of distinct lines of inference may be taken, none leading into another. That this must be the case is indeed evident without going into the logic of relatives, from the vast multitude of theorems deducible from the few incomplex premisses of the theory of numbers. But ordinary logic has nothing but barren sorites to explain how this can be. Since Kant, especially, it has been customary to say that deduction only elicits what was implicitly thought in the premisses; and the famous distinction of analytical and synthetical judgments is based upon that notion. But the logic of relatives shows that this is not the case in any other sense than one which reduces it to an empty form of words. Matter entirely foreign to the premisses may appear in the conclusion. Moreover, so far is it from being true, as Kant would have it, that all reasoning is reasoning in Barbara, that that inference itself is discovered by the microscope of relatives to be resolvable into more than half a dozen distinct steps. In minor points the doctrines of ordinary logic are so constantly modified or reversed that it is no exaggeration to say that deductive logic is completely metamorphosed by the study of relatives.

642. One branch of deductive logic, of which from the nature of things ordinary logic could give no satisfactory account, relates to the vitally important matter of abstraction. Indeed, the student of ordinary logic naturally regards abstraction, or the passage from "the rose smells sweet" to "the rose has perfume," to be a quasi-grammatical matter, calling for little or no notice from the logician. The fact is, however, that almost every great step in mathematical reasoning derives its importance from the fact that it involves an abstraction. For by means of abstraction the transitory elements of thought, the {epea pteroenta}, are made substantive elements, as James terms them, {epea apteroenta}. †1 It thus becomes possible to study their relations and to apply to these relations discoveries already made respecting analogous relations. In this way, for example, operations become themselves the subjects of operations.

To take a most elementary example — from the idea of a particle moving, we pass to the idea of a particle describing a line. This line is then thought as moving, and so as generating a surface; and so the relations of surfaces become the subject of thought. An abstraction is an ens rationis whose being consists in the truth of an ordinary predication. A collection, or system, is an abstraction or abstract ens; and thus the whole doctrine of number is founded on the operation of abstraction. If we conceive an object to be a collective whole, but to be so in such a way that it has no part which is not itself a collective whole in the same way, then, if the collection is of the nature of a sorite, it is a general, whose parts are distinguished merely as having additional characters; but if the collection is a set, whose members have other relations to one another, it is a continuum. The logic of continua is the most important branch of the logic of relatives, and mathematics, especially geometrical topic, or topical geometry, has its developement retarded from the lack of a developed logic of continua.

643. Literature: relatives have, since Aristotle, been a recognized topic of logic. The first germ of the modern doctrine appears in a somewhat trivial remark of Robert Leslie Ellis. De Morgan did the first systematic work in his fourth memoir on the syllogism in 1860 (Cambridge Philosophical Transactions, x. 331-358); he here sketched out the theory of dyadic relations. C. S. Peirce, in 1870, †1 extended Boole's algebra so as to apply to them, and after many attempts produced a good general algebra of logic, together with another algebra specially adapted to dyadic relations (Studies in Logic, by members of the Johns Hopkins University, 1883, Note B, 187-203). †2 Schröder developed the last in a systematic manner (which brought out its glaring defect of involving hundreds of merely formal theorems without any significance, and some of them quite difficult) in the third volume of his Exakte Logik (1895). Schröder's work contains much else of great value. . . .

§9. Transposition †3

644. Transposition consists in transferring a term from the subject to the predicate, or the reverse, with no change in the character of the connection; as, No artists who are bankers are clever, No artists are clever bankers, No bankers are clever artists, None are at once artists and bankers and clever; or as All but a is b, All but b is a. Any proposition may be "transformed" into other exactly equivalent forms: e.g. the transformation may consist in the change from one sort of connection to another (change of copula, in the extended meaning of that term), as — to take a compound proposition as an example — It never rains but it pours = always either it pours or it does not rain, but this is not transposition.

645. Certain copulas permit transposition simply, with no variation in the quality of the term transposed (as in the instances just given); but with the non-symmetrical copulas there must be a change from positive to negative or the reverse (and, if the proposition is complex, from the conjunctive to the alternative combination and the reverse), if the change can be made at all: He who is an astronomer and un-devout is mad = Any astronomer is mad or devout = All are mad or devout or not astronomers. When both the whole subject and the whole predicate is transposed the change is commonly called contraposition if the copula is non-symmetrical (All a is b = All non-b is non-a; None but a is b = None but non-b is non-a), but simple conversion if it is symmetrical (No a is b = No b is a, Some a is b = Some b is a). The usual discussion in the logics of the doctrine of the equivalence of propositions is greatly simplified by taking this more general view of the subject.


cover
The Collected Papers of Charles Sanders Peirce. Electronic edition.
Volume 3: Exact Logic
Endmatter
Appendix: On NonionsE
<hide milestones><show milestones><hide table of contents><show table of contents>

Appendix: On NonionsE †1

646. Readers of Professor Sylvester's communication entitled Erratum in the last number of these Circulars have perhaps inferred that my conduct in the matter there referred to had been in fault. Professor Sylvester's Erratum relates to his "Word upon Nonions," printed in the Johns Hopkins University Circulars No. 17, p. 242. In that article appears this sentence: "These forms [i.e. a certain group of nine Forms belonging to the algebra of Nonions] can be derived from an algebra given by Mr. Charles S. Peirce, (Logic of Relatives, 1870)." The object of Professor Sylvester's "Erratum" would seem to be to say that this sentence was inserted by me in his proof-sheet without his knowledge or authority on the occasion of the proof being submitted to me to supply a reference, and to repudiate the sentence, because he "knows nothing whatever" of the fact stated. But I think this view of Professor Sylvester's meaning is refuted by simply citing the following testimony of Professor Sylvester himself, printed in the Johns Hopkins University Circulars, No. 15, p. 203.

"Mr. Sylvester mentioned . . . that . . . he had come upon a system of Nonions, the exact analogues of the Hamiltonian Quaternions . . . Mr. Charles S. Peirce, it should be stated, had to the certain knowledge of Mr. Sylvester arrived at the same result many years ago in connection with his theory of the logic of relatives; but whether the result had been published by Mr. Peirce, he was unable to say."

This being so, I think that on the occasion of Professor Sylvester's publishing these forms I was entitled to some mention, if I had already published them, and a fortiori if I had not. When the proof-sheet was put into my hands, the request made to me, by an oral message, was not simply to supply a reference but to correct a statement relating to my work in the body of the text. And I had no reason to suppose that having thus submitted his text to me, Professor Sylvester would omit to look at his proof-sheet after it left my hands to see whether or not he approved of such alteration as I might have proposed. At any rate, when from these causes Professor Sylvester's "Word upon Nonions" had been published with the above statement concerning me, would it have been too much to expect that he should take the trouble to refer to my memoir in order to see whether the statement was not after all true, before publicly protesting against it?

647. I will now explain what the system of Nonions consists in and how I have been concerned with it.

The calculus of Quaternions, one of the greatest of all mathematical discoveries, is a certain system of algebra applied to geometry. A quaternion is a four-dimensional quantity; that is to say, its value cannot be precisely expressed without the use of a set of four numbers. It is much as if a geographical position should be expressed by a single algebraical letter; the value of this letter could only be defined by the use of two numbers, say the Latitude and Longitude. There are various ways in which a quaternion may be conceived to be measured and various different sets of four numbers by which its value may be defined. Of all these modes, Hamilton, the author of the algebra, selected one as the standard. Namely, he conceived the general quaternion q to be put into the form

q = xi + yj + zk + w,

where x, y, z, w, are four ordinary numbers, while i, j, k, are peculiar units, subject to singular laws of multiplication. For ij = -ji, the order of the factors being material, as shown in this multiplication table, where the first factor is entered at the side, the second at the top, and the product is found in the body of the table.

1 i j k
1 1 i j k
i i -1 k -j
j j -k -1 i
k k j -i -1

As long as x, y, z, and w in Hamilton's standard tetranomial form are confined to being real numbers, as he usually supposed them to be, no simpler or more advantageous form of conceiving the measurement of a quaternion can be found. But my father, Benjamin Peirce, †1 made the profound, original, and pregnant discovery that when x, y, z, w are permitted to be imaginaries, there is another very different and preferable system of measurement of a quaternion. Namely, he showed that the general quaternion, q, can be put into the form

q = xi + yj + zk + w l,

where x, y, z, w, are real or imaginary numbers, while i, j, k, l are peculiar units whose multiplication obeys this table.

i j k l
i i j 0 0
j 0 0 i j
k k l 0 0
l 0 0 k l

A quaternion does not cease to be a quaternion by being measured upon one system rather than another. Any quantity belonging to the algebra is a quaternion; the algebra itself is "quaternions." The usual formulae of the calculus have no reference to any tetranomial form, and such a form might even be dispensed with altogether.

While my father was making his investigations in multiple algebra I was, in my humble way, studying the logic of relatives and an algebraic notation for it; and in the ninth volume of the Memoirs of the American Academy of Arts and Sciences, †2 appeared my first paper upon the subject. In this memoir, I was led, from logical considerations that are patent to those who read it, to endeavor to put the general expression of any linear associative algebra into a certain form; namely as a linear expression in certain units which I wrote thus:

(u1:u1) (u1:u2) (u1:u3), etc.,
(u2:u1) (u2:u2) (u2:u3), etc.,
(u3:u1) (u3:u2) (u3:u3), etc.,
etc. etc. etc.

These forms, in their multiplication, follow these rules:

(ua:ub) (ub:uc) = (ua:uc) (ua:ub) (uc:ud) = 0.

I said, "I can assert, upon reasonable inductive evidence, that all such algebras can be interpreted on the principles of the present notation in the same way," †1 and consequently can be put into this form. I afterwards published a proof of this. †2 I added that this amounted to saying that "all such algebras are complications and modifications of the . . . Hamilton's quaternions." †3 What I meant by this appears plainly in the memoir. It is that any algebra that can be put into the form proposed by me is thereby referred to an algebra of a certain class (afterwards named quadrates by Professor Clifford) which present so close an analogy with quaternions that they may all be considered as mere complications of that algebra. Of these algebras, I gave as an example, the multiplication table of that one which Professor Clifford afterward named nonions. †P1 This is the passage: †4

It will be seen that the system of nonions is not a group but an algebra; that just as the word "quaternion" is not restricted to the three perpendicular vectors and unity, so a nonion is any quantity of this nine-fold algebra.

So much was published by me in 1870; and it then occurred either to my father or to me (probably in conversing together) that since this algebra was thus shown (through his form of quaternions) to be the strict analogue of quaternions, there ought to be a form of it analogous to Hamilton's standard tetranomial form of quaternions. That form, either he or I certainly found. I cannot remember, after so many years, which first looked for it; whichever did must have found it at once. I cannot tell what his method of search would have been, but I can show what my own must have been. The following course of reasoning was so obtrusive that I could not have missed it.

Hamilton's form of quaternions presents a group of four square-roots of unity. Are there, then, in nonions, nine independent cube-roots of unity, forming a group? Now, unity upon my system of notation was written thus:

(u1:u1) + (u2:u2) + (u3:u3).

Two independent cube-roots of this suggest themselves at once; they are

(u1:u2) + (u2:u3) + (u3:u1)

(u3:u2) + (u2:u1) + (u1:u3).

In fact these are hinted at in my memoir, p. 53.[?] Then, it must have immediately occurred to me, from the most familiar properties of the imaginary roots of unity, that instead of the coefficients

1, 1, 1,

I might substitute

1, g, g2

or

1, g2 g

where g is an imaginary cube-root of unity. The nine cube-roots of unity thus obtained are obviously independent and obviously form a group. Thus the problem is solved by a method applicable to any other quadrate.

648. My father, with his strong partiality for my performances, talked a good deal about the algebra of nonions in general and these forms in particular; and they became rather widely known as mine. Yet it is clear that the only real merit in the discovery lay in my father's transformation of quaternions. In 1875, when I was in Germany, my father wrote to me that he was going to print a miscellaneous paper on multiple algebra and he wished to have it accompanied by a paper by me, giving an account of what I had found out. I wrote such a paper, and sent it to him; but somehow all but the first few pages of the manuscript were lost, a circumstance I never discovered till I saw the part that had reached him (and which he took for the whole) in print. I did not afterward publish the matter, because I did not attach much importance to it, and because I thought that too much had been made, already, of the very simple things I had done.

I here close the narrative. The priority of publication of the particular group referred to belongs to Professor Sylvester. But most readers will agree that he could not have desired to print it without making any allusion to my work, and that to say the group could be derived from my algebra was not too much.


Endnotes

3

†1 Proceedings of the American Academy of Arts and Sciences, vol. 7, pp. 250-61, March 1867.

3

†2 I.e., a inline image b is the class of those things which are a not-b, b not-a or both a and b.

4

†1 I.e., if a,b = 0, a + b = a inline image b. Logical addition allows conjunction, arithmetic addition is exclusive.

4

†2 (2) and (5) embody the law of tautology — one of the features which distinguish the Boolian from the ordinary arithmetical calculus, limiting it to the numbers 1 and 0.

5

†1 I.e., if x is a minimum, a inline image b inline image a-b inline image a-(a,b); if a maximum, a inline image b inline image ab + a-b inline image a.

5

†2 I.e., Class b must be contained in Class a; b may of course be null.

5

†P1 So that, for example, ā denotes not-a. [Elec. ed. note: The tilde (~) is used to mark negation in this electronic edition rather than a line above the symbol(s) negated.]

5

†3 I.e., the class a inline image b is equal to the indeterminate class of that which is both a and b, plus the class of that which is a but not b, under the condition that there are no b's which are not a.

6

†1 As [0 inline image (x inline image x)] inline image [0 inline image x inline image x], 0 added to a class is the class. 0 represents the minimum which results from subtracting a class from itself.

6

†2 If x is a minimum, a;b=a,b=a; if a maximum, a;b, inline image ā, + ab.

6

†3 The class a;b is equal to the class of that which is both a and b plus the indeterminate class of what is neither a nor b on the condition that there are no a's which are not b.

6

†4 As a:b represents the maximum or upper limit of a;b (see 8) unity represents the maximum which results from dividing any class by itself. As (1 inline image x;x) inline image (x,1 inline image x) the result of multiplying a class by 1 is the class.

6

†5 As a:b inline image a + , 0:x inline image (0 + inline image ).

7

†1 See Lewis' Survey of Symbolic Logic, pp. 58-67, 82 and 132-174 for a very clear presentation and a development of these "transformations."

7

†2 f(x) = x,a + ,b.

7

†P1 The proof offered for this is fallacious inasmuch as i and j have not been proved to be independent of x.—1870. [Peirce follows this remark with a proof which is too long and of insufficient interest to be reproduced].

7

†P2 Identity (18') is reducible to (18) by development by second member by (18).—1870.

8

†P1 a;b, c must always be taken as (a;b), c, not as a;(b, c).

8

†1 Laws of Thought, vol. 2, p. 87ff.

9

†1 Ibid., p. 10.

9

†2 I.e., from the logical relations of class identity in extension to the "arithmetical" relations of numerical equality. Cf. 44.

9

†3 'Arithmetical' multiplication is represented by juxtaposing the terms.

13

†1 Cf. 2.674.

15

†1 Logic of Chance, ch. 9, section 24.

15

†P1 See a notice, Venn's Logic of Chance, in the North American Review for July 1867 [vol. 9].

16

†1 Proceedings of the American Academy of Arts and Sciences, vol. 7, pp. 402-412, September 1867.

16

†2 No other part seems to have been written.

16

†3 See Paper No. I.

17

†1 a and b are independent if they are summations of terms (4 and 5) each of whose members is distinct (1 and 2), so that there is a class of the terms of a and b together (3) and a term in a has a corresponding member in b (6). There are as many members of a,b as there are combinations of a member of a with one of b. Cf. 33.

18

†1 —because in transfinite arithmetic finite quantities can be added to infinites without affecting the total — ℵ + x = ℵ + y = ℵ where x and y are finite and ℵ is the smallest transfinite cardinal.

20

†1 Originally m.

20

†2 Originally u.

23

†1 See 4n.

23

†2 See 2.467. Cf. Paul Weiss, Erkenntnes Bd. 2, H. 4, S. 242-8 where it is shown that all the logical propositions are variations of some such form as PQ+P̅Q+P+~(PQ)

24

†1 The ideas here presented for the derivation of arithmetic from logic are somewhat similar to those employed in the Principia Mathematica, Whitehead and Russell, vol 2, section A (1912).

24

†2 See vol. 2, bk. II, ch. 5. for a discussion of these terms.

24

†3 This leads to a somewhat similar definition of a cardinal number as that given in the Principia Mathematica. But see 4.333 where this paper is characterized as being the worst Peirce ever published.

25

†P1 Thus, in one point of view, identity is a species of equality, and, in another, the reverse is the case. This is because the Being of the copula may be considered on the one hand (with De Morgan [Formal Logic, p. 59]) as a special description of "inconvertible, transitive relation," while, on the other hand, all relation may be considered as a special determination of Being. If a Hegelian should be disposed to see a contradiction here, an accurate analysis of the matter will show him that it is only a verbal one.

25

†1 Cf. Principia Mathematica, vol. 2, section A.

27

†1 Memoirs of the American Academy, vol. 9, pp. 317-78 (1870). Reprinted separately by Welch, Bigelow and Company, Cambridge, Mass., 1870, pp. 1-62. "In 1870 I made a contribution to this subject [logic] which nobody who masters the subject can deny was the most important excepting Boole's original work that ever has been made." — From the "Lowell Lectures," 1903.

27

†2 "On the Syllogism No. IV, and on the Logic of Relations," pp. 331-58, dated 1859.

28

†P1 I use the sign ⤙ in place of ≦. My reasons for not liking the latter sign are that it cannot be written rapidly enough, and that it seems to represent the relation it expresses as being compounded of two others which in reality are complications of this. It is universally admitted that a higher conception is logically more simple than a lower one under it. Whence it follows from the relations of extension and comprehension, that in any state of information a broader concept is more simple than a narrower one included under it. Now all equality is inclusion in, but the converse is not true; hence inclusion in is a wider concept than equality, and therefore logically a simpler one. On the same principle, inclusion is also simpler than being less than. The sign ≦ seems to involve a definition by enumeration; and such a definition offends against the laws of definition.

29

†P1 I write a comma below the sign of addition, except when (as is the case in ordinary algebra) the corresponding inverse operation (subtraction) is determinative, [i.e., except when the addition is arithmetical.]

29

†1 I.e., arithmetical.

29

†2 See 27.

29

†3 Cf. 55, 69f.

29

†4 I.e., "non-relative," or what was before called "logical."

29

†5 Cf. 73, 74n.

30

†1 I.e., arithmetical. See 75.

30

†2 The symbolism of the earlier papers is here slightly modified: the simple conjunction of terms now represents relative, instead of arithmetical, multiplication, and the dot is introduced to represent arithmetical multiplication. The comma, however, is still retained for logical multiplication.

30

†3 See 36.

30

†4 Cf. 71 and 72.

30

†P1 In the notation of quaternions Hamilton has assumed

(xy)z=x(zy) instead of (xy)z=x(yz),
although it appears to make but little difference which he takes. Perhaps we should assume two involutions, so that
(xy)z=x(yz), z(yx)=(zy)x.
But in this paper only the former of these is required. [See 113ff. for the latter.]
31

†1 See 77.

31

†2 I.e., logical; see 10.

31

†3 See 10 (24).

33

†1 The symbol inline image represents the base of Naperian logarithms, the symbol inline image represents π and inline image represents the square root of the negative in Benjamin Peirce's Linear Associative Algebras, §15ff (1870).

34

†1 Cf. the discussion on the categories in vol. 1, bk. III, and on signs in vol. 2, bk. II.

34

†2 Cf. 421 and 1.347.

34

†3 Cf. 69n.

35

†1 "On the Structure of the Syllogism," Section 1, Cambridge Philosophical Transactions, vol. 8 (1846); Formal Logic, p. 37 (1847).

35

†2 Cf. 2.518ff.

35

†3 Laws of Thought, p. 27.

35

†P1 According to De Morgan, Formal Logic, p. 334. De Morgan refers to the first edition of Drobisch's Logic. The third edition contains nothing of the sort.

36

†1 Cf. 42-44. A class is the extension of a collection; see e.g., 537n.

36

†2 Op. cit., p. 33.

37

†1 In 3.

37

†2 Ch. 6, §63; ch. 15, §177ff.

37

†P1 In another book [Substitution of Similars (1869) and subsequent works] he uses the sign ·|· instead of +.

38

†1 Cf. 82.

39

†1 h is a variable designating an unspecified H. See 84, 94, 111.

40

†1 See 73.

41

†1 In 55.

41

†2 l, s is the logical product of l and s; ls, on the other hand, is the relative product of l and s. See 54n, 68.

42

†1 A servant of herself who is also a woman is the same as a servant of a woman?

43

†1 "Something" is whatever is identical with an undetermined thing.

43

†2 "Anything" is whatever is identical with itself.

43

†P1 It will often be convenient to speak of the whole operation of affixing a comma and then multiplying, as a commutative multiplication, the sign for which is the comma. But though this is allowable, we shall fall into confusion at once if we ever forget that in point of fact it is not a different multiplication, only it is multiplication by a relative whose meaning — or rather whose syntax — has been slightly altered; and that the comma is really the sign of this modification of the foregoing term.

44

†1 Op. cit., p. 243ff.

45

†1 This is more accurately read as: a servant of all those who are either men or women.

45

†2 Cf. J. N. Lambert. "Sechs Versuche einer Zeichenkunst in der Vernunftlehre"; in Logische u. Philosophische Abhandlungen, vol. 1, ed. J. Bernouilli, Berlin, (1782) for the use of this "Newtonian Formula" in an intensional logic of absolute terms.

46

†1 I.e., a servant and lover of every woman is a servant of every woman and a lover of every woman.

46

†P1 "The same" substituted for "an"; "some" for "a," "every" for "all" — ink correction on C. S. P.'s own copy; cf. 145.

47

†P1 "Follows from last because [it is] negative of 𝐛̅mw"—marginal note.

47

†1 See 145.

48

†1 This proposition is the source of the famous so-called paradoxes of material implication.

48

†2 This sentence seems to have been misplaced and should have appeared after (22) or (23).

50

†1 I.e., by (21) 0xax inline image 0x and as by (12) ax + 0x = ax, 0xax.

53

†1 On his own copy, Peirce substitutes the condition "x is an unlimited relative," for "x > 0."

54

†1 See 198.

55

†1 In his Pure Logic.

56

†1 Where c = not, this becomes the formula for contraposition. See 142 and 2.550. In 186, however, a different formula for contraposition is given.

57

†1 Cf. the definition of individuals in 611-13.

58

†P1 The absolute individual can not only not be realized in sense or thought, but cannot exist, properly speaking. For whatever lasts for any time, however short, is capable of logical division, because in that time it will undergo some change in its relations. But what does not exist for any time, however short, does not exist at all. All, therefore, that we perceive or think, or that exists, is general. So far there is truth in the doctrine of scholastic realism. But all that exists is infinitely determinate, and the infinitely determinate is the absolutely individual. This seems paradoxical, but the contradiction is easily resolved. That which exists is the object of a true conception. This conception may be made more determinate than any assignable conception; and therefore it is never so determinate that it is capable of no further determination.

59

†1 See 96ff.

59

†2 See 100ff.

59

†3 See 121ff.

60

†1 Π' signifies logical multiplication and Σ' signifies logical addition.

60

†2 A servant of every woman is a servant of a woman.

60

†3 A lover-of-a-servant of every woman is a lover-of-a-servant of a woman.

61

†1 See Lewis' Survey of Symbolic Logic, p. 87n for a proof of this theorem.

61

†2 A lover of every servant of all women is a lover of a servant of every woman.

61

†3 A lover of every servant-of-a-woman is to a woman a lover of all her servants.

61

†P1 lsw ⤙ lsw and lsw ⤙ (ls)w invariably holds—marginal note.

62

†1 On Peirce's copy a line was drawn through the vinculum in each of these cases with the comment, "Crossed are not universally true."

64

†1 I.e., 1-(xy).

64

†2 I.e., -(xy).

64

†3 I.e., -(x', x'') - (x', x''') etc.

64

†4 I.e., -(x', x'', x''', x'''').

65

†P1 It makes another resemblance between 1 and infinity that log 0 = -1.

67

†1 In 56.

68

†1 I.e., each correlate has only one relate though a given relate may have many correlates.

69

†1 I.e., as a lover of none but servants.

69

†2 A lover of all servants is not a non-lover of a servant,

69

†3 A lover of none but servants is not a lover of a non-servant. Cf. 116.

69

†4 Formal Logic, p. 341.

69

†5 Cf. ib., p. 343; see also 244.

70

†1 An x of none but y's is a non-x of all non-y's; i.e., xy = -x-y.

72

†1 Cf. (92), (93) and 332.

72

†2 Or: Whatever is a lover-of-a-servant to nothing but women . . . etc.

75

†1 Cf. 602ff.

78

†1 See Lewis, Survey of Symbolic Logic, p. 103, for a symbolic and analytic account of some of these propositions.

80

†P1 Linear Associative Algebras. By Benjamin Peirce. 4to, lithographed. Washington. 1870. [Published with notes and addenda by C. S. Peirce in The American Journal of Mathematics, vol. 4, pp. 97-229 (1881); see paper no. VIII of this volume.]

81

†1 Ibid., (1612).

81

†2 See Appendix to this volume.

82

†1 By (161).

83

†1 In his "New Elements of Geometry With a Complete Theory of Parallels," Gelehrte Schriften der Universität, Kasan, 1836-1838.

84

†P1 The researches of Lobatchewsky furnish no solution of the question concerning the apriority of space. For though he has shown that it is conceivable that space should have such properties that two lines might be in a plane and inclined to one another without ever meeting, however far produced, yet he has not shown that the facts implied in that supposition are inconsistent with supposing space to retain its present [Euclidean] nature and the properties only of the things in it to change. For example, in Lobatchewsky's geometry a star at an infinite distance has a finite parallax. But suppose space to have its present properties, and suppose that there were one point in the universe towards which anything being moved should expand, and away from which being moved should contract. Then this expansion and contraction might obey such a law that a star, the parallax of which was finite, should be at an infinite distance measured by the number of times a yard-stick must be laid down to measure off that distance. I have not seen Beltrami's investigations, ["Saggio di interpretazione della geometrica non-euclidea," Giorn. di Matem., 6, 1868.] but I understand that they do show that something of this sort is possible. Thus, it may be that, make what suppositions you will concerning phenomena, they can always be reconciled to our present geometry or be shown to involve implicit contradictions. If this is so—and whether it is or not is a completely open question—then the principles of geometry are necessary, and do not result from the specialities of any object cognized, but from the conditions of cognition in general. In speaking of the conditions of cognition, in general, I have in view no psychological conception, but only a distinction between principles which, if the facts should present a sufficient difficulty, I may always logically doubt, and principles which it can be shown cannot become open to doubt from any difficulty in my facts, as long as they continue to be supposed in all logical procedure.

But, waiving this point, Lobatchewsky's conclusions do not positively overthrow the hypothesis that space is a priori. For he has only shown that a certain proposition, not usually believed to be axiomatical, is conceivably false. That people may be doubtful or even mistaken about a priori truth does not destroy all important practical distinction between the two kinds of necessity. It may be said that if Lobatchewsky's geometry is the true one, then space involves an arbitrary constant, which value cannot be given a priori. This may be; but it may be that the general properties of space, with the general fact that there is such a constant, are a priori, while the value of the constant is only empirically determined.

It appears to me plain that no geometrical speculations will settle the philosophy of space, which is a logical question. If space is a priori, I believe that it is in some recondite way involved in the logic of relatives.

85

†1 Cf. 225 and 585.

86

†P1 "As we speak of self-loving, etc., the former of these classes should be called self-relatives" — marginal note.

86

†1 Cf. 230.

86

†P2 "The idea of a copula is different. These should be called assimilative." — Marginal note. See 592.

86

†2 Op. cit., §25; see also 593.

87

†P1 "If such reciprocation is admissible but not necessary they may be called reciprocal."—Marginal note.

87

†P2 "Quædam sunt relatione equiparantiæ, quædam disquiparantiæ. Primæ sunt relationes similium nominum, secundæ relationes dissimilium nominum. Exemplum primi est quando idem nomen ponitur in recto et in obliquo, sicut simile simili est simile.... Exemplum secundi est quando unum nomen ponitur in recto sed aliud in obliquo, sicut pater est filii pater et non oportet quod sit patris pater." Ockham Quodlibetum 6, qu. 20. See also his Summa Logices, pars 1, cap. 52. "Relativa equiparantiæ: quæ sunt synonyma cum suis correlativis.... Relativa diquiparantiæ: quæ non sunt synonyma cum suis correlativis." Pschlacher in Petr. Hisp. The same definitions substantially may be found in many late mediæval logics.

87

†1 Formal Logic, p. 345

87

†P3 "Instead of self-relatives, better concurrents."—Marginal note.

87

†P4 "Insert 'only' between contains and 'elements'; for 'non-cyclic' substitute 'acyclic.'" — Marginal note.

87

†2 Cf. 233.

87

†P5 "Insert 'only' between contains and 'elements'; for 'non-cyclic' substitute 'acyclic.'" — Marginal note.

88

†1 Op. cit., p. 346.

88

†2 For if R is transitive then aRb.bRcaRc; if R is also equiparant then aRb.bRa is true and aRa is a necessary consequent.

88

†P1 "Should be called concatenated" — marginal note.

89

†P1 "Duplex est relatio: scilicet rationis et realis. Unde relatio rationis est quae fit per actum comparativum intellectus, ut sunt secundæ intentiones; sed relatio realis est duplex, scilicet aptitudinalis et actualis. Aptitudinalis est quæ non requirit terminum actu existere sed solum in aptitudine; cujusmodi sunt omnes propriæ passiones, omnes aptitudines, et omnes inclinationes; et tales sunt in illo prædicamento reductive in quo sunt illa quorum sunt propriæ passiones. Sed relatio actualis est duplex, scilicet, intrinsecus adveniens, et extrinsecus adveniens. Intrinsecus adveniens est quae necessario ponitur positis extremis in quacunque etiam distantia ponantur, ut similitudo, paternitas, equalitas. Extrinsecus adveniens est quæ necessario non ponitur, positis extremis, sed requiritur debita approximatio extremorum; cujusmodi sunt sex ultima praedicamenta, scilicet, actio, passio, quando, ubi, situs, et habitus." Tartaretus.

89

†1 Cf. 8 and 81 (25), (26.)

89

†2 These two together contain De Morgan's Theorem to the effect that the negative of a logical product is the logical sum of the negatives of its factors; and that the negative of a sum (z=not) is the product of the negatives of the summands.

90

†1 Laws of Thought, p. 62.

93

†1 This should be 1(1-B) ⤙ 1(1-A)

94

†1 In 146.

96

†1 The second half of this equation should be: inline image13inline imagexyz.

98

†1 In 112.

99

†1 Proceedings of the American Academy of Arts and Sciences, pp. 392-94, vol. 10, (1875).

99

†2 See 294 for another approach to this same problem.

100

†1 See Benjamin Peirce's Linear Associative Algebras, op. cit., p. 195.

100

†2 Ibid., p. 188.

101

†1 Ibid., p. 209.

101

†2 Ibid., p. 202.

102

†1 Proceedings of the American Academy of Arts and Sciences, vol. 13, pp. 115-16, (1877).

102

†2 "Die mechanik nach den Principien der Ausdehnungslehre," Bd. 12, H. 2, S. 222-40, (1877).

103

†1 Originally 'vectors'; corrected by Peirce.

103

†2 Originally 'relation'; corrected by Peirce.

104

†1 American Journal of Mathematics, vol. 3, pp. 15-57 (1880), with Peirce's marginal corrections, and the printed corrections of September 15, 1880, in which he says, "The manuscript left my hands in April last before I had seen several important publications — Mr. McColl's third paper, Prof. Wundt's Logik, etc."

104

†2 The editors have changed 'Chapter' to 'Part.'

104

†P1 "The whole of these two parts is bad, first, because it does not treat the subject from the point of view of pure mathematics, as it should have done; and second because the fundamental propositions are not made out. I follow too much in the footsteps of ordinary numerical algebra, and the sketch of the algebra of the copula is very insufficient." — from the Lowell Lectures, 1903.

105

†1 Cf. 2.146, 2.148.

106

†P1 Deductive logic, perhaps, does not involve the principle that there is any special character in the peripheral excitation but only that reasoning proceeds by habits that are consistent. Deductive — consistency of thought with itself. Inductive —consistency of the world (Uniformity of Nature). — marginal note, c. 1882.

106

†1 Cf. vol. 2, bk. III, ch. 1, §§1, 2, and ch. 2, Part I.

107

†P1 Though the leading principle itself is not present to the mind, we are generally conscious of inferring on some general principle. [Cf. 2.186 ff.]

108

†P1 This dash was used by Boole, but not over other than class-signs.

111

†P1 The general doctrine of this section is contained in my paper, On the Natural Classification of Arguments, 1867 [vol. 2, Bk. III, ch. 2].

111

†P2 There is a difference of opinion among logicians as to whether ⤙ or = is the simpler relation. But in my paper on the Logic of Relatives [47n.], I have strictly demonstrated that the preference must be given to ⤙ in this respect. The term simpler has an exact meaning in logic; it means that whose logical depth is smaller; that is, if one conception implies another, but not the reverse, then the latter is said to be the simpler. Now to say that A = B implies that A ⤙ B, but not conversely. Ergo, etc. It is to no purpose to reply that A ⤙ B implies A= (A that is B); it would be equally relevant to say that A ⤙ B implies A = A. Consider an analogous case. Logical sequence is a simpler conception than causal sequence, because every causal sequence is a logical sequence but not every logical sequence is a causal sequence; and it is no reply to this to say that a logical sequence between two facts implies a causal sequence between some two facts whether the same or different. The idea that = is a very simple relation is probably due to the fact that the discovery of such a relation teaches us that instead of two objects we have only one, so that it simplifies our conception of the universe. On this account the existence of such a relation is an important fact to learn; in fact, it has the sum of the importances of the two facts of which it is compounded. It frequently happens that it is more convenient to treat the propositions A ⤙ B and B ⤙ A together in their form A = B; but it also frequently happens that it is more convenient to treat them separately. Even in geometry we can see that to say that two figures A and B are equal is to say that when they are properly put together A will cover B and B will cover A; and it is generally necessary to examine these facts separately. So, in comparing the numbers of two lots of objects, we set them over against one another, each to each, and observe that for every one of the lot A there is one of the lot B, and for every one of the lot B there is one of the lot A.

In logic, our great object is to analyse all the operations of reason and reduce them to their ultimate elements; and to make a calculus of reasoning is a subsidiary object. Accordingly, it is more philosophical to use the copula ⤙ apart from all considerations of convenience. Besides, this copula is intimately related to our natural logical and metaphysical ideas; and it is one of the chief purposes of logic to show what validity those ideas have. Moreover, it will be seen further on that the more analytical copula does in point of fact give rise to the easiest method of solving problems of logic.

112

†1 I.e., -(A = B).

112

†2 "On the Structure of the Syllogism, and on the Application of the Theory of Probabilities to Questions of Argument and Authority." Transactions, Cambridge Philosophical Society, vol. 8, pp. 379-408, (1849). The paper was read and dated 1846.

112

†3 I.e., it is false that A ⤙ B.

113

†P1 In consequence of the identification in question, in S ⤙ P, I speak of S indifferently as subject, antecedent, or premiss, and of P as predicate, consequent, or conclusion.

113

†1 See Laws of Thought, p. 62f.

113

†2 See 138.

113

†P2 Equally unsuccessful is Mr. Jevons' attempt to overcome the difficulty by omitting particular propositions, 'because we can always substitute for it [some] more definite expressions if we like.' The same reason might be alleged for neglecting the consideration of not. But in fact the form A ~⤙ B is required to enable us to simply deny A ⤙ B.

113

†3 To express such a particular proposition disjunctively, change the quantity and quality of the antecedent and the consequent and deny their disjunction. Cf. 196.

114

†1 See 178.

115

†1 See vol. 2, Bk. III, ch. 1, §3 for a later analysis of this quadrant.

115

†2 The readings in this column are not precise.

115

†3 The terms in this column are taken from De Morgan's later papers.

115

†4 Cf. vol. 2, Bk. III, ch. 3, §3.

116

†P1 In this connection see De Morgan, "On the Syllogism," No. V., 1862. [Transactions, Cambridge Philosophical Society, vol. 4, p. 467, (1864), read and dated 1863.]

116

†P2 Mr. Hugh McColl (Calculus of Equivalent Statements, Second Paper, 1878, [Proceedings, London Mathematical Society, vol. 9, p. 183 (1877)]), makes use of the sign of inclusion several times in the same proposition. He does not, however, give any of the formulæ of this section.

117

†P1 "On the Syllogism," No. II., 1850, [Transactions, Cambridge Philosophical Society, vol. 9, (1851), p. 104].

117

†P2 That the validity of syllogism is not deducible from the principles of identity, contradiction, and excluded middle, is capable of strict demonstration. The transitiveness of the copula is, however, implied in the identification of the copula-relation with illation, because illation is obviously transitive.

117

†P3 The conception of substitution (already involved in the mediæval doctrine of descent), as well as the word, was familiar to logicians before the publication of Mr. Jevons's Substitution of Similars. [see vol. 8] This book argues, however, not only that inference is substitution, but that it and induction in particular consist in the substitution of similars. This doctrine is allied to Mill's theory of induction.

117

†P4 This must have been in Boole's mind from the first. De Morgan ("On the Syllogism," No. II., 1850, p. 83) goes too far in saying that "what is called elimination in algebra is called inference in logic," if he means, as he seems to do, that all inference is elimination. [Cf. 2.442f.]

117

†1 See 91n.

119

†P1 De Morgan, Syllabus, 1860, p. 18.

120

†1 Cf. 2.597-8.

121

†P1 An oversight has here been committed. For from Ā = (A ⤙ x) follows not merely (16) but also (19), (20), and (21), and thus all the properties of the negative which concern syllogistic. But this does not affect the view taken of the subject, nor the division of the moods according to the properties of the negative on which they depend; for whatever is shown in the text to be deducible from Ā = (A ⤙ x) is in fact deducible from (16). — Sept., 1880.

121

†P2 Aristotle and De Morgan have particular conclusions from two universal premisses. These are all rendered illogical by the significations which I attach to ⤙ and ~⤙.

121

†1 P̅ and C̅ here represent some such forms as S ~⤙ M and S ~⤙ P.

123

†1 See 2.458.

124

†1 This should be: either (x ~⤙ y) or y.

125

†P1 The symbol 0 is used by Boole; the symbol ∞ replaces his 1, according to a suggestion in my Logic of Relatives, 1870 [88].

125

†P2 These forms of definition are original. The algebra of non-relative terms was given by Boole (Mathematical Analysis of Logic, 1847). Boole's addition was not the same as that in the text, for with him whatever was common to the two terms added was taken twice over in the sum. The operations in the text were given as complements of one another, and with appropriate symbols, by De Morgan ("On the Syllogism," No. III., 1858 [Cambridge Philosophical Transactions, vol. 10], p. 185). For addition, sum, parts, he uses aggregation, aggregate, aggregants; for multiplication, product, factors, he uses composition, compound, components. Mr. Jevons (Formal Logic [Pure Logic?], 1864)—I regret that I can only speak of this work from having read it many years ago, and therefore cannot be sure of doing it full justice—improved the algebra of Boole by substituting De Morgan's aggregation for Boole's addition. The present writer, not having seen either De Morgan's or Jevons's writings on the subject, again recommended the same change (On an Improvement in Boole's Calculus of Logic, 1867 [3]), and showed the perfect balance existing between the two operations. In another paper, published in 1870 [47], I introduced the sign of inclusion into the algebra.

In 1872, Robert Grassmann, brother of the author of the Ausdehnungslehre, published a work entitled 'Die Formenlehre oder Mathematik,' the second book of which gives an algebra of logic identical with that of Jevons. The very notation is reproduced, except that the universe is denoted by T instead of U, and a term is negatived by drawing a line over it, as by Boole, instead of by taking a type from the other case, as Jevons does. Grassmann also uses a sign equivalent to my ⤙. In his third book, he has other matter which he might have derived from my paper of 1870. Grassmann's treatment of the subject presents inequalities of strength; and most of his results had been anticipated. Professor Schröder, of Karlsruhe, in the spring of 1877, produced his Operationskreis des Logikkalkuls. He had seen the works of Boole and Grassmann, but not those of De Morgan, Jevons, and me. He gives a fine development of the algebra, adopting the addition of Jevons, and he exhibits the balance between + and X by printing the theorems in parallel columns, thus imitating a practice of the geometricians. Schröder gives an original, interesting, and commodious method of working with the algebra. Later in the same year, Mr. Hugh McColl, apparently having known nothing of logical algebra except from a jejune account of Boole's work in Bain's Logic, published several papers on a Calculus of Equivalent Statements, [Proceedings, London Mathematical Society, Series 1, vol. IX, pp. 177-186], the basis of which is nothing but the Boolian algebra, with Jevons's addition and a sign of inclusion. Mr. McColl adds an exceedingly ingenious application of this algebra to the transformation of definite integrals.

127

†P1 Remark that the proofs of the lettered propositions follow the enunciations — 1880.

127

†P2 Logic of Relatives (§4) gives a×ba. The other formulæ, equally obvious, I do not find anywhere.

127

†P3 The first of these given by Boole for his addition, was retained by Jevons in changing the addition. The second was first given by me (1867) [See 4].

128

†1 "It seems that (a+bc ⤙ (a×c)+(b×c) cannot be proved from the definitions. The propositions L are needed"—a marginal note prompted apparently by Schröder's criticisms in his Vorlesungen über die Algebra der Logik, Bd. 1, Kap. 6.

On February 14, 1904, Peirce wrote Prof. E. V. Huntington of Harvard as follows:

"Dear Mr. Huntington: Should you decide to print the proof of the distributive principle (and this would not only relieve me from a long procrastinated duty, but would have a certain value for exact logic, as removing the eclipse under which the method of developing the subject followed in my paper in vol. 3 has been obscured) I should feel that it was incumbent upon me, in decency, to explain its having been so long withheld. The truth is that the paper aforesaid was written during leisure hours gained to me by my being shut up with a severe influenza. In writing it, I omitted the proof, as there said, because it was 'too tedious' and because it seemed to me very obvious. Nevertheless, when Doctor Schröder questioned its possibility, I found myself unable to reproduce it, and so concluded that it was to be added to the list of blunders, due to the grippe, with which that paper abounds—a conclusion that was strengthened when Schröder thought he demonstrated the indemonstrability of the law of distributiveness. (I must confess that I never carefully examined his proof, having my table loaded with logical books for the perusal of which life was not long enough.) It was not until many years afterwards that, looking over my papers of 1880 for a different purpose, I stumbled upon this proof written out in full for the press, though it was eventually cut out, and, at first, I was inclined to think that it employed the principle that all existence is individual, which my method, in the paper in question, did not permit me to employ at that stage. I venture to opine that it fully vindicates my characterisation of it as 'too tedious'. But this is how I have a new apology to make to exact logicians."

This letter and the proof were used by Professor Huntington in his "Sets of Independent Postulates for the Algebra of Logic," Transactions, American Mathematical Society, vol. 5, p. 300n. (1904), and in proof of proposition 22a. The proof is also to be found in Lewis's Survey of Symbolic Logic, p. 128 (5.5). A more elegant proof is to be found in the Principia Mathematica. See 384n.

130

†1 Cf. Mrs. Ladd-Franklin's Antilogism in her "On the Algebra of Logic," Studies in Logic, edited by C. S. Peirce, Little, Brown & Co., Boston, 1883.

132

†1 In 9.

132

†2 In 12.

132

†3 These two embody De Morgan's principle of duality.

132

†4 This should be: ā + ⤙ ~(a×b).

135

†1 I.e., {(x+y) ⤙ z }.

135

†2 I.e., {x ⤙ (y×z)}.

135

†3 I.e., {(x+y) ~⤙ z}.

135

†4 I.e., {x ~⤙ (y×z)}.

136

†1 The Laws of Thought, pp. 146-9.

139

†1 Cf. 4.121.

139

†2 See 93, 2.646 and 4.121-22.

139

†3 Cf. 4.118-9.

140

†1 Cf. 2.356.

140

†P1 In my Logic of Relatives, 1870 [§6], I have used this expression ['simple relatives'] to designate what I now call dual relatives.

142

†1 Four lines of formula are here deleted, in accordance with Peirce's subsequent marginal comment. They involved the invalid use of the law of association in connection with triple relatives.

142

†2 Σ(A:B) represents the logical sum of individual relatives; Π(α:β), represents the logical product of simple relatives.

143

†1 LM indicates the application of L on M. L says that any formula of the form (1:2):3 is to be changed to (3:1):2. As M means that (1:2):3 is to be changed to (2:3):1, the application of L on M yields the original. The rest of the formulæ are to be understood in a similar way.

143

†2 I.e., I results by substituting a for b, b for a and c for c in the equation for b; and so on with the rest.

144

†1 See 136an.

144

†2 See 136dn.

145

†P1 The relative 0 ought to be considered as at once a concurrent and an alio-relative, and the relative ∞ as at once the negative of a concurrent and the negative of an alio-relative. [Cf. 585-6.] The statements in the text require to be modified to this extent. [This note, apparently a correction made after receiving proofs, was published at the end of the original paper.]

146

†P1 These numbers are every fifth of the series:
1̅*,- 7,10,-3,2̅*, -5,5,2,3, 5̅*,0,5,7,10,1̅5*, 15,20,27,37,~5̅2̅*, where ux+ux+3 = ux+4. But only holds up to 203 and is therefore valueless.— marginal note.

147

†1 See 136cn.

147

†2 See 136hn.

150

†P1 The first three of these were studied by De Morgan ("On the Syllogism," No. IV.); the last is new. The above names for the first three (except the adjective external suggested by Grassmann's operation) are given in my Logic of Relatives.

151

†P1 A similar table is given by De Morgan. Of course, it lacks the symmetry of this, because he had not the fourth operation. [Cf. 112.]

158

†1 The American Journal of Mathematics, vol. 4, pp. 85-95 (1881), as subsequently corrected.

158

†2 J. S. Mill, e.g. in Logic, bk. II, ch. 6, §2-3.

159

†P1 For example, in the ordinary algebra of imaginaries two quantities may both result from the addition of quantities of the form a2+b2i to the same quantity without either being in this relation to the other.

159

†1 But see 4.121.

160

†1 See 564.

160

†2 I.e., semi-limited and infinite; see 4.107.

160

†3 Cf. 4.150ff.

166

†1 I.e., c is a one-one relation.

166

†2 is the converse of c.

166

†3 Originally 'c'd with.'

166

†4 Originally 'c'd by.'

166

†5 I.e., counting involves the establishment of a one-one correlation between the members of a given class and the natural numbers.

168

†P1 This long proof is quite unnecessary. The whole thing depends not on Fermat's mode of reasoning but on De Morgan's. — marginal note. [I.e., it depends, not on mathematical induction but on the syllogism of transposed quantity.]

169

†P1 It may be remarked that when we reason that a certain proposition, if false of any number, is false of some smaller number, and since there is no number (in a semi-limited system) smaller than every number, the proposition must be true, our reasoning is a mere logical transformation of the reasoning that a proposition, if true for n, is true for 1+n, and that it is true for 1.

170

†1 This mode of reasoning was uncovered by De Morgan and called the syllogism of transposed quantity. See 402 and 4.103f.

171

†1 The American Journal of Mathematics, vol. 4, pp. 221-29, (1881), an addendum to Benjamin Peirce's Linear Associative Algebras published posthumously in the same volume with notes by C. S. P. These notes throw considerable light on the significance and relationship of algebra to the logic of relatives.

171

†P1 I have used a11 etc., in place of the a1, etc., used by my father in his text.

171

†P2 Any one of them multiplied by 0 gives 0.

172

†P1 If B = 0, of course the result is 0.

173

†P1 A brief proof of this theorem, perhaps essentially the same as the above, was published by me in the Proceedings of the American Academy of Arts and Sciences, for May 11, 1875. [150-51.]

173

†1 Paper No. IX.

173

†2 The reference is to the Linear Associative Algebras, p. 188.

176

†P1 The idempotent basis having been shown to be arithmetical unity, we are free to use the letter i to denote another unit.

180

†1 Dated, Baltimore, January 7, 1882, pp. 1-6, with a postscript of January 16, 1882. To judge from a search through the technical journals and from Peirce's reference in 294, this paper was privately printed.

180

†P1 I have usually restricted the coefficients to one or other of two values; but the more general view was distinctly recognized in my paper of 1870.

181

†1 Cf. 133.

182

†1 This should be: (1̅)ii = 0.

182

†2 Cf. 242.

182

†P1 "On De Morgan's Extension of the Algebraic Processes," American Journal of Mathematics, vol. 3, no. 3.

183

†1 See 297-305.

185

†1 150-51.

185

†2 Cf. the analysis of signs, 2.274 and in the present volume, 359f.

185

†3 There is no paper by this title known to have been published by Sylvester, though a paper entitled 'Lectures on the Principles of Universal Algebra,' was published in the American Journal of Mathematics, vol. 6, pp. 270-86, (1884). A number of papers on algebra published in 1881-82 are to be found in the Collected Mathematical Papers of J. J. Sylvester, ed. by H. F. Baker, Cambridge University Press, vol. 3 (1904-12).

185

†P1 "Description of a Notation for the Logic of Relatives." Memoirs, American Academy of Arts and Sciences, vol. 9, 1870. [III.] "On the Application of Logical Analysis to Multiple Algebra." Proceedings of the same Academy, 1875, May 11. [IV.] "Note on Grassman's Calculus of Extension." Ibid. 1877, Oct. 10. [V.] "On the Algebra of Logic." American Journal of Mathematics, vol. 3. [VI.]

187

†1 Johns Hopkins University Circulars, No. 13, p. 179, (1882).

189

†1 Johns Hopkins University Circulars, No. 19, pp. 3-4, (1882), read before the University Mathematical Society, October 18, 1882.

189

†2 The first and second paragraph were originally transposed.

193

†1 The rest of this paper was added on October 30.

195

†1 Note B, pp. 187-203, Johns Hopkins Studies in Logic, ed. by C. S. Peirce, Little, Brown & Co., Boston, 1883.

197

†1 As this is not intended to exclude being a lover of a benefactor, but only being a non-lover of a non-benefactor, the following alternative expressions may be clearer: "lover of all non-benefactors"; "a non-lover only of benefactors," or "Either a lover of X or X is a benefactor." Relative addition is the denial of transaddition (243), so that the following are equivalent: -~(lb), -lb, l-b, lb, and -(lb). In 473 it is shown that $~l ⤙ b is also equivalent to the above.

197

†2 Cf. 118 (143) (144).

197

†3 Cf. 352.

197

†4 I.e., it is not exclusive.

198

†1 These two formulæ together show that lovers of servants or benefactors of servants are the same as lovers or benefactors of servants.

198

†2 These two formulæ together show that lovers and benefactors of all non-servants are the same as lovers of all non-servants and benefactors of all non-servants.

198

†3 Cf. 249, 1ii.

198

†4 Cf. 136 and 227.

199

†P1 Sometimes important. 1 ⤙ l + $ [i.e., identity implies to be either a lover or not loved by] and l,$ ⤙ 𝖓 [i.e., to love and not be loved by implies otherness] — marginal note.

200

†1 I.e., Identity implies to love all that is loved by.

200

†2 I.e., to love something that is not loved by implies otherness.

200

†3 "On a New Algebra of Logic," Studies in Logic, pp. 87, 88.

202

†1 2.521ff.

202

†2 The present paper was rewritten to serve as the thirteenth chapter of the Grand Logic of 1893-94, where the above six propositions are replaced by the following five:

$āp†β̅ Every a has every β
ă(p†β̅) Some a has all β
ăp†β̅ Every β is some a or other
($āp Some β is all a
$āpβ Some β is in some a;
and the various syllogisms are modified accordingly. There are no other changes of consequence. These are reduced to four by Schröder, Algebra der Logik, III, 1, 470.
202

†3 For a reduction in the number of these forms see Schröder, ibid., 470f.

204

†1 See 2.526n for definition of this term.

205

†1 Op. cit., p. 87.

206

†P1 The sums of 331.

207

†1 More clearly: For some i and k, i is a lover of all non-benefactors of k; or for some i and some k every j is such that either i loves j or j is a benefactor of k.

208

†1 This sentence is slightly different from the original, in accordance with Peirce's correction.

209

†1 No record of this communication has been found in the official Proceedings of the London Mathematical Society; nor has any letter from Mr. Schlötel come to hand in a search through Peirce's correspondence.

209

†2 Die Logik, neu bearbeitet, 1854.

210

†1 The American Journal of Mathematics, vol. 7, No. 2, pp. 180-202, (1885); reprinted pp. 1-23.

210

†2 See vol. 2, bk. II, for a detailed analysis of signs.

210

†3 More frequently called 'symbols'; the word 'token' is later (in 4.537) taken to apply to what in 2.245 is called a 'sinsign.'

211

†P1 See Proceedings, American Academy of Arts and Sciences, vol. 7, p. 294, May 14, 1867. [1.558.]

212

†1 See 4.356.

212

†2 "On the Diagrammatic and Mechanical Representations of Propositions and Reasoning." Philosophical Magazine, ser. 5, vol. 10, pp. 1-15 (1880).

212

†P1 Studies in Logic, by members of the Johns Hopkins University; Boston, Little, Brown and Co., 1883.

212

†3 Ibid., p. 74.

214

†1 If this proposition be added to the postulates of Boolean algebra and if the terms of that algebra be interpreted as propositions, a propositional calculus is secured. From an historical standpoint this is of tremendous significance.

216

†1 The Laws of Thought, pp. 47ff.

216

†2 I.e., ~().

216

†3 I.e., -(xy) or +.

216

†4 I.e., -()z= -{-(xz)-(yz)}

216

†5 I.e., -{-(xy)} = -(x̄~z)-().

217

†1 In paper No. VI, part I.

218

†1 Cf. 527.

220

†1 Strictly speaking, the modus ponens is expressed as {(xy).x} ⤙ y. The given proposition states that if x is true then provided 'x then y' is true, y is true.

220

†2 I.e. (xy) ⤙ {(yz) ⤙ (xz)}, which is often called the nota notæ or the dictum de omni. See 383, and vol. 2, Bk. III, ch. 4, 14, and cf. Joseph's An Introduction to Logic, p. 296n and 308n.

220

†3 This is not an exact reading of the given formula even as modified in the last note. The formula for this expression and thus for Barbara is: {(xy) ~⤙ (y ~⤙ z)} ⤙ (xz).

222

†P1 It is interesting to observe that this reasoning is dilemmatic. In fact, the dilemma involves the fifth icon. The dilemma was only introduced into logic from rhetoric by the humanists of the renaissance; and at that time logic was studied with so little accuracy that the peculiar nature of this mode of reasoning escaped notice. I was thus led to suppose that the whole non-relative logic was derivable from the principles of the ancient syllogistic, and this error {But it was not an error!!! See my original demonstration. — marginal note. [See 200n].] is involved in chapter 2 of my paper in the third volume of this Journal [No. VI]. My friend, Professor Schröder, detected the mistake and showed that the distributive formulæ

(x+y)zxz+yz

(x+z)(y+z) ⤙ xy+z

could not be deduced from syllogistic principles. [This matter is discussed at length by Schröder in his Vorlesungen über die Algebra der Logik, Bd. 1, 12, (1890)]. I had myself independently discovered and virtually stated the same thing. (Studies in Logic, p. 189 [331].) There is some disagreement as to the definition of the dilemma (see Keynes's excellent Formal Logic, p. 241); but the most useful definition would be a syllogism depending on the above distribution formulæ. The distribution formulæ

xz+yz ⤙ (x+y)z

xy+z ⤙ (x+z)(y+z)

are strictly syllogistic. DeMorgan's added moods are virtually dilemmatic depending on the principle of excluded middle.

225

†1 This should be: ~(+y)+z.

225

†2 This should be: +y.

226

†1 Op. cit., p. 80f.

227

†1 If the six vertices of a hexagon lie three and three on two straight lines, the three points of intersection of the opposite sides lie on a straight line.

227

†2 Op. cit., p. 79.

228

†P1 I will just remark, quite out of order, that the quantification may be made numerical; thus producing the numerically definite inferences of DeMorgan and Boole. Suppose at least 2/3 of the company have white neckties and at least 3/4 have dress coats. Let w mean 'he has a white necktie,' and d 'he has a dress coat.' Then, the two propositions are

(2/3)(w) and (3/4)(d).

These are to be multiplied together. But we must remember that xy is a mere abbreviation for ~(+), and must therefore write

~(~((2/3)w) + ~((3/4)d)).

Now ~((2/3)w) is the denial of (2/3)w, and this denial may be written (>(1/3)), or more than 1/3 of the universe (the company) have not white neckties. So ~((3/4)d) = (>(1/4)). The combined premisses thus become

~((>(1/3)) + (>(1/4)))
Now (>(1/3)) + (>(1/4)) gives May be (1/3 + 1/4)(+).
Thus we have May be (7/12)(+)
and this is (At least 5/12)(+),
which is the conclusion.
228

†1 This is the seventh icon?

228

†2 This is the eighth icon?

230

†1 See Lady's and Gentleman's Diary for 1850, p. 48 for the original presentation of 'Kirkman's School-Girls Problem.' See also W. W. R. Ball, Mathematical Recreations and Essays, Ch. IX, MacMillan & Co., 5 ed. (1911), where a reference is given to Benjamin Peirce's method of solution and his enunciation of a corresponding problem.

230

†2 Peirce defines a syntheme in the Century Dictionary, p. 6139, ed. 1889, as "a system of groups of objects comprising every one of a larger set just once, twice or other given number of times. The groups may be divided into subgroups subject to various conditions."

231

†1 Obviously a misprint for ΣiΠjxij ⤙ ΠjΣixij.

232

†1 See 403A and B at end of article.

232

†2 See 403C and M.

233

†1 See 403H.

235

†1 See 403I.

235

†2 See 403J.

236

†1 Cf. 21 (7).

236

†2 vαw.

236

†3 I.e., any two terms which are relates in any one-one correspondence to any term are identical; any two terms which are the correlates by one-one correspondence to any term are identical; and any case of terms related one to one is always a one-one correspondence.

237

†1 Transactions Cambridge Philosophical Society, vol. 10, pp. *355-*358, (1864).

237

†2 Ibid., p. *356.

237

†3 See 564 and 4.103f.

237

†P1 Another of De Morgan's examples [Formal Logic, p. 168] is this: "Suppose a person, on reviewing his purchases for the day, finds, by his counterchecks, that he has certainly drawn as many checks on his banker (and maybe more) as he has made purchases. But he knows that he paid some of his purchases in money, or otherwise than by checks. He infers then that he has drawn checks for something else except that day's purchases. He infers rightly enough." Suppose, however, that what happened was this: He bought something and drew a check for it; but instead of paying with the check, he paid cash. He then made another purchase for the same amount, and drew another check. Instead, however, of paying with that check, he paid with the one previously drawn. And thus he continued without cessation, or ad infinitum. Plainly the premisses remain true, yet the conclusion is false.

239

†1 This undated note seems to have been written for publication in some issue of The American Journal of Mathematics, shortly subsequent to that in which the previous article appeared. Why it was not published is unknown.

243

†1 The second formula is incorrect. The second half should read:

a + xb + a x; or a + xb + ax̄b.

243

†2 Cf. 2.517-27.

244

†1 399.

245

†1 I.e., every a is identical with every b.

247

†1 This should be 1̅eh.

248

†1 I.e., any two terms are identical and related or they are not identical and the said relation does not relate them.

248

†2 I.e., every two terms are related by some relation.

250

†1 The Open Court, vol. 6, pp. 3391-4 (1892).

251

†1 Or possibly in some other Renaissance writing. My memory may deceive me; and my library is precious small.

253

†1 Cf. 2.593ff.

254

†1 "On the Syllogism," II, Transactions, Cambridge Philosophical Society, vol. 9, p. 104, (1851).

255

†1 See 136a.

257

†1 The Open Court, vol. 6, pp. 3416-8, (1892).

258

†1 Nothing of the kind seems to have been published by Peirce, and there is no record of any relevant manuscript.

258

†2 Page 5057, ed. of 1889; see also 571f.

259

†P1 In this connection, see James's, Principles of Psychology, vol. 1, pp. 237-271; Briefer Course, pp. 160 et seq. James is no logician, but it is not difficult to trace a connection between the points he makes and the theory of inference.

260

†1 Paper no. XVI and see vol. 4, bk. II.

261

†1 Cf. discussion in vol. 2, bk. II, ch. 4, on the nature of propositions. The previous chapters of the same book should clarify what follows.

262

†1 Cf. 2.287n.

262

†P1 Nature, in connection with a picture, copy, or diagram, does not necessarily denote an object not fashioned by man, but merely the object represented, as something existing apart from the representation.

262

†2 Cf. 469f, 1.289f, 1.346.

262

†P2 Thus, CO, which appears as such a radicle in formic acid, makes of itself a saturated compound.

263

†1 Cf. 63 and 1.347.

263

†2 Cf. vol. 1, bk. III.

264

†1 This is the last paper on logic to be published in The Open Court.

264

†P1 Philosophical Transactions for 1886 [pp. 1-70]. No logician should fail to study this memoir.

264

†P2 I use this word in its proper sense, and not to mean unlike, as Mr. Kempe does.

266

†1 The Monist, vol. 7, pp. 19-40, (1896).

266

†P1 Vorlesungen über die Algebra der Logik, (Exakte Logik). Von Dr. Ernst Schröder, Ord. Professor der Mathematik an der technischen Hochschule zu Karlsruhe in Baden. Dritter Band. Algebra und Logik der Relative. Leipsic: B. G. Teubner. 1895. Price, 16M.

268

†1 See 2.30.

270

†1 La philosophie positive, deuxèiame leçon.

270

†2 Cf. the classification of sciences in vol. 1, Bk. II.

270

†3 By Peirce, p. 5397, ed. of 1889.

272

†1 The first and second parts are the topics of bks. II and III of vol. 2; the third is discussed in vol. 5 and 6.

272

†2 Opera Omnia Collecta, T. 1, pp. 45-76. L. Durand.

272

†3 1.559.

273

†1 Bd. 1, S. 118.

273

†2 Cf. 2.19.

279

†1 Acad. Quaest. II, 143.

279

†2 E.g. Sextus Empiricus, Adv. Math. VIII, 113-17.

279

†3 The Diodorians in opposition to the Philonians deny that material implication expresses what is usually meant by "if ... then ...." For contemporary discussion see G. E. Moore, Philosophical Studies, p. 276ff; C. I. Lewis, A Survey of Symbolic Logic, esp. p. 324ff, and Paul Weiss, "Relativity in Logic," The Monist, October, 1928; "The Nature of Systems," The Monist, April/July, 1929; "Entailment and the Future of Logic," Proceedings, Seventh International Congress of Philosophy; E. J. Nelson, "Intensional Relations," Mind, October, 1930.

280

†1 Quæstiones in Octo libror Physicorum Aristotelis, L. 1, qu. II.

282

†1 Cf. 2.352 and 2.618.

282

†2 Vol. 5, bk. II, ch. 3.

283

†1 182-197.

283

†2 The editors have not considered it worth publishing. But see 4.277ff. and vol. 4, Bk. II.

284

†1 See 472 and 2.532-5.

284

†2 Symbolic Logic, p. 39ff.

284

†3 The principle of duality is expressible in the formulæ: -(a+b) = (āb̄) and -(ab) = (ā+b̄).

286

†1 See vol. 2, bk. II, ch. 1, for a discussion of the "ethics of terminology."

286

†2 Peirce's first contribution to 'exact' logic is published in the Appendix to vol. 2.

288

†1 The Monist, vol. 7, pp. 161-217, (1897).

288

†2 Cf. vol. 5, bk. II, ch. 5.

288

†P1 Algebra und Logik der Relative. Leipsic: B. G. Teubner. 1895. Price, 16 M.

291

†1 See 42-44, 1.83, 2.227, 2.364 and 4.235.

294

†P1 The Pythagoreans, who seem first to have used these words, probably attached a patronymic signification to the termination. A triad was derivative of three, etc.

295

†1 In this section Peirce presents his "Entitative Graphs." The "Existential Graphs" are to be found in vol. 4, bk. II.

295

†2 Part 1, 1886, pp. 1-70.

295

†3 Cf. 1.289f., 1.346 and 421.

296

†1 See J. J. Sylvester, "Chemistry and Algebra," Mathematical Papers, vol. III, No. 14; W. K. Clifford; "Remarks on the Chemico-Algebraic Theory," Mathematical Papers, No. 28.

297

†1 Meyer used the volumes as abscissæ and the weights as ordinates. See Das Natürliche System der Chemischen Elemente, Meyer u. Mendejeff, Leipzig, 1895.

297

†2 See 475f.

299

†1 I.e., s = s.

300

†1 i.e., s†$ but not ~; or perhaps more clearly (s) ⤙ ( ⤙ $) but not s.

300

†P1 Professor Schröder proposes to substitute the word "symmetry" for convertibility, and to speak of simply convertible modes of junction as "symmetrical." Such an example of wanton disregard of the admirable traditional terminology of logic, were it widely followed, would result in utter uncertainty as to what any


Figure 6

writer on logic might mean to say, and would thus be utterly fatal to all our efforts to render logic exact. Professor Schröder denies that the mode of junction in "lover of a servant" is "symmetrical," which word in practice he makes synonymous with "commutative," applying it only to such junctions as that between "lover" and "servant" in "Adolphus is at once lover and servant of Eugenia." Commutativity depends on one or more polyadic relatives having two like blanks as shown in Fig. 6.

305

†P1 In my method of graphs, the spots represent the relatives, their bonds the hecceities; while in Mr. Kempe's method, the spots represent the objects, whether individuals or abstract ideas, while their bonds represent the relations. Hence, my own exclusive employment of bonds between pairs of spots does not, in the least, conflict with my argument that in Mr. Kempe's method such bonds are insufficient.

306

†1 I.e., evenly encircled bonds have π as their quantifier, while the others have Σ.

309

†P1 "On the Natural Classification of Arguments." Proceedings of the American Academy of Arts and Sciences. [2.477.]

312

†1 The use of one such logical constant is shown by Peirce to be sufficient for the development of Boolian Algebra. See e.g., 4.1ff.

314

†1 Cf. 332-4.

315

†1 Symbolic Logic, p. 39ff.

316

†1 See 47n.

317

†1 But see 396.

319

†1 This should be yj.

320

†1 1885, see No. XIII.

320

†2 This does not seem to have been done.

321

†1 See S. 296.

321

†2 Peirce wrote inline image= x u · a which conforms neither to Schröder nor to the illustration in the text.

322

†1 Cf. 2.316.

324

†1 Compare the physiological explanation of deductions, inductions and abductions in 2.643.

325

†1 See S.161ff.

325

†2 See S.163.

325

†3 See S.165.

328

†1 This notation is not exactly that of Schröder, who writes a = 1aβγ0, and ā = 0āβ̄γ̄1. See S. 205.

329

†1 See S.231.

330

†1 See S. 239.

332

†1 For l1 inline image $2, = $1 inline image $2, = 2, inline image 1

332

†2 See vol. 4, bk. I, ch. 4.

333

†1 See vol. 4, bk. I, ch. 7.

333

†2 "Kalkül der Abzählenden Geometrie," 1879.

333

†3 See 4.219ff.

333

†4 See 6.450.

333

†5 See e.g. 374 and 442.

334

†P1 For the simple reason that the real world is a part of the ideal world. namely, that part which sufficient experience would tend ultimately (and therefore definitively), to compel Reason to acknowledge as having a being independent of what he may arbitrarily, or willfully, create.—Marginal note, 1908.

334

†P2 That is to say each is vaguely, not distinctly, possible. — Marginal note, 1908.

335

†P1 Or linearly [by taking the diagonals] as follows. But there the small primes come earlier: 1; 2, 3; 5; 4, 6, 9; 10, 15; 7; 8, 12, 18, 27; 20, 30, 45; 14, 21; 35; 16, 24, 36, 54, 81; 40, 60, 90, 135; 28, 42, 63; 70, 105; 11, etc." — Marginal note, 1908.

336

†1 Cf. 1.61, 1.591ff, 2.156.

336

†2 Where?

336

†3 See 4.552n.

336

†4 Obviously a misprint for: ΣiΠjij

337

†1 (1) ΣiΠjij ⤙ ΠjΣilij (2); (3) ΣjΠilij ⤙ ΠiΣjlij (4). (2) and (3) are contradictories; (1) and (4) are contradictories. If (1) were true (2) would be true. If (3) were true (4) would be true. If (1) and (3) were true, (2) and (3), which are contradictories, would both be true. Though (1) and (3) thus cannot both be true, they may both be false. They are related to one another as an A and an E; their contradictories (4) and (2) must be related to one another as an I and an O — both can be and one at least must be true.

337

†2 See 2.517ff.

338

†P1 I prefer to speak of a member of a collection as a subject of it rather than as an object of it; for in this way I bring to mind the fact that the collection is virtually a quality or class-character. [A collection is a rhema or propositional function. Its members are those subjects which make it a true proposition. See 66.]

339

†1 I.e., all the i's which are members of P, are related to the j's which are members of Q, and there is no k distinct from an i which has the same relation to the j's which the i's have to the j's.

339

†P1 It must be remembered that to a person familiar with the algebra all such series of steps become evident at first glance.

339

†2 Cf. 532.

343

†1 See "Ueber eine elementare Frage der Mannigfaltigkeitslehre," (1890-1), Georg Cantor Gesammelte Abhandlung, herausg. E. Zermelo, S. 278-81, Berlin, (1932) where Cantor shows that 2n is always greater than n.

343

†P1 Inasmuch as the above theorem is, as I believe, quite opposed to the opinion prevalent among students of Cantor, and they may suspect that some fallacy lurks in the reasoning about wishes, I shall here give a second proof of a part of the theorem, namely that there is an endless succession of infinite multitudes related to one another as above stated, a relation entirely different, by the way, from those of the orders of infinity used in the calculus. I shall not be able to prove by this second method, as is proved in the text, that there are no higher multitudes, and in particular no maximum multitude. The ways of distributing a collection into two houses are equal to the possible combinations of members of that collection (including zero); for these combinations are simply the aggregates of individuals put into either one of the houses in the different modes of distribution. Hence, the proposition is that the combinations of whole numbers are more multitudinous than the whole numbers, that the combinations of combinations of whole numbers are still more multitudinous, the combinations of combinations of combinations again more multitudinous, and so on without end.

I assume the previously proved proposition that of any two collections there is one which can be placed in one-to-one correspondence with a part or the whole of the other. This obviously amounts to saying that the members of any collection can be arranged in a linear series such that of any two different members one comes later in the series than the other.

A part may be equal to the whole; as the even numbers are equal in multitude to all the numbers (since every number has a double distinct from the doubles of all other numbers, and that double is an even number). Hence, it does not follow that because one collection can be placed in one-to-one correspondence to a part of another, it is less than that other, that is, that it cannot also, by a rearrangement, be placed in one-to-one correspondence with the whole. This makes an inconvenience in reasoning which can be overcome in a manner I proceed to describe.

Let a collection be arranged in a linear series. Then, let us speak of a section of that series, meaning the aggregate of all the members which are later than (or as late as) one assignable member and at the same time earlier than (or as early as) a second assignable member. Let us call a series simple if it cannot be severed into sections each equal in multitude to the whole. A series not simple itself. may be conceivably severed into simple sections, or it may be so arranged that it cannot be so severed (for example the series of rational fractions arranged in the order of their magnitudes). But suppose two collections to be each ranged in a linear series, and suppose one of them, A, is in one-to-one correspondence with a part of the other B. If now the latter series, B, can be severed into simple sections, in each of which it is possible to find a member at least as early in the series as any member of that section that is in correspondence with a member of the other collection A, and also a member at least as late in the series as any member of that section that is in correspondence with any member of the other collection, and if it is also possible to find a section of the series, B, equal to the whole series, B, in which it is possible to find a member later than any member that is in correspondence with any member of the collection, A, then I say that the collection, B, is greater than the collection, A. This is so obvious that I think the demonstration may be omitted.

Now, imagine two infinite collections, the α's and the β's, of which the β's are the more multitudinous. I propose to prove that the possible combinations of β's are more multitudinous than the possible combinations of α's. For let the pairs of conjugate combinations (meaning by conjugate combinations a pair each of which includes every member of the whole collection which the other excludes) of the β's be arranged in a linear series; and those of the α's in another linear series. Let the order of the pairs in each of the two series be subject to the rule that if of two pairs one contains a combination composed of fewer members than either combination of the other pair, it shall precede the latter in the series. Let the order of the pairs in the series of pairs of combinations of β's be further determined by the rule that where the first rule does not decide, one of two pairs shall precede the other whose smaller combination (this rule not applying where one [?] combinations are equal) contains fewer β's which are in correspondence with α's in one fixed correspondence of all the α's with a part of the β's.

In this fixed correspondence each α has its β, while there is an infinitely greater multitude of β's without α's than with. Let the two series of pairs of combinations be so placed in correspondence that every pair of unequal combinations of α's is placed in correspondence with that pair of combinations of β's of which the smaller contains only the β's corresponding in the fixed correspondence to the smaller combination of α's; and let every pair of equal combinations of α's be put into correspondence with a pair of β's of which the smaller contains only the β's belonging in the fixed correspondence to one of the combinations of α's.

Then it is evident that each series will generally consist of an infinite multitude of simple sections. In none of these will the combinations be more multitudinous than those of the β's. In some, the combinations of α's will be equal to those of the β's; but in an infinitely greater multitude of such simple sections and each of these infinitely more multitudinous, the combinations of β's will be infinitely more multitudinous than those of the α's. Hence it is evident that the combinations of the β's will on the whole be infinitely more multitudinous than those of the α's.

That is if the multitude of finite numbers be α, and 2a = b, 2b = c, 2c = d, etc. a < b < c < d < etc. ad infinitum.

It may be remarked that the finite combinations of finite whole numbers form no larger a multitude than the finite whole numbers themselves; i.e. they are at least enumerable. But there are infinite collections of finite whole numbers; and it is these which are infinitely more numerous than those numbers themselves.

345

†1 Cf. 4113.

345

†2 This is the last of the published papers on Schröder.

346

†1 Educational Review, pp. 209-16, (1898).

346

†2 Metaphysica 1061a 28-1061b 3; 1061b 21-25.

346

†P1 Davidson, Aristotle and the ancient educational ideals. Appendix: The Seven Liberal Arts. (New York: Charles Scribner's Sons.)

347

†P1 Brouillon, Proiet d'une atteinte aux événemens des rencontres du cône avec son plan, 1639.

348

†P1 In his Linear associative algebras, [p. 97, published in the American Journal of Mathematics, vol. 4, (1881), pp. 97-229; see No. VIII.]

349

†P1 See this well put in Thomson and Tait's Natural Philosophy, §447.

352

†P1 The Mathematical Psychology of Boole and Gratry.

352

†1 No further articles were published in the Educational Review. The following, however, was part of the original article and seems to have been omitted only because of lack of space. What follows is taken from paginated page proofs.

353

†P1 I notice that writers of school arithmetics shrink from accepting the correct name of their art, Vulgar Arithmetic. It is a pity we have lost Chaucer's word "augrim," which, etymologically meaning "the art of the Chorasmian," is free from all objection. Still, we cannot find fault with these writers who adopt no more high sounding title for their subject than Practical Arithmetic.

354

†1 This should be: C is not R'd by B.

354

†2 Given the precepts

I. (ΠaΠb) - (a R b) inline image - (b R a)

II. (ΠaΠbΠc) a R b inline image b R c inline image c R a

through the use of the propositions of logic and the principles of substitution and inference, the following are some of the theorems that can be derived.

A. b) - (b R b)
[I:(b/a)] - (b R b) inline image - (b R b) (1)
(1) ⤙ - (b R b)
B. aΠb) a R b inline image b R a
[II:(b/c)] a R b inline image b R b inline image b R a (1)
(A) · (1) ⤙ a R b inline image b R a
C. aΠbΠc) b R a ⤙ (b R c inline image c R a)
I = b R a ⤙ - (a R b) (1)
II = -(a R b) ⤙ (b R c inline image c R a) (2)
(1)·(2)⤙ [b R a ⤙ (b R c inline image c R a)]
D. aΠbΠc)(b R a · c R b) ⤙ c R a
[I:(c/a)] - (c R b) inline image - (b R c) (1)
= c R b ⤙ - (b R c) (2)
C = b R a ⤙ [-(b R c) ⤙ c R a] (3)
= [b R a . -(b R c)] ⤙ c R a (4)
(2) · (4) ⤙ (b R a · c R b) ⤙ c R a
E. aΠbΠc) - (b R a) inline image - (c R b) inline image -(a R c)
D = -(b R a) inline image - (c R b) inline image c R a (1)
[I:(c/b)] -(a R c) inline image -(c R a) (2)
(1) · (2) ⤙ -(b R a) inline image -(c R b) inline image -(a R c)
355

†1 4 = -(c R a) ⤙ -(b R a) inline image -(c R b), which by [(n/c),(m/a),(x/b)] is -(nRm)⤙ -(xRm) inline image -(nRx).

355

†P1 The word hardly is in older English "hard." N follows hard upon M, that is, solidly up against M, with nothing between them.

359

†1 The proof sheet ends here with the note "to be continued." The ms. of this paper has not been uncovered but similar papers are published in vol. 4, bk. I.

360

†1 A letter to the Editor of Science, vol. 2, pp. 430-33, March 16, 1900.

360

†2 Stetigkeiten u. Irrationalen Zahlen, 2te Auf., 1892; Was sind u. was sollen die Zahlen, 1888.

360

†3 "Beiträge zur Begründung der transfiniten Mengenlehre," Georg Cantor Gesammelte Abhandlung, herausg. E. Zermelo, S. 282-351, Berlin, (1932).

360

†4 258, 281ff., 402.

360

†5 But see Was sind u. was sollen die Zahlen, §64.

361

†1 "Mr. Charles Peirce, as I understand his statements in the Monist, appears to stand almost alone amongst recent mathematical logicians outside of Italy, in still regarding the Calculus as properly to be founded upon the conception of the actually infinite and infinitesimal." The World and the Individual, 1st Series, p. 562n.

361

†2 548.

364

†1 Op. cit., p. 562n.

365

†1 Georg Cantor Gesammelte Abhandlung, S. 275-6.

366

†1 Printed separately in eight pages, circa 1903, apparently intended as the second part of A Syllabus of Certain Topics of Logic, published as a supplement to the Lowell Lectures of 1903. See vol. 1, bk. II, ch. 1, note.

367

†P1 It is far better to invent a word for a purely technical conception than to use an expression liable to be corrupted by being employed by loose writers. I reduplicate the first syllable of relation to form this word, with little reference to the meaning of the syllable as a preposition. Still, relations of this kind are the only ones that might be asserted of the same relates transposed; and the reduplication of the preposition re connotes such transposition.

367

†P2 I must, with pain and shame, confess that in my early days I showed myself so little alive to the decencies of science that I presumed to change the name of this branch of logic, a name established by its author and my master, Augustus De Morgan, to "the logic of relatives." I consider it my duty to say that this thoughtless act is a bitter reflection to me now, so that young writers may be warned not to prepare for themselves similar sources of unhappiness. I am the more sorry, because my designation has come into general use.

369

†1 I.e.,

1. extraloves al̄b

2. contraloves -(alb)

3. juxtaloves -(al̄b)

4. reloves al̆b

5. coloves (alc)(cl̆b)

6. ultraloves (alc)(c$l̄b)

7. transloves -[(al̄c)(cl̆b)]

8. superloves {-[(al̄c)(cl̆b)][(ald)(d$l̄b)]}

369

†2 See note to 4.

369

†3 The following schedule may be of aid in this section:

1. ΣiΣjrij — lation

2. ΣiΣjrij — contralation;

3. ΠiΠj~rij — extralation; r = 0

4. ΠiΠjrij — juxtalation; r = ∞

5. ΣiΠjrij — perlation

6. ΣiΠjr̄ij — contraperlation

7. ΠiΣjr̄ij — extraperlation

8. ΠiΣjrij — juxtaperlation

9. ΣjΠirij — reperlation

10. ΣjΠir̄ij — contrareperlation

11. ΠjΣiij — extrareperlation

12. ΠjΣirij — juxtareperlation

13. ΣiΠjΣkrik · rjk — conlation

14. ΣiΠjΣkrik · rjk — ultralation

15. ΠiΣjΠk - (ik . rjk) — translation

16. ΠiΠjΠkΣl - (ik · rjk)(ril · j l) — superlation

17. ΣiΣjrij∞ — essential perlation

18. ΣiΣjrij — essential reperlation

19. ΣiΣjij∞ — contressentiperlation

1 = -3, 5 = -7, 9 = -11; 2 = -4; 6 = -8; 10 = -12, 5 ⤙ 12; 9⤙8; 6⤙11; 10⤙7; 17.18⤙3 inline image 4.

†P1 Since this shows he felt the obligation, it is the more lamentable that his treatment of my notation showed no such scrupulosity. That notation had been most maturely considered and thoroughly put to the test; and his changes were, without exception, for the worse. I did not say this while he lived, out of regard for his feelings. His additions to the notation, to express what I had afforded no means of expressing, stand, of course, on a different footing; and I should be bound to follow him, here.

371

†1 Juxtaperlation.

371

†2 Better: ". . . there is nothing to which everything stands."

372

†1 I.e., ΣiΠjrij ⤙ ΠjΣirij; ΣjΠirij ⤙ ΠiΣjrij.

372

†2 ΣiΠjr̄ i j ⤙ ΠjΣir̄ i j; ΣjΠir̄ i j ⤙ ΠiΣjr̄ i j.

373

†1 [[Image will not appear in pop-up, please follow link. - NLX Ed.]]

374

†1 The following schedule may be of aid in this section:

1. Πirii —suilation

2. ΠiΠjiijjrij —contrasuilation—alio-relative

3. ΣiΣjiijjrij —extrasuilation

4. Σirii —juxtasuilation—self-relative

 

5. ΠiΠjrij·ij —ambilation

6. ΠiΠjrii rjjij —contrambilation—concurrency

7. ΣiΣjij·ij —extrambilation

8. ΣiΣjrij·ij —juxtambilation—opponency

 

9. ΣiΠjrij·ij —peneperlation

10. ΣjΠirij·ij —penereperlation

11. ΣiΠjij·ij —penecontraperlation

12. ΣjΠirij·ij —penecontrareperlation

1 ⤙ 4; 5 ⤙ 8; 9 ⤙ 8; 10 ⤙ 8; 11 ⤙ 7; 12 ⤙ 7.

1.5 = ∞. 2.5 = N.

1.6 = 1. 2.6 = O.

375

†1 [[Image will not appear in pop-up, please follow link. - NLX Ed.]]

376

†1 See 136.

376

†2 This should be: peneperlation.

376

†3 On one of his copies of this syllabus C. S. P. writes: "Here the fund for the printing gave out." The rest of this paper is from manuscript.

376

†4 The following schedule may be of use in this section:

1. ΠiΠkΣj iRj · jRk granilation

2. ΣiΣkΠj - (iRj · jRk) extragranilation

3. ΠiΠkΣj - (iRj · jRk) contragranilation

4. ΣiΣkΠj iRj · jRk juxtagranilation

5. ΠiΠkΠlΠnΣjΣm iRj · jRk · lRm · mRn · ΠjΣm jRm · ΠmΣj jRm spicalation

6. ΠiΣjΠk iRj · jRkiRk transitive extraspicalation

7. ΠiΠjΣk iRj · k R jiRk idempotent contraredultralation

377

†1 See 576.

377

†2 "On the Syllogism IV," Transactions Cambridge Philosophical Society, vol. 10, p. 346 (1859).

378

†1 I.e., if (B:C) and (C:F) imply (B:F), then if (C:X) holds, (B:X) holds, X standing for any correlate; and if (Y:C) holds then (Y:F) holds, Y standing for any relate.

379

†1 I.e., BC, CA, CE, CG, FC, GD ⤙ BA, BD, BE, BG, CD, FA, FD, FE, FG.

380

†1 See 136b.

380

†2 See 136b.

380

†3 See his Linear Associative Algebras, §25.

380

†4 See editors' note to §2.

381

†1 See editors' note to 2.

381

†2 Novum Organon Renovatum, bk. II, ch. IV.

381

†3 Syllabus of Logic, §§124, 144.

381

†4 See Prantl, Geschichte der Logik, Bd. I, S. 521.

382

†1 Cf. 136e.

382

†P1 I am not here making infinity a limit but in making the state of exceeding the finite the limit of successive increments by unity.

383

†1 "The Lowell Lectures," 1903, for which this syllabus was intended as a supplement.

383

†2 See 136c.

383

†3 Cf. Prantl, op. cit., Bd. III, S. 153, 282.

383

†4 Cf. Prantl, op. cit., Bd. III, S. 290.

383

†5 Formal Logic, p. 345.

384

†1 Cf. 121ff.

384

†2 Algebra der Logik, 3.1, S. 11-12 and passim.

384

†P1 But in the mathematical books, that which fills it is called an Element.

387

†1 See 2.406ff.

387

†2 I.e., Speculative Grammar, for which see vol. 2, bk. II.

387

†3 The remainder of this syllabus is to be found in 2.233 seq., where triadic relations, and, in particular, the divisions of signs are treated at length.

388

†1 Dictionary of Philosophy and Psychology, ed. by J. M. Baldwin, Macmillan & Co., N.Y., vol. 1, p. 518, by Peirce and H. B. Fine. 2d ed. 1911.

388

†2 "Untersuchung über Gegenstände der höhere Geodäsie; 2 Abhandlung;" Abh. d. Königl. G. d. W. zu Göttingen, 2ter u. 3ter Bd.

388

†3 See vol. 7.

390

†1 Dictionary of Philosophy and Psychology, vol. 1, p. 537-38.

390

†2 See Prantl, op. cit., Bd. I, S. 661, 684.

390

†3 Vol. 2, bk. II, ch. 4, §12.

390

†4 Cf. 1.435.

392

†1 Vol. 2, bk. II, ch. 4, 12.

392

†2 Dictionary of Philosophy and Psychology, vol. 1, p. 574.

392

†3 See 77-80.

392

†4 See 113-118.

393

†1 Dictionary of Philosophy and Psychology, vol. 2, pp. 24-27.

394

†1 Vol. 4, bk. II, ch. 1., §1.

395

†1 §8.

395

†2 A number of elementary definitions of such familiar terms as "aggregation," "absorption," "associative," "commutative," etc. have been omitted.

398

†1 See 47n.

399

†1 Johns Hopkins Studies in Logic, p. 87ff.

399

†2 Dictionary of Philosophy and Psychology, vol. 2, p. 117-18, by Peirce and H. B. Fine.

400

†1 See Georg Cantor Gesammelte Abhandlung S. 282, Berlin, (1932).

401

†1 Dictionary of Philosophy and Psychology, vol. 2, pp. 315-16.

401

†2 See Analytica Posteriora, 76b, 26-31.

403

†1 Dictionary of Philosophy and Psychology, vol. 2, p. 338.

404

†1 Ibid., vol. 2, pp. 447-50.

408

†1 Principles of Psychology, I, 243.

409

†1 See Paper No. III.

409

†2 See Paper No. XII.

409

†3 Dictionary of Philosophy and Psychology, vol. 2, p. 713, by Christine Ladd-Franklin and Peirce.

411

†1 Johns Hopkins University Circulars, No. 22, pp. 86-88, April, 1883, entitled "A Communication from Mr. Peirce."

413

†1 In his Linear Associative Algebras, 1870.

413

†2 Paper No. III.

414

†1 See 130.

414

†2 See 150-51.

414

†3 See 130.

†P1 It would have been more accurately analogical, perhaps, to call it novenions.

414

†4 §127, beginning with "for example," was here reproduced.