1911 Encyclopædia Britannica/Psychology

From Wikisource
Jump to navigation Jump to search
5206301911 Encyclopædia Britannica, Volume 22 — PsychologyJames Ward

PSYCHOLOGY (ψυχή, the mind or soul, and λόγος, theory), which can only be more strictly defined by an analysis of what “mind” means.

1. In the several natural sciences the scope and subject-matter of each are so evident that little preliminary discussion is called for. But with psychology, however much it is freed from metaphysics, this is different. It is indeed ordinarily assumed that its subject-matter can be at once defined. “It is what you can perceive by consciousness or The Science of “Mind.” reflection or the internal sense,” says one, “just as the subject-matter of optics is what you can perceive by sight.” Or, “psychology is the science of the phenomena of mind,” we are told again, “and is thus marked off from the physical sciences, which treat only of the phenomena of matter.” But, whereas nothing is simpler than to distinguish between seeing and hearing, or between the phenomena of heat and the phenomena of gravitation, a very little reflection may convince us that we cannot in the same fashion distinguish internal from external sense, or make clear to ourselves what we mean by phenomena of mind as distinct from phenomena of matter.

To every sense there corresponds a sense-organ; the several senses are distinct and independent, so that no one sense can add to or alter the materials of another: the possession of five senses, e.g. furnishing no data as to the character of a possible sixth. Moreover, sense-impressions are passively received and occur in the first instance without regard Internal and External. to the feeling or volition of the recipient and without any manner of relation to the “contents of consciousness” at the moment. Now such a description will apply but very partially to the so-called internal sense. For we do not by means of it passively receive impressions differing from all previous presentations, as the sensations of color for one “couched” differ from all he has experienced before: the new facts consist rather in the recognition of certain relations among pre-existing presentations, i.e. are due to our mental activity and not to a special mode of what has been called our sensitivity. For when we taste we cannot hear that we taste, when we see we cannot smell that we see; but when we taste we may be conscious that we taste, when we hear we may be conscious that we hear. Moreover, the facts so ascertained are never independent of feeling and volition and of the contents of consciousness at the time, as true sensations are. Also if we consult the physiologist we learn that there is no evidence of any organ or centre that could be regarded as the “physical basis” of this inner sense; and, if self-consciousness alone is temporarily in abeyance and a man merely “beside himself,” such state of delirium has little analogy to the functional blindness or deafness that constitutes the temporary suspension of sight or hearing.

To the concept of an internal perception or observation the preceding objections do not necessarily apply—that is to say, this concept may be so defined that they need not. But then in proportion as we escape the change of assuming a special sense which furnishes the material for such perception or observation, in that same proportion are we compelled to seek for some other mode of distinguishing its subject-matter. For, so far as the mere mental activity of perceiving or observing is concerned, it is not easy to see any essential difference in the process whether what is observed be psychical or physical. It is quite true that the so-called psychological observation is more difficult, because the facts observed are often less definite and less persistent, and admit less of actual isolation than physical facts do; but the process of recognizing similarities or differences, the dangers of mal-observation or non-observation, are not materially altered on that account. It may be further allowed that there is one difficulty peculiarly felt in psychological observation, the one most inaccurately expressed by saying that here the observer and the observed are one. But this difficulty is surely in the first instance due to the very obvious fact that our powers of attention are limited, so that we cannot alter the distribution of attention at any moment without altering the contents of consciousness at that moment. Accordingly, where there are no other ways of surmounting this difficulty, the psychological observer must either trust to representations at a later time, or he must acquire the power of taking momentary glances at the psychological aspects of the phase of consciousness in question. And this one with any aptitude for such studies can do with so slight a diversion of attention as not to disturb very seriously either the given state or that which immediately succeeds it. But very similar difficulties have to be similarly met by physical observers in certain special cases, as, e.g. in observing and registering the phenomena of solar eclipse; and similar aptitudes in the distribution of attention have to be acquired, say, by extempore orators or skilful surgeons. Just as little, then, as there is anything that we can with propriety call an inner sense, just so little can we find in the process of inner perception any satisfactory characteristic of the subject-matter of psychology. The question still is: What is it that is perceived or observed? and the readiest answer of course is: Internal experience as distinguished from external, what takes place in the mind as distinct from what takes place without.

This answer, it must be at once allowed, is adequate for most purposes, and a great deal of excellent psychological work has been done without ever calling it in question. But the distinction between internal and external experience is not one that can be drawn from the standpoint of psychology, at least not at the outset. From this standpoint it appears to be either (1) inaccurate or (2) not extra-psychological. As to (1), the boundary between the internal and the external was, no doubt, originally the surface of the body, with which the subject or self was identified; and in this sense the terms are of course correctly used. For a thing may, in the same sense of the word, be in one space and therefore not in—i.e. out of—another; but we express no intelligible relation if we speak of two things as being one in a given room and the other in last week. Any one is at liberty to say if he choose that a certain thing is “in his mind”; but if in this way he distinguishes it from something else not in his mind, then to be intelligible this must imply one of two statements—either that the something else is actually or possibly in some other mind, or, his own mind being alone considered, that at the time the something else does not exist at all. Yet, evident as it seems that the correlatives in and not-in must apply to the same category, whether space, time, presentation (or non-presentation) to a given subject, and so forth, we still find psychologists more or less consciously confused between “internal,” meaning presented in the psychological sense, and “external,” meaning not “not-presented” but corporeal or oftener extra-corporeal. But (2), when used to distinguish between presentations (some of which, or some relations of which with respect to others, are called “internal,” and others or other relations, “external”), these terms are at all events accurate; only then they cease to mark off the psychological from the extra-psychological, inasmuch as psychology has to analyse this distinction and to exhibit the steps by which it has come about. But we have still to examine whether the distinction of phenomena of Matter and phenomena of Mind furnishes a better dividing line than the distinction of internal and external.

A phenomenon, as commonly understood, is what is manifest, sensible, evident, the implication being that there are eyes to see, ears to hear, and so forth—in other words, that there is presentation to a subject; and wherever there is presentation to a subject it will be allowed that we are in the domain of psychology. But in talking of physical phenomena Mental and Material. we, in a way, abstract from this fact of presentation. Though consciousness should cease, the physicist would consider the sum total of objects to remain the same: the orange would still be round, yellow and fragrant as before. For the physicist—whether aware of it or not—has taken up a position which for the present may be described by saying that phenomenon with him means appearance or manifestation, or—as we had better say—object, not for a concrete individual, but rather for what Kant called Bewusstsein überhaupt, or, as some render it, the objective consciousness, i.e. for an imaginary subject freed from all the limitations of actual subjects save that of depending on “sensibility” for the material of experience. However, this is not all, for, as we shall see presently, the psychologist also occupies this position; at least if he does not his is not a true science. But, further, the physicist leaves out of sight altogether the facts of attention, feeling, and so forth, all of which actual presentation entails. From the psychological point of view, on the other hand, the removal of the subject removes not only all such facts as attention and feeling, but all presentation or possibility of presentation whatever. Surely, then, to call a certain object, when we abstract from its presentation, a material phenomenon, and to call the actual presentation of this object a mental phenomenon, is a clumsy and confusing way of representing the difference between the two points of view. For the terms “material” and “mental” seem to imply that the two so-called phenomena have nothing in common, whereas the same object is involved in both, while the term “phenomenon” implies that the point of view is in each case the same, when in truth what is emphasized by the one the other ignores.

2. Paradoxical though it may be, we must then conclude that psychology cannot be defined by reference to a special subject-matter Standpoint of Psychology. as such concrete sciences, for example, as mineralogy and botany can be; and, since it deals in some sort with the whole of experience, it is obviously not an abstract science in any ordinary sense of that term. To be characterized at all, therefore, apart from metaphysical assumptions, it must be characterized by the standpoint from which this experience is viewed. It is by way of expressing this that widely different schools of psychology define it as subjective, all other positive sciences being distinguished as objective. But this seems scarcely more than a first approximation to the truth, and, as we have seen incidentally, is apt to be misleading. The distinction rather is that the standpoint of psychology is what is sometimes termed “individualistic,” that of the so-called object-sciences being “universalistic,” both alike being objective in the sense of being true for all, consisting of what Kant would call judgments of experience. For psychology is not a biography in any sense, still less a biography dealing with idiosyncrasies, and in an idiom having an interest and a meaning for one subject only, and incommunicable to any other. Locke, Berkeley and Hume have been severely handled because they regarded the critical investigation of knowledge as a psychological problem, and set to work to study the individual mind simply for the sake of this problem. But none the less their standpoint was the proper one for the science of psychology itself; and, however surely their philosophy was foredoomed to a collapse, there is no denying a steady psychological advance as we pass from Locke to Hume and his modern representatives. By “idea” Locke tells us he means “Whatsoever is the object of the understanding when a man thinks” (i.e. is conscious), and having, as it were, shut himself within such a circle of ideas he finds himself powerless to explain his knowledge of a world that is assumed to be independent of it; but he is able to give a very good account of some of these ideas themselves. He cannot justify his belief in the world of things whence certain of his simple ideas “were conveyed” any more than Robinson Crusoe could have explored the continents whose products were drifted to his desert island, though he might perhaps survey the island itself well enough. Berkeley accordingly, as Professor Fraser happily puts it, abolished Locke's hypothetical outer circle. Thereby he made the psychological standpoint clearer than ever—hence the truth of Hume's remark, that Berkeley's arguments “admit of no answer”; at the same time the epistemological problem was as hopeless as before—hence again the truth of Hume's remark that those arguments “produced no conviction.” Of all the facts with which he deals, the psychologist may truly say that their esse is percipi, inasmuch as all his facts are facts of presentation, are ideas in Locke's sense, or objects which imply a subject. Before we became conscious there was no world for us; should our consciousness cease, the world for us ceases too; had we been born blind, the world would for us have had no colour; if deaf, it would have had no sounds; if idiotic, it would have had no meaning. Psychology, then, never transcends the limits of the individual. But now, though this Berkeleyan standpoint is the standpoint of psychology, psychology is not pledged to the method employed by Berkeley and by Locke. Psychology may be individualistic without being confined exclusively to the introspective method. There is nothing to hinder the psychologist from employing materials furnished by his observations of other men, of infants, of the lower animals, or of the insane; nothing to hinder him taking counsel with the philologist or even the, physiologist, provided always he can show the psychological bearings of those facts which are not directly psychological. The standpoint of psychology is individualistic; by whatever methods, from whatever sources its facts are ascertained, they must—to have a psychological import—be regarded as having place in, or as being part of, some one's consciousness or experience. In this sense, i.e. as presented to an individual, “the whole choir of heaven and furniture of earth” may belong to psychology, but otherwise they are psychological nonentities. In defining psychology, however, the propriety of avoiding the terms mind or soul, which it implies, is widely acknowledged; mind because of the disastrous dualism of mind and matter, soul because of its metaphysical associations. Hence F. A. Lange's famous mot: modern psychology is Psychologie ohne Seele. But consciousness, which is the most frequent substitute, is continually confused with self-consciousness, and so is apt to involve undue stress on the subjective as opposed to the objective, as well as to emphasize the cognitive as against the conative factors. Experience, it is maintained, is a more fundamental and less ambiguous term. Psychology then is the science of individual experience. The problem of psychology, in dealing with this complex subject-matter, is in general—first, to ascertain its ultimate constituents, and, secondly, to determine and explain the laws of their interaction.

General Analysis.

3. In seeking to make a first general analysis of experience, we must start from individual human experience, for this alone is what we immediately know. From this standpoint we must endeavour to determine the “irreducible minimum” involved, so that our concept may apply to all lower forms of experience as well. Etymologically experience connotes practical acquaintance, efficiency and skill as the result of trial—usually repeated trial—and effort. Many recent writers on comparative psychology propose to make evidence of experience in this sense the criterion of psychical life. The ox knoweth his owner and the ass his master's crib, and so would pass muster; but the ant and the bee, who are said to learn nothing, would, in spite of their marvellous instinctive skill, be regarded as mere automata in Descartes's sense. That this criterion is decisive on the positive side will hardly be denied; the question how far it is available negatively we must examine later on. But it will be well first briefly to note some of the implications of this positive criterion: Experience is the process of becoming expert by experiment. The chief implication, no doubt, is that which in psychological language we express as the duality of subject and object. Looking at this relation as the comparative psychologist has to do, we find that it tallies in the main with the biological relation of organism and environment. The individuality of the organism corresponds to, though it is not necessarily identical with, the psychological subject, while to the environment and its changes corresponds the objective continuum or totum objectivum as we shall call it. This correspondence further helps us to see still more clearly the error of regarding individual experience as wholly subjective, and at the same time helps us to find some measure of truth in the naïve realism of Common Sense. As these points have an important bearing on the connexion of psychology and epistemology, we may attempt to elucidate them more fully.

Though it would be unwarrantable to resolve a thing, as some have done, into a mere meeting-point of relations, yet it is perhaps as great a mistake to assume that it can be anything determinate in itself apart from all relations to other things. By the physicist this mistake can hardly be made: for him action and reaction are strictly correlative: a material system can do no work on itself. For the biologist, again, organism and environment are invariably complementary. But in psychology, when presentations are regarded as subjective modifications, we have this mistaken isolation in a glaring form, and all the hopeless difficulties of what is called “subjective idealism” are the result. Subjective modifications no doubt are always one constituent of individual experience, but always as correlative to objective modifications or change in the objective continuum. If experience were throughout subjective, not merely would the term subjective itself be meaningless, not merely would the conception of the objective never arise, but the entirely impersonal and intransitive process that remained, though it might be described as absolute becoming, could not be called even solipsism, least of all real experience. Common Sense, then, is right in positing, wherever experience is inferred, (1) a factor answering to what we know as self, and (2) another factor answering to what each of us knows as the world. It is further right in regarding the world which each one immediately knows as a coloured, sounding, tangible world, more exactly as a world of sensible qualities. The assumption of naive realism, that the world as each one knows it exists as such independently of him, is questionable. But this assumption goes beyond individual experience, and does not, indeed could not, arise at this standpoint.

Answering to the individuality and unity of the subjective factor, there is a corresponding unity and individuality of the objective. Every Ego has its correlative Non-Ego, whence in the end such familiar saying as quot homines tot sententiae and the like. The doctrine of Leibnitz, that “each monad is a living mirror. . . representative of the universe according to its point of view,” will, with obvious reservations, occur to many as illustrative here. In particular, Leibnitz emphasized one point on which psychology will do well to insist. “Since the world is a plenum,” he begins, “all things are connected together and everybody acts upon every other, more or less, according to their distance, and is affected by their reaction; hence each monad is a living mirror,”[1] &c. Subject and Object, or (as it will be clearer in this connexion to say) Ego and Non-Ego, are then not merely logically a universe, but actually the universe, so that, as Leibnitz put it, “He who sees all could read in each what is happening everywhere” (Monadology, § 61). Though every individual experience is unique, yet the more Ego₁ is similar to Ego₂ the more their complementaries Non-Ego₁, Non-Ego₂ are likewise similar; much as two perspective projections are more similar the more adjacent their points of sight, and more similar as regards a given position the greater its distance from both points. No doubt we must also make a very extensive use of the hypothesis of subconsciousness, just as Leibnitz did, before we can say that the universe is the objective factor in each and every individual's experience. But we shall have in any case to allow that, besides the strictly limited “content” rising above the threshold of consciousness, there is an indefinite extension of the presentational continuum beyond it. And the Leibnitzian Monadology helps us also to clear up a certain confusion that besets terms such as “content of consciousness,” or “finite centre of experience”—a barbarous but intelligible phrase that has recently appeared—the confusion, that is, with a mosaic of mutually exclusive areas, or with a scheme of mutually exclusive logical compartments. Consciousnesses, though in one respect mutually exclusive, do not limit each other in this fashion. For there is a sense in which all individual experiences are absolutely the same, though relatively different as to their point of view, i.e. as to the manner in which for each the same absolute whole is sundered into subjective and objective factors.

This way of looking at the facts of mind helps, again, to dispel the obscurity investing such terms as subjective, intersubjective, transsubjective and objective, as these occur in psychological or epistemological discussions. For the psychologist must maintain that no experience is merely subjective: it is only epistemologists (notably Kant) who so describe individual experience, because objects experienced in their concrete particularity pertain, like so many idiosyncrasies, to the individual alone. In contrast with this, epistemologists then describe universal experience—the objects in which are the same for every experiment—as objective experience par excellence. And so has arisen the time-honoured opposition of Sense-knowledge and Thought-knowledge: so too has arisen the dualism of Empiricism and Rationalism, which Kant sought to surmount by logical analysis. It is in the endeavour to supplement this analysis by a psychological genesis that the terms intersubjective and transsubjective prove useful. The problem for psychology is to ascertain the successive stages in the advance from the one form of experience or knowledge to the other. “When ten men look at the sun or the moon,” said Reid, “they all see the same individual object.” But according to Hamilton this statement is not “philosophically correct . . . the truth is that each of these persons sees a different object. . . . It is not by perception but by a process of reasoning that we connect the objects of sense with existences beyond the sphere of immediate knowledge.”[2] Now it is to this “beyond” that the term transsubjective is applied, and the question before us is: How do individual subjects thus get beyond the immanence or immediacy with which all experience begins? By a “process of reasoning,” it is said. But it is at least true in fact, whether necessarily true or not, that such reasoning is the result of social intercourse. Further, it will be generally allowed that Kant's Analytik, before referred to, has made plain the insufficiency of merely formal reasoning to yield the categories of Substance, Cause and End, by which we pass from mere perceptual experience to that wider experience which transcends it. And psychology, again, may claim to have shown that in fact these categories are the result of that reflective self-consciousness to which social intercourse first gives rise.

But such intercourse, it has been urged, presupposes the common ground between subject and subject which it is meant to explain. How, it is asked, if every subject is confined to his own unique experience, does this intersubjective intercourse ever arise? If no progress towards intellective synthesis were possible before intersubjective intercourse began, such intercourse, as presupposing something more than immediate sense-knowledge, obviously never could begin.[3] Let us illustrate by an analogy which Leibnitz's association of experience with a “point of view” at once suggests. If it were possible for the terrestrial astronomer to obtain observations of the heavens from astronomers in the neighbouring stars, he would be able to map in three dimensions constellations which now he can only represent in two. But unless he had ascertained unaided the heliocentric parallax of these neighbouring stars, he would have no means of distinguishing them as near from the distant myriads besides, or of understanding the data he might receive; and unless he had first of all determined the still humbler geocentric parallax of our sun, those heliocentric parallaxes would have been unattainable. So in like manner we may say “intersubjective parallax” presupposes what we may call “subjective parallax,” and even this the psychological duality of object and subject. But such subjective parallax or acquaintance with other like selves is the direct outcome of the extended range in time which memory proper secures; and when in this way self has become an object, resembling objects become other selves or “ejects,” to adopt with slight modification a term originated by the late W. K. Clifford. We may be quite sure that his faithful dog is as little of a solipsist as the noble savage whom he accompanies. Indeed, the rudiments of the social factor are, if we may judge by biological evidence, to be found very early. Sexual union in the physiological sense occurs in all but the lowest Metazoa, pairing and courtship are frequent among insects, while “among the cold-blooded fishes the battle of the stickleback with his rivals, his captivating manœuvres to lead the female to the nest which he has built, his mad dance of passion around her, and his subsequent jealous guarding of the nest, have often been observed and admired.”[4] Among birds and mammals we find not merely that these psychological aspects of sexual life are greatly extended, but we find also prolonged education of offspring by parents and imitation of the parents by offspring. Even language, or, at any rate “the linguistic impulse,” is not wholly absent among brutes.[5] Thus as the sensori-motor adjustments of the organism to its environment generally advance in complexity and range, there is a concomitant advance in the variety and intimacy of its relations specially with individuals of its kind. It is therefore reasonable to assume no discontinuity between phases of experience that for the individual are merely objective and phases that are also effective as well; and once the effective level is attained, some interchange of experience is possible. So disappears the great gulf fixed betwixt subjective or individual and inter subjective or universal experience by rival systems in philosophy.

4. From this preliminary epistemological discussion we may pass on to the psychological analysis of experience itself. As to this, there is in the main substantial agreement; the elementary facts of mind cannot be expressed in less than three propositions—“I feel somehow,” “I know something,” “I do something.” But here at once there arises an important question, viz. What after all are we to understand by the subject of these propositions? The proposition “I feel somehow” is not equivalent to “I know that I feel somehow.” To identify the two would be to confound consciousness with self-consciousness. We are no more confined to our own immediate observations here than elsewhere; but the point is that, whether seeking to analyse one's own consciousness or to infer that of a lobster, whether discussing the association of ideas or the expression of emotions, there is always an individual self or “subject” in question. It is not enough to talk of feelings or volitions: what we mean is that some individual—man or worm—feels, strives, acts, thus or thus. Obvious as this may seem, it has been frequently either forgotten or gainsaid. It has been forgotten among details or through the assumption of a medley of faculties, each treated as an individual in turn, and among which the real individual was lost. Or it has been gainsaid, because to admit that all psychological facts pertain to an experiencing subject or experiment seemed to imply that they pertained to a particular spiritual substance, which was simple, indestructible, and so forth; and it was manifestly desirable to exclude such assumptions from psychology as a science aiming only at a systematic exposition of what can be known and verified by observation. But, however, much assailed or Subject or Ego. disowned, the concept of a “mind” or conscious subject is to be found implicitly or explicitly in all psychological writers whatever—not more in Berkeley, who accepts it as a fact, than in Hume, who treats it as a fiction. This being so, we are far more likely to reach the truth eventually if we openly acknowledge this inexpugnable assumption, if such it prove, instead of resorting to all sorts of devious periphrases to hide it. Now wherever the word Subject, or its derivatives, occurs in psychology we might substitute the word Ego and analogous derivatives, did such exist. But Subject is almost always the preferable term; its impersonal form is an advantage, and it readily recalls its modern correlative Object. Moreover, Ego has two senses, distinguished by Kant as pure and empirical, the latter of which was, of course, an object, the Me known, while the former was subject always, the I knowing. By pure Ego or Subject it is proposed to denote here the simple fact that everything experienced is referred to a Self experiencing. This psychological concept of a self or subject, then, is after all by no means identical with the metaphysical concepts of a soul or mind-atom, or of mind-stuff not atomic; it may be kept as free from metaphysical implications as the concept of the biological individual or organism with which it is so intimately connected.

The attempt, indeed, has frequently been made to resolve the former into the latter, and so to find in mind only such an Attempts to extrude the Ego. individuality as has an obvious counterpart in this individuality of the organism, i.e. what we may call an objective individuality. But such procedure owes all its plausibility to the fact that it leaves out of sight the difference between the biological and the psychological standpoints. All that the biologist means by a dog is “the sum of the phenomena which make up its corporeal existence.”[6] And, inasmuch as its presentation to any one in particular is a point of no importance, the fact of presentation at all may be very well dropped out of account. Let us now turn to psychology: Why should we not here follow Huxley and take “the word ‘soul’ simply as a name for the series of mental phenomena which make up an individual mind”?[7] Surely the moment we try distinctly to understand this question we realize that the cases are different. “Series of mental phenomena” for whom? For any passer-by such as might take stock of our biological dog? No, obviously only for that individual mind itself; yet that is supposed to be made up of, to be nothing different from, the series of phenomena. Are we, then, (1) quoting J. S. Mill's words, “to accept the paradox that something which ex hypothesi is but a series of feelings, can be aware of itself as a series”?[8] Or (2) shall we say that the several parts of the series are mutually phenomenal, much as A may look at B, who was just now looking at A? Or (3) finally, shall we say that a large part of the so-called series, in fact every term but one, is phenomenal for the rest—for that one?

As to the first, paradox is too mild a word for it; even contradiction will hardly suffice. It is as impossible to express “being aware of” by one term as it is to express an equation or any other relation by one term: what knows can no more be identical with what is known than a weight with what it weighs. If a series of feelings is what is known or presented, then what knows, what it is presented to, cannot be that series of feelings, and this without regard to the point Mill mentions, viz. that the infinitely greater part of the series is either past or future. The question is not in the first instance one of time or substance at all, but simply turns upon the fact that knowledge or consciousness is unmeaning except as it implies something knowing or conscious of something. But it may be replied: Granted that the formula for consciousness is something doing something, to put it generally; still, if the two somethings are the same when I touch myself or when I see myself, why may not agent and patient be the same when the action is knowing or being aware of; why may I not know myself—in fact, do I not know myself? Certainly not; agent and patient never are the same in the same act; such terms as self-caused, self-moved, self-known, et id genus omne, either connote the incomprehensible or are abbreviated expressions—as, e.g. touching oneself) when one's right hand touches one's left.

And so we come to the alternative: As one hand washes the other, may not different members of the series of feelings be subject and object in turn? Compare, for example, the state of mind of a man succumbing to temptation (as he pictures himself enjoying the coveted good and impatiently repudiates scruples of conscience or dictates of prudence) with his state when, filled with remorse, he sides with conscience and condemns this “former self”—the “better self” having meanwhile become supreme. Here the cluster of presentations and their associated sentiments and motives, which together played the role of self in the first situation, have—only momentarily it may be true, but still have—for the time the place of not-self; and under abnormal circumstances this partial alternation may become complete alienation, as in what is called “double consciousness.” Or again, the development of self-consciousness might be loosely described as taking the subject or self of one stage as an object in the next—self being, e.g. first identified with the body and afterwards distinguished from it. But all this, however true, is beside the mark; and it is really a very serious misnomer to speak, as e.g. Herbert Spencer does, of the development of self-consciousness as a “differentiation of subject and object.” It is, if anything, a differentiation of object and object, i.e. in plainer words, it is a differentiation among presentations—a differentiation every step of which implies just that relation to a subject which it is supposed to supersede.

There still remains the alternative, expressed in the words of J. S. Mill, viz. “the alternative of believing that the Mind or Ego is something different from any series of feelings or possibilities of them.” To admit this, of course, is to admit the necessity of distinguishing between Mind or Ego, meaning the unity or continuity of consciousness as a complex of presentations, and Mind or Ego as the subject to which this complex is presented. In dealing with the body from the ordinary biological standpoint no such necessity arises. But, whereas there the individual organism is spoken of unequivocally, in psychology, on the other hand, the individual mind may mean either (i.) the series of feelings or “mental phenomena” above referred to; or (ii.) the subject of these feelings for whom they are phenomena; or (iii.) the subject of these feelings or phenomena plus the series of feelings or phenomena themselves, the two being in that relation to each other in which alone the one is subject and the other a series of feelings, phenomena or objects. It is in this last sense that Mind is used in empirical psychology.[9] Its exclusive use in the first sense is favoured only by those who shrink from the speculative associations connected with its exclusive use in the second. But psychology is not called upon to transcend the relation of subject to object or, as we may call it, the fact of presentation. On the other hand, as has been said, the attempt to ignore one term of the relation is hopeless; and equally hopeless, even futile, is the attempt, by means of phrases such as consciousness or the unity of consciousness, to dispense with the recognition of a conscious subject.

5. We might now proceed to inquire more closely into the character and relations of the three invariable constituents of Feeling. psychical life which are broadly distinguished as cognitions, feelings and conations. But we should be at once confronted by a doctrine which, strictly taken, amounts almost to a denial of this tripartite classification of the facts of mind-the doctrine, viz. that feeling alone is primordial and invariably present wherever there is consciousness at all. Every living creature, it is said, feels, though it may never do any more; only the higher animals, and these only after a time, learn to discriminate and identify and to act with a purpose. This doctrine, as might be expected, derives its plausibility partly from the vagueness of psychological terminology, and partly from the intimate connexion that undoubtedly exists between feeling and cognition on the one hand and feeling and volition on the other. As to the meaning of the term, it is plain that further definition is requisite for a word that may mean (a) a touch, as feeling of roughness; (b) an organic sensation, as feeling of hunger; (c) an emotion, as feeling of anger; (d) feeling proper, as pleasure or pain. But, even taking feeling in the last, its stricter sense, it has been maintained that all the more complex forms of consciousness are resolvable into, or at least have been developed from, feelings of pleasure and pain. The only proof of such position, since we cannot directly observe the beginnings of conscious life, must consist of considerations such as the following. So far as we can judge, we find feeling everywhere; but, as we work downwards from higher to lower forms of life, the possible variety and the definiteness of sense-impressions both steadily diminish. Moreover, we can directly observe in our own organic sensations, which seem to come nearest to the whole content of primitive or infantile experience, an almost entire absence of any assignable quale. Finally, in our sense experience generally, we find the element of feeling at a maximum in the lower senses and the cognitive element at a maximum in the higher. But the so-called intellectual senses are the most used, and use (we know) blunts feeling and favours intellection, as we see in chemists, who sort the most filthy mixtures by smell and taste without discomfort. If, then, feeling predominates more and more as we approach the beginning of conscious life, may we not conclude that it is its only essential constituent? On the contrary, such a conclusion would be rash in the extreme. Two lines, e.g. may get nearer and nearer and yet will never meet, if the rate of approach is simply proportional to the distance. A triangle may be diminished indefinitely, and yet we cannot infer that it becomes eventually all angles, though the angles get no less and the sides do. Before, then, we decide whether pleasure or pain alone can ever constitute a complete experience, it may be well to inquire into the connexion between feeling and cognition, on the one hand, and between feeling and conation on the other, so far as we can now observe. And this is an inquiry which will help us towards an answer to our main question, namely, that concerning the nature and connexions of what are commonly regarded as the three ultimate facts of mind.

Broadly speaking, in any state of mind that we can directly observe, what we find is (1) that we are aware of a certain change Relation of Feeling to Cognition and Conation. in our sensations, thoughts or circumstances, (2) that we are pleased or pained with the change, and (3) that we act accordingly. We never find that feeling directly alters—i.e. without the intervention of the action of which it prompts—either our sensations or situation, but that regularly these latter with remarkable promptness and certainty alter it. We have not first a change of feeling, and then a change in our sensations, perceptions and ideas; but, these changing, change of feeling follows. In short, feeling appears to be an effect, which therefore cannot exist without its cause, though in different circumstances the same immediate cause may produce a different amount or even a different state of feeling. Turning from what we may call the receptive phase of an experience to the active or appetitive phase, we find in like manner that feeling is certainly not—in such cases as we can clearly observe—the whole of what we experience at any moment. True, in common speech we talk of liking pleasure and disliking pain; but this is either tautology, equivalent to saying we are pleased when we are pleased and pained when we are pained; or else it is an allowable abbreviation, and means that we like pleasurable objects and dislike painful objects, as when we say we like feeling warm and dislike feeling hungry. But feeling warm or feeling hungry, we must remember, is not pure feeling in the stricter sense of the word. Within the limits of our observation, then, we find that feeling accompanies some more or less definite presentation which for the sake of it becomes the object of appetite or aversion; in other words, feeling implies a relation to a pleasurable or painful presentation or situation, that, as cause of feeling or as end of the action to which feeling prompts, is doubly distinguished from it. Thus the very facts that lead us to distinguish feeling from cognition and conation make against the hypothesis that consciousness can ever be all feeling.

But, as already said, the plausibility of this hypothesis is in good part due to a laxity in the use of terms. Most psychologists Feeling and Sensation distinct. before Kant, and some even to the present day, speak of pleasure and pain as sensations. But it is plain that pleasure and pain are not simple ideas, as Locke called them, in the sense in which touches and tastes are—that is to say, they are never like these localized or projected, nor are they elaborated in conjunction with other sensations and movements into percepts or intuitions of the external. This confusion of feeling with sensations is largely consequent on the use of one word pain both for certain organic sensations and for the purely subjective state of being pained. But such pains not only are always more or less definitely localized—which of itself is so far cognition, they are also distinguished as shooting, burning, gnawing, &c., all which symptoms indicate a certain objective quality. Accordingly psychologists have been driven by one means or another to recognize two “aspects” (Bain), or “properties” (Wundt), in what they call a sensation, the one a “sensible or intellectual” or “qualitative,” the other an “affective” or “emotive,” aspect or property. The term “aspect” is figurative and obviously inaccurate; even to describe pleasure and pain as properties of sensation is a matter open to much question. But the point which at present concerns us is simply that when feeling is said to be the primordial element in consciousness more is usually included under feeling than pure pleasure and pain, viz. some characteristic or quality by which one pleasurable or painful sensation is distinguishable from another. No doubt, as we go downwards in the chain of life the qualitative or objective elements in the so-called sensations become less and less definite; and at the same time organisms with well-developed sense-organs give place to others without any clearly differentiated organs at all. But there is no ground for supposing even the amoeba itself to be affected in all respects the same whether by changes of temperature or of pressure or by changes in its internal fluids, albeit all of these changes will further or hinder its life and so presumably be in some sort pleasurable or painful. On the whole, then, there are grounds for saying that the endeavour to represent all the various facts of consciousness as evolved out of feeling is due to a hasty striving after simplicity, and has been favoured by the ambiguity of the term feeling itself. If by feeling we mean a certain subjective state varying continuously in intensity and passing from time to time from its positive phase (pleasure) to its negative phase (pain), then this purely pathic state implies an agreeing or disagreeing something which psychologically determines it. If, on the other hand, we let feeling stand for both this state and the cause of it, then, perhaps, a succession of such “feelings” may make up a consciousness; but then we are including two of our elementary facts under the name of one of them. The simplest form of psychical life, therefore, involves not only a subject feeling but a subject having qualitatively distinguishable presentations which are the occasion of its feeling.

6. We may now try to ascertain what is meant by cognition as an essential element in this life, or, more exactly, what we are to understand by the term presentation. It was an important step onwards for psychology when Locke introduced that “new way of ideas” which Stillingfleet found alternately so amusing and so dangerous. By ideas Presentation. Locke told him he meant “nothing but the immediate objects of our minds in thinking”; and it was so far a retrograde step when Hume restricted the term to certain only of these objects, or rather to these objects in a certain state, viz. as reproduced ideas or “images.” And, indeed, the history of psychology seems to show that its most important advances have been made by those who have kept closely to this way of ideas; the establishment of the laws of association with their many fruitful applications and the whole Herbartian psychology may suffice as instances (see Herbart). The truth is that the use of such a term is itself a mark of an important generalization, one which helps to free us from the mythology and verbiage of the “faculty-psychologists.” All the various mental facts spoken of as sensations, movements, percepts, images, intuitions, concepts, notions, have two characteristics in common: (1) they admit of being more or less attended to, and (2) they can be variously combined together and reproduced. It is here proposed to use the term presentation to denote them all, as being the best English equivalent for what Locke meant by idea and what Kant and Herbart called a Vorstellung.

A presentation has then a twofold relation—first, directly to the subject, and, secondly, to other presentations. The former relation answers to the fact that a presentation is attended to, that the subject is more or less conscious of it: it is “in his mind” or presented. As presented to a subject a presentation might with advantage be called an object, or perhaps a psychical object, to distinguish it from what are called objects apart from presentation, i.e. conceived as independent of any particular subject. Locke, as we have seen, did so call it; still, to avoid possible confusion, it may turn out best to dispense with the frequent use of object in this sense. But on one account, at least, it is desirable not to lose sight altogether of this, which is after all the stricter as well as the older signification of object, namely, because it enables us to express definitely, without implicating any ontological theory, what we have so far seen reason to think is the fundamental fact in experience. Instead of depending mainly on that vague and treacherous word “consciousness,” or committing ourselves to the position that ideas are modifications of a certain mental substance or identical with the subject to whom they are presented, we may leave all this on one side, and say that ideas are objects, and the relation of objects to subjects—that whereby the one is object and the other subject—is presentation; and it is because only objects sustain this relation that they may be spoken of simply as presentations. On the side of the subject this relation implies what, for want of a better word, may be called attention, extending the denotation Attention. of this term so as to include even what we ordinarily call inattention. Attention so used will thus cover part of what is meant by consciousness—so much of it, that is, as answers to being mentally active, active enough at least to “receive impressions.” Attention on the side of the subject implies intensity on the side of the object: we might indeed almost call intensity the matter of a presentation, without which it is a nonentity.[10]

The inter-objective relations of presentations, on which their second characteristic, that of revivability and associability depends, though of the first importance in themselves, hardly call for examination in a general analysis like the present. But there is one point still more fundamental that we cannot wholly pass by: it Continuity of Consciousness. is—in part at any rate—what is commonly termed the unity or continuity of consciousness. From the physical standpoint and in ordinary life we can talk of objects that are isolated and independent and in all respects distinct individuals. The screech of the owl, for example, has physically nothing to do with the brightness of the moon: either may come or go without changing the order of things to which the other belongs. But psychologically, for the individual percipient, they are parts of one whole; the more his attention is given to the one the more it is taken from the other. Also the actual recurrence of the one will afterwards entail the re-presentation of the other also. Not only are they still parts of one whole, but such distinctness as they have at present is the result of a gradual differentiation.

It is quite impossible for us now to imagine the effects of years of experience removed, or to picture the character of our infantile presentations before our interests had led us habitually to concentrate attention on some and to ignore others. In place of the many things which we can now see and hear, not merely would there then be a confused presentation of the whole field of vision and of a mass of undistinguished sounds, but even the difference between sights and sounds themselves would be without its present distinctness. Thus the further we go back the nearer we approach to a total presentation having the character of one general continuum in which differences are latent. There is, then, in psychology, as in biology, what may be called a principle of “progressive differentiation or specialization”;[11] and this, as well as the facts of reproduction and association, forcibly suggests the conception of a certain objective continuum forming the background or basis to the several relatively distinct presentations that are elaborated out of it—the equivalent, in fact, of that unity and continuity of consciousness which has been supposed to supersede the need for a conscious subject.

There is one class of objects of special interest even in a general survey, viz. movements or motor presentations. These, like sensory presentations, admit of association and reproduction, and seem to attain to such distinctness as they possess in adult human experience by a gradual differentiation out of an original diffused mobility which Motor Presentations. is little besides emotional expression. Of this, however, more presently. It is primarily to such dependence upon feeling that movements owe their distinctive character, the possession, that is, under normal circumstances, of definite and assignable psychical antecedents, in contrast to sensory presentations, which are devoid of them. We cannot psychologically explain the order in which particular sights and sounds occur; but the movements that follow them, on the other hand, can be adequately explained only by psychology. The twilight that sends the hens to roost sets the fox to prowl, and the lion's roar Subjective Selection. which gathers the jackals scatters the sheep. Such diversity in the movements, although the sensory presentations are similar, is due, in fact, to what we might call the principle of “subjective or hedonic selection”—that, out of all the manifold changes of sensory presentation which a given individual experiences, only a few are the occasion of such decided feeling as to become objects of possible appetite or aversion. It is thus by means of movements that we are more than the creatures of circumstances and that we can with propriety talk of subjective selection. The representation of what interests us comes then to be associated with the representation of such movements as will secure its realization, so that—although no concentration of attention will secure the requisite intensity to a pleasurable object present only in idea—we can by what is strangely like a concentration of attention convert the idea of a movement into the fact, and by means of the movement attain the coveted reality.

7. And this has brought us round naturally to the third of the commonly accepted constituents of experience. What is conation, or rather conative action? For there are two Conation. questions often more or less confused, the question of motive or spring of action, as it is sometimes called—why is there action at all? and the question of means—how do definite actions come about? The former question relates primarily to the connexion of conation and feeling. It is only the latter question that we now raise. In ordinary voluntary movement we have first of all an idea or re-presentation of the movement, and last of all the actual movement itself—a new presentation which may for the present be described as the filling out of the re-presentation, which thereby attains that intensity, distinctness and embodiment we call reality. How does this change come about? The attempt has often been made to explain it by a reference to the more uniform, and apparently simpler, case of reflex action, including under this term what are called sensori-motor and ideo-motor actions. In all these the movement seems to be the result of a mere transference of intensity from the associated sensation or idea that sets on the movement. But when by some chance or mischance the same sensory presentation excites two or more nascent motor changes that conflict, a temporary block is said to occur; and, when at length one of these nascent motor changes finally prevails, then, it is said, “there is constituted a state of consciousness which displays what we term volition.”[12] But this assumption that sensory and motor ideas are associated before volition, and that volition begins where automatic or reflex action ends, is due to that inveterate habit of confounding the psychical and the physical which is the bane of modern psychology. How did these particular sensory and motor presentations ever come to be associated? The only psychological evidence we have of any very intimate connexion between sensory and motor representations is that furnished by our acquired dexterities, i.e. by such movement as Hartley[13] styled “secondarily automatic.” But then all these have been preceded by volition: as Herbert Spencer says, “the child learning to walk wills each movement before making it.” Surely, then, a. psychologist should take this as his typical case and prefer to assume that all automatic actions that come within his ken at all are in this sense secondarily automatic, i.e. to say that either in the experience of the individual or of his ancestors, volition or something analogous to it, preceded habit.

But, if we are thus compelled by a sound method to regard sensori-motor actions as degraded or mechanical forms of voluntary actions, instead of regarding voluntary actions as gradually differentiated out of something physical, we have not to ask: What happens when one of two alternative movements is executed? but the more general question: What happens when any movement is made in consequence of feeling? It is obvious that on this view the simplest definitely purposive movement must have been preceded by some movement simpler still. For any distinct movement purposely made presupposes the ideal presentation, before the actual realization, of the movement. But such ideal presentation, being a re-presentation, equally presupposes a previous actual movement of which it is the so-called mental residuum. There is then, it would seem, but one way left, viz. to regard those movements which are immediately expressive of pleasure or pain as primordial, and to regard the so-called voluntary movements as elaborated out of these. The vague and diffusive character of these primitive emotional manifestations is really a point in favour of this position. For such “diffusion” is evidence of an underlying continuity of motor presentations parallel to that already discussed in connexion with sensory presentations, a continuity which, in each case, becomes differentiated in the course of experience into comparatively distinct and discrete movements and sensations respectively.[14]

But whereas we can only infer, and that in a very roundabout fashion, that our sensations are not absolutely distinct but are parts of one massive sensation, as it were, we are still liable under the influence of strong emotion directly to experience the corresponding continuity in the case of movement. Such motor-continuum we may suppose is the psychical counterpart of that permanent readiness to act, or rather that continual nascent acting, which among the older physiologists was spoken of as “tonic action.” This “skeletal tone,” as it is now called, is found to disappear more or less completely from a limb when its sensory nerves are divided. “In the absence of the usual stream of afferent impulses passing into it, the spinal cord ceases to send forth the influences which maintain the tone.”[15] And a like intimate dependence, we have every reason to believe, obtains throughout between sensation and movement. We cannot imagine the beginning of life but only life begun. The simplest picture, then, which we can form of a concrete state of mind is not one in which there are movements before there are any sensations or sensations before there are any movements, but one in which change of sensation is followed by change of movement, the link between the two being a change of feeling.

Having thus simplified the question, we may now ask again: How is this change of movement through feeling brought about? Dependence of Action of Feeling. The answer, as already hinted, appears to be: By a change of attention. We learn from such observations as psychologists describe under the head of fascination, imitation, hypnotism, &c., that the mere concentration of attention upon a movement is often enough to bring the movement to pass. But, of course, in such cases neither emotion nor volition is necessarily implied; but none the less they show the close connexion that exists between attention and movement. Everybody, too, must often have observed how the execution of any but mechanical movements arrests attention to thoughts or sensations, and how, vice versa, a striking impression or thought interrupts him in the performance of skilled movements. Let us suppose, then, that we have at any given moment a certain distribution of attention between sensory and motor presentations; a change in that distribution then will mean a change in the intensity of some of all of these. But, in the case of motor presentations, change of intensity means change of movement. Such changes are, however, quite minimal in amount so long as the given presentations are not conspicuously agreeable or disagreeable. So soon as they are, however, there is evidence of a most intimate connexion between feeling and attention; but it is hardly possible adequately to exhibit this evidence without first attempting to ascertain the characteristics of the presentations, or groups of presentations, that are respectively pleasurable and painful, and this must occupy us later on.

8. We are now at the end of our analysis, and the results may perhaps be most conveniently summarized by first throwing Primordial Facts of Mind. them into a tabular form and then appending a few remarks by way of indicating the main purport of the table. Taking no account of the specific difference between one concrete state of mind and another, and supposing that we are dealing with presentations in their simplest form, i.e. as sensations and movements, we have:—

A SUBJECT
(1) non-voluntarily attending
to changes in the
sensory-continuum;[16]
= Presentation
of sensory
OBJECTS.
[Cognition]
(2) being, in consequence,
either pleased or pained;
[Feeling]
and (3) by voluntary attention
or “innervation”
producing changes in
the motor-continuum.[16]
= Presentation
of motor
[Conation]

Of the three phases or functions, thus analytically distinguishable, but not really separable, the first and the third correspond in the main with the receptive and active states or powers of the older psychologists. The second, being more difficult to isolate, was long overlooked; or, at all events, its essential characteristics were not distinctly marked, so that it was confounded either with (1) which is its cause, or with (3), its effect. But perhaps the most important of all psychological distinctions is that which traverses both the old bipartite and the prevailing tripartite analysis, viz. that between the subject on the one hand, as acting and feeling, and the objects of this activity on the other. With this distinction clearly before us, instead of crediting the subject with an indefinite number of faculties or capacities, we must seek to explain not only reproduction, association, &c., but all varieties of thinking and acting, by the laws pertaining to ideas or presentations, leaving to the subject only the one power of variously distributing that attention upon which the intensity of a presentation in part depends. What we call activity in the narrower sense (as e.g. purposive movement and intellection) is but a special form of this single subjective activity, although a very important one.

According to this view, then, presentations, attention, feeling, are not to be regarded as three co-ordinate genera, each of which is a complete “state of mind or consciousness,” i.e. as being all alike included under this one supreme category. There is, as Berkeley long ago urged, no resemblance between activity and an idea; nor is it easy to see anything common to pure feeling and an idea, unless it be that both possess intensity. Classification seems, in fact, to be here out of place. Instead, therefore, of the one summum genus, state of mind or consciousness, with its three co-ordinate subdivisions—cognition, emotion, conation—our analysis seems to lead us to recognize three distinct and irreducible components—attention, feeling, and objects or presentations—as together, in a certain connexion, constituting one concrete state of mind or psychosis. Of such concrete states of mind or psychoses we may then say—so far agreeing with the older, bipartite psychology—that there are two forms, corresponding to the two ways in which attention may be determined and the two classes of objects attended to in each, viz. (1) the sensory or receptive attitude, when attention is non-voluntarily determined, i.e. where feeling follows the act of attention; and (2) the motor or active attitude, where feeling precedes the act of attention, which is thus determined voluntarily.

Attention.

9. Instead of a congeries of faculties we have assumed a single subjective activity and have proposed to call this attention. Some further explication of this position seems to be desirable. We start with the duality of subject and object as fundamental. We say of man, mouse, or monkey that it feels, perceives, remembers, infers, strives, and so forth. Leaving aside the first term, it is obvious that all the rest imply both an activity and an object. Is it possible to resolve these instances into a form in which the assumed diversity of the act will appear as a diversity of the object? At first sight it looks rather as if the kind of activity might vary while the object remained the same; that e.g. we perceived an object and later on remembered or desired it. It would then be most natural to refer these several activities to corresponding faculties of perception, memory and desire. This, indeed, is the view embodied in common speech, and for practical purposes it is doubtless the simplest and the best. Nevertheless, a more thorough analysis shows that when the supposed faculty is different the object is never entirely and in all respects the same. Thus in perception, e.g. we deal with “impressions” or primary presentations, and in memory and imagination with “ideas” (in the later sense) or secondary presentations. In desire the want of the object gives it an entirely different setting, adding a new characteristic, that of value or worth, so that its acquisition becomes the end of a series of efforts or movements. The older psychology, by its acceptance of the Cartesian doctrine that all the facts of immediate experience are to be interpreted as subjective modifications, failed to distinguish adequately between the subject as active and the objects of its activity. Hence the tendency to rest content with the popular distinction of various faculties in spite of the underlying sameness implied in the common application of “conscious” to them all. In fact, Locke's definition of idea (in the older and wider sense) as the immediate object of consciousness or thinking was censured by Reid as “the greatest blemish in the Essay on Human Understanding.” But, accepting this definition as implied in the duality of subject and object, and accepting too the underlying sameness which the active form “conscious” undeniably implies, we have simply to ask: “Which is the better term to denote this common element—consciousness or attention?”

Consciousness, as the vaguest, most protean and most treacherous of psychological terms, will hardly serve our purpose. Attention, on the other hand, has an invariable active sense, and there is an appropriate verb, to attend. But many things, it may be said, are presented while few are attended to; if attention is to be made coextensive with the activity implied in consciousness, will not the vital distinction between attention and inattention be lost? In fact, however, this distinction implies a covert comparison, not an absolute contrast. In everyday life we recognize many degrees of attention, ranging from an extreme of intense concentration to one of complete remission, as Locke long ago pointed out.[17] Between these extremes there is perfect continuity, and not a difference of kind; to apply the one term attention to the whole range is very like applying the one term magnitude to large and small quantities alike.

But it is not enough to show that when we commonly talk of different faculties we also find psychological differences of object, and to assert that if there is one common factor in all psychical activity this factor is attention. To make our position secure it is needful to show directly that all the various faculties with which a subject can be credited are resolvable into attention and various classes or relations or states of presentations that are attended to. How far this is possible remains to be seen as we proceed. In the case of the so-called “intellectual powers” the position is generally conceded, but so far as the voluntary or active powers are concerned it is as generally denied. Now, in so far as volition implies not merely action, overt or intended, but also motives, in so far also it must be acknowledged it contains a factor not resolvable into attention to motor presentations. This further factor, which has been called “the volitional character of feeling,” we here leave aside. Apart from this direct spring of action, then, the question is whether the active process itself differs from the cognitive or receptive process save in being attention to a special class of objects. First of all, it is noteworthy that both have the same characteristics. Thus, what Hamilton called “the law of limitation” holds of each alike and of either with respect to the other; and it holds too not only of the number of presentations but also of the intensity. We can be absorbed in action just as much as in perception or thought; also, as already said, movements, unless they are mechanical, inhibit ideas; and vice versa, ideas, other than associated trains, arrest movements. Intoxication, hypnotism or insanity, rest or exhaustion, tell on apperception as well as on innervation. The control of thoughts, equally with the control of movements, requires effort; and as there is a strain peculiar to intently listening or gazing, which is known to have a muscular concomitant, so too there is a strain characteristic of recollection and visualization, which may quite well turn out to be muscular too. When movements have to be associated, the same continuous attention is called for as is found requisite in associating sensory impressions; and, when such associations have become very intimate, dissociation is about equally difficult in both cases.

There is one striking fact that brings to light the essential sameness of apperception and innervation, cited by Wundt for this very purpose. In so-called “reaction-time” experiments it is found, when the impression to be registered follows on a premonitory signal after a certain brief interval, that then the reaction (registering the impression) is often instantaneous; the reaction-time, in other words, is nil. In such a case the subject is aware not of three separate events, (1) the perception of the impression; (2) the reaction; (3) the perception of this; but the fact of the impression is realized and the registering movement is actualized at once and together the subject is conscious of one act of attention and one only.

Theory of Presentations.

10. We come now to the exposition of the objects of attention or consciousness, i.e. to what we may call the objective or presentational factor of psychical life. The treatment of this will fall naturally into two divisions. In the first we shall have to deal with its general characteristics and with the fundamental processes which all presentation involves. In view of its general and more or less hypothetical character we may call it the theory of presentation. We can then pass on to the special forms of presentations, known as sensations, percepts, images, &c., and to the special processes to which these forms lead up.

This exposition will be simplified if we start with a supposition that will enable us to leave aside, at least for the present, the Assumption of a Psychological Individual. difficult question of heredity. We know that in the course of each individual's life there is more or less of progressive differentiation or development. Further, it is believed that there has existed a series of sentient individuals beginning with the lowest form of life and advancing continuously up to man. Some traces of the advance already made may be reproduced in the growth of each human being now, but for the most part such traces have been obliterated. What was experience in the past has become instinct in the present. The descendant has no consciousness of his ancestor's failures when performing by “an untaught ability” what they slowly and perhaps painfully acquired. But, if we are to attempt to follow the genesis of mind from its earliest dawn, it is the primary experience rather than the eventual instinct that we have first of all to keep in view. To this end, then, it is proposed to assume that we are dealing with one individual who has continuously advanced from the beginning of psychical life, and not with a series of individuals of whom all save the first inherited certain capacities from their progenitors. The life-history of such an imaginary individual, that is to say, would correspond with all that was new in the experience of a certain typical series of individuals each of whom advanced a certain stage in mental differentiation. On the other hand, from this history would be omitted that inherited reproduction of the net results, so to say, of ancestral experience, that innate tradition by which alone, under the actual conditions of existence, progress is possible. The process of thus reproducing the old might differ as widely from that of producing the new as electrotyping does from engraving. However, the point is that as psychologists we know nothing directly about it; neither can we distinguish precisely at any link in the chain of life what is old and inherited—original in the sense of Locke and Leibnitz——from what is new or acquired—original in the modern sense. But we are bound as a matter of method to suppose all complexity and differentiation among presentations to have been originated, i.e. experimentally acquired, at some time or other. So long, then, as we are concerned primarily with the progress of this differentiation we may disregard the fact that it has not actually been, as it were, the product of one hand dealing with one tabula rasa to use Locke's—originally Aristotle's—figure, but of many hands, each of which, starting with a reproduction of what had been wrought on the preceding tabulae, put in more or fewer new touches before devising the whole to a successor who would proceed in like manner.

11. What is implied in this process of differentiation and what is it that becomes differentiated?—these are the questions The Presentation Continuum. to which we must now attend. Psychologists have usually represented mental advance as consisting fundamentally in the combination and recombination of various elementary units, the so-called sensations and primitive movements: in other words, as consisting in a species of “mental chemistry.” If we are to resort to physical analogies at all—a matter of very doubtful propriety—we shall find in the growth of a seed or an embryo far better illustrations of the unfolding of the contents of consciousness than in the building up of molecules: the process seems much more a segmentation of what is originally continuous than an aggregation of elements at first independent and distinct. Comparing higher minds or stages of mental development with lower—by what means such comparison is possible we need not now consider—we find in the higher conspicuous differences between presentations which in the lower are indistinguishable or absent altogether. The worm is aware only of the difference between light and dark. The steel-worker sees half a dozen tints where others see only a uniform glow. To the child, it is said, all faces are alike; and throughout life we are apt to note the general, the points of resemblance, before the special, the points of difference. But even when most definite, what we call a presentation is still part of a larger whole. It is not separated from other presentations, whether simultaneous or successive, by something which is not of the nature of presentation, as one island is separated from another by the intervening sea, or one note in a melody from the next by an interval of silence. In our search for a theory of presentations, then, it is from this “continuity of consciousness” that we must take our start. Working backwards from this as we find it now, we are led alike by particular facts and general considerations to the conception of a totum objectivum or objective continuum which is gradually differentiated, thereby giving rise to what we call distinct presentations, just as some particular presentation, clear as a whole, as Leibnitz would say, becomes with mental growth a complex of distinguishable parts. Of the very beginning of this continuum we can say nothing; absolute beginnings are beyond the pale of science. Experience advances as this continuum is differentiated, every differentiation being a change of presentation. Hence the commonplace of psychologists—We are only conscious as we are conscious of change.

But “change of consciousness” is too loose an expression to take the place of the unwieldy phrase differentiation of a Gradual Differentiation of Presentation-Continuum. presentation-continuum, to which we have been driven. For not only does the term “consciousness” confuse what exactness requires us to keep distinct, an activity and its object, but also the term “change” fails to express the characteristics which distinguish new presentations from other changes. Differentiation implies that the simple becomes complex or the complex more complex; it implies also that this increased complexity is due to the persistence of former changes; we may even say such persistence is essential to the very idea of development or growth. In trying, then, to conceive our psychological individual in the earliest stages of development we must not picture him as experiencing a succession of absolutely new sensations, which, coming out of nothingness, admit of being strung upon the “thread of consciousness” like beads picked up at random, or cemented into a mass like the bits of stick and sand with which the young caddis covers its nakedness. The notion, which Kant has done much to encourage, that psychical life begins with a confused manifold of sensations—devoid not only of logical but even of psychological unity—is one that becomes more inconceivable the more closely we consider it. An absolutely new presentation, having no sort of connexion with former presentations till the subject has synthesized it with them, is a conception for which it would be hard to find a warrant either by direct observation, by inference from biology, or in considerations of an a priori kind. At any given moment we have a certain whole of presentations, a “field of consciousness,” psychologically one and continuous; at the next we have not an entirely new field but a partial change within this field. Many who would allow this in the case of representations, i.e. where idea succeeds idea by the workings of association, would demur to it in the case of primary presentations or sensations. “For,” they would say, “may not silence be broken by a clap of thunder, and have not the blind been made to see?” To urge such objections is to miss the drift of our discussion, and to answer them may serve to make it clearer. Where silence can be broken there are representations of preceding sounds and in all probability even subjective presentations of sound as well; silence as experienced by one who has heard is very different from the silence of Condillac's statue before it had ever heard. The question is rather whether such a conception as that of Condillac's is possible; supposing a sound to be, qualitatively, entirely distinct from a smell, could a field of consciousness consisting of smells be followed at once by one in which sounds had part? And, as regards the blind coming to see, we must remember not only that the blind have eyes but that they are descended from ancestors who could see. What nascent presentations of sight are thus involved it would be hard to say; and the problem of heredity is one that we have for the present left aside.

The view here taken is (1) that at its first appearance in psychical life a new sensation or so-called elementary presentation is really a partial modification of some pre-existing presentation which thereby becomes as a whole more complex than it was before; and (2) that this complexity and differentiation of parts never become a plurality of discontinuous presentations, having a distinctness and individuality such as the atoms or elementary particles of the physical world are supposed to have. Beginners in psychology, and some who are not beginners, are apt to be led astray by expositions which set out from the sensations of the special senses, as if these furnished us with the type of an elementary presentation. The fact is we never experience a mere sensation of colour, sound, touch, and the like; and what the young student mistakes for such is really a perception, a sensory presentation combined with various sensory and motor presentations and with representations—and having thus a definiteness and completeness only possible to complex presentations. Moreover, if we could attend to a pure sensation of sound or colour by itself, there is much to justify the suspicion that even this is complex and not simple, and owes to such complexity its clearly marked specific quality. In certain of our vaguest and most diffused organic sensations there is probably a much nearer approach to the character of the really primitive presentations.

In such sensations we can distinguish three variations, viz. variations of quality, of intensity, and of what Bain called Diffusion and Restriction. massiveness, or, as we shall say, extensity. This last characteristic, which everybody knows who knows the difference between the ache of a big bruise and the ache of a little one, between total and partial immersion in a bath, is, as we shall see later on, an essential element in our perception of space. But it is certainly not the whole of it, for in this experience of massive sensation alone it is impossible to find other elements which an analysis of spatial intuition unmistakably yields. Extensity and extension, then, are not to be confounded. Now, we find, even at our level of mental evolution, that an increase in the intensity of a sensation is apt to entail an increase in its extensity too. In like manner we observe a greater extent of movement in emotional expression when the intensity of the emotion increases. Even the higher region of imagination is no exception, as is shown by the whirl and confusion of ideas incident to delirium, and, indeed, to all strong excitement. But this “diffusion” or “radiation,” as it has been called, diminishes as we pass from the class of organic sensations to the sensations of the five senses, from movements expressive of feeling to movements definitely purposive, and from the tumult of ideas excited by passion to the steadier sequences determined by efforts to think. Increased differentiation seems, then, to be intimately connected with increased “restriction.” Probably there may be found certain initial differentiations which for psychology are ultimate facts that it cannot explain. As already said, the very beginning of experience is beyond us, though it is our business—working from within—to push back our analysis as far as we can. But some differentiations being given, then it may be safely said that, in accordance with what we have called the principle of subjective selection (see § 6), attention would be voluntarily concentrated upon certain of these and upon the voluntary movements specially connected with them. To such subjectively initiated modifications of the presentation-continuum, moreover, we may reasonably suppose “restriction” to be in large measure due. But increased restriction would render further differentiation of the given whole of presentation possible, and so the two processes might supplement each other. These processes have now proceeded so far that at the level of human consciousness we find it hard to form any tolerably clear conception of a field of consciousness in which an intense sensation, no matter what, might—so to say—diffuse over the whole. Colours, e.g. are with us so distinct from sounds that—except as regards the excitement of attention or the drain upon it—there is nothing in the intensest colour to affect the simultaneous presentation of a sound. But at the beginning whatever we regard as the earliest differentiation of sound might have been incopresentable with the earliest differentiation of colour, if sufficiently diffused, much as a held of sight all blue is now incopresentable with one all red. Or, if the stimuli appropriate to both were active together, the resulting sensation might have been not a blending of two qualities, as purple is said to be a blending of red and violet, but rather a neutral sensation without the specific qualities of either. Now, on the other hand, colours and sounds are necessarily so far localized that we are directly aware that the eye is concerned with the one and the ear with the other. This Incopresentability. brings to our notice a fact so ridiculously obvious it has never been deemed worthy of mention, although it has undeniably important bearings—the fact, viz. that certain sensations or movements are an absolute bar to the simultaneous presentation of other sensations or movements. We cannot see an orange as at once yellow and green, though we can feel it at once as both smooth and cool; we cannot open and close the same hand at the same moment, but we can open one hand while closing the other. Such incopresentability or contrariety is thus more than mere difference, and occurs only between presentations belonging to the same sense or to the same group of movements. Strictly speaking, it does not always occur even then; for red and yellow, hot and cold, are presentable together provided they have certain other differences which we shall meet again presently as differences of “local sign.”

12. In the preceding paragraphs we have had occasion to distinguish between the presentation-continuum or whole field Retentiveness. of consciousness, as we may for the present call it, and those several differentiations within this field which are ordinarily spoken of as presentations, and to which—now that their true character as parts is clear—we too may confine the term. But it will be well in the next place, before inquiring more closely into their characteristics, to consider for a moment that persistence of preceding modifications which the principle of progressive differentiation implies. This persistence is best spoken of as retentiveness. It is often confused with memory, though this is something much more complex and special; for in memory there is necessarily some contrast of past and present, whereas here there is simply the persistence of the old. But what is it that persists? On our theory we must answer, the continuum as differentiated, not the particular differentiation as an isolated unit. If psychologists have erred in regarding the presentations of one moment as merely a plurality of units, they have erred in like manner concerning the so-called residua of such presentations. As we see a certain colour or a certain object again and again, we do not go on accumulating images or representations of it, which are somewhere crowded together like shades on the banks of the Styx; nor is such colour, or whatever it be, the same at the hundredth time of presentation as at the first, as the hundredth impression of a seal on wax would be. There is no such lifeless fixity in mind. The explanations of perception most in vogue are far too mechanical and, so to say, atomistic; but we must fall back upon the unity and continuity of our presentation continuum if we are to get a better. Suppose that in the course of a few minutes we take half a dozen glances at a strange and curious flower. We have not as many complex presentations which we might symbolize as F₁, F₂,....F₆. But rather, at first only the general outline is noted, next the disposition of petals, stamens, &c., then the attachment of the anthers, position of the ovary, and so on; that is to say, symbolizing the whole flower as [p′ (ab) s′ (cd) o′ (fg)], we first apprehend say [p′ . . s′ . . o′], then [p′ (ab) s′ . . o′.], or [p′ (a . .) s′ (c . .) o′ (f . .)], and so forth. It is because the traits first attended to persist that the later form an addition to them till the complex is at length complete. There is nothing in this instance properly answering to what are known as the reproduction and association of ideas; in the last and complete apprehension as much as in the first vague and inchoate one the flower is there as a primary presentation. There is a limit, of course, to such a procedure, but the instance taken, we may safely say, is not such as to exceed the bounds of a simultaneous field of consciousness. Assuming then that such increase of differentiation through the persistence of preceding differentiations holds of the presentation-continuum as a whole, we conclude that, in those circumstances in which we now have a specific sensation of, say, red or sweet, there would be for some more primitive experience nothing but a vague, almost organic, sensation, which, however, would persist, so that on a repetition of the circumstances it could be again further differentiated. The earlier differentiations, in short, do not disappear like the waves of yesterday in the calm of today, nor yet last on like old scars beside new ones; but rather the two are blended and combined, so that the whole field of consciousness, like a continually growing picture, increases indefinitely in complexity of pattern.

13. Assimilation.—This process, in which later differentiations blend with and thereby further restrict and specialize what is retained of earlier and less definite presentations, is thus a further implication of the principle of the progressive development of the presentational continuum. When not ignored altogether this further process has been commonly regarded as merely a simple form of “association,” its peculiarity being, as it was supposed, that the presentations associated—though numerically distinct—were in quality perfectly identical. In point of fact, both these assumptions seem to be erroneous and due to the so-called psychologist's fallacy.[18] For the experiencing subject there is apparently at this stage-as we have already urged-neither the numerical distinctness nor the qualitative identity which the words “past impression (A₁)” and “present impression (A₂)” suggest. Still the connexion between this process of mere blending or fusion, which we shall call assimilation, and the process of association proper is so close, and the detailed analysis called for so complex, that we must needs defer further discussion till we come to treat of association as a whole (cf. below, § 24). It may then be possible to show that we have here to do with a process much simpler and more fundamental than association. But it is at least clear at once that if the term association is to be correctly used it will imply that the presentations associated are from the first distinct, are attended to as distinct, are associated solely in consequence of such attention, and remain to the last distinguishable.

In view of the intimate connexion between differentiation, retentiveness and assimilation it will sometimes be convenient to refer to all three together as constituting what we may call the plasticity of the presentational continuum.

14. This will be the most convenient place to take note of Relativity. certain psychological doctrines which, though differing in some material respects, are usually included under the term Law of Relativity.

a. Hobbes's Sentire semper idem et non sentire ad idem recidunt is often cited as one of the first formulations of this law; and if we take it to apply to the whole field of consciousness it becomes at once true and trite: a field of consciousness unaltered either by change of impression or of idea would certainly be a blank and a contradiction. Understood in this sense the Law of Relativity amounts to what Hamilton called the Law of Variety: “that we are conscious only as we are conscious of difference.”[19] But, though consciousness involves change, it is still possible that particular presentations in the field of consciousness may continue unchanged indefinitely. When it is said that “a constant impression is the same as a blank,” what is meant turns out to be something not psychological at all, as, e.g., our insensibility to the motion of the earth or to the pressure of the air—cases in which there is obviously no presentation, nor even any evidence of nervous change. Or else this paradox proves to be but an awkward way of expressing what we may call accommodation, whether physiological or psychological. Thus the skin soon adapts itself to certain seasonal alterations of temperature, so that heat or cold ceases to be felt: the sensation ceases because the nervous change, its proximate physical counterpart, has ceased. Again, there is what James Mill calls “an acquired incapacity of attention,” such that a constant noise, for example, in which we have no interest, is soon inaudible. In such a case of psychological accommodation we should expect also to find on the physiological side some form of central reflection or isolation more or less complete. As a rule, no doubt, impressions do not continue constant for more than a very short time; still there are sad instances enough in the history of disease, bodily and mental, to show that such a thing can quite well happen, and that such constant impressions (and “fixed ideas,” which are in effect tantamount to them), instead of becoming blanks, may dominate the entire consciousness, colouring or bewildering everything.

b. From the fact that the field of consciousness is continually changing it has been supposed to follow, not only that a constant presentation is impossible, but—as a further consequence—that every presentation is essentially nothing but a transition or difference. “All feeling,” says Bain, the leading exponent of this view, “is two-sided. . . . We may attend more to one member of the couple than to the other. . . . We are more conscious of heat when passing to a higher temperature, and of cold when passing to a lower. The state we have passed to is our explicit consciousness, the state we have passed from is our implicit consciousness.” But the transition need not be from heat to cold, or vice versa: it can equally well take place from a neutral state, which is indeed the normal state, of neither heat nor cold; a new-born mammal, e.g. must experience cold, having never experienced heat. Again, suppose a sailor becalmed gazing for a whole morning upon a stretch of sea and sky, what sensations are implicit here? Shall we say yellow as the greatest contrast to blue, or darkness as the contrary of light, or both? What, again, is the implicit consciousness when the explicit is sweet; is it bitter or sour, and from what is the transition in such a case? For one thing it seems clear that the transition of attention from one presentation to another and the differences between the presentations themselves are distinct facts. It is strange that the psychologist who has laid such stress on neutral states of surprise as being akin to feeling and so distinct from special presentations, should in any way confound the two. The mistake is perhaps accounted for by the fact that Bain, in common with the rest of his school, nowhere distinguishes between attention and the presentations that are attended to. If “change of impression” and being conscious or mentally alive are the same thing, it is then manifestly tautologous to say that one is the indispensable condition of the other. If they are not the same thing, then the succession of shocks or surprises cannot wholly determine the impressions which successively determine them.

But we have still to consider whether the impressions themselves are nothing but differences or contrasts. “We do not know any one thing of itself but only the difference between it and another thing,” said Bain. But it is plain we cannot speak of contrast or difference between two states or things as a contrast or difference, if the states or things are not themselves presented; the so-called contrast or difference would then be itself a single presentation, and its supposed “relativity” but an inference. Difference is not more necessary to the presentation of two objects than two objects to the presentation of difference. And, what is more, a difference between presentation is not at all the same thing as the presentation of that difference. The former must precede the latter; the latter, which requires active comparison, need not follow. There is an ambiguity in the words “know,” “knowledge,” which Bain seems not to have considered: “to know” may mean either to perceive or apprehend, or it may mean to understand or comprehend.[20] Knowledge in the first sense is only what we shall have presently to discuss as the recognition or assimilation of an impression (see below, § 18); knowledge in the latter sense is the result of intellectual comparison and is embodied in a proposition. Thus a blind man who cannot know light in the first sense can know about light in the second if he studies a treatise on optics. Now in simple perception or recognition we cannot with any exactness say that two things are perceived: straight is a thing, i.e. a definite object presented; not so not-straight, which answers to no definite object at all. Only when we rise to intellectual knowledge is it true to say: “No one could understand the meaning of a straight line without being shown a line not straight, a bent or crooked line.”[21] Two distinct presentations are necessary to the comparison that is here implied; but we must first recognize our objects before we can compare them, and this further step we may never take. We need, then, to distinguish between the comparativity of intellectual knowledge, which we must admit—for it rests at bottom on a purely analytical proposition—and the “differential theory of presentations,” which, however plausible at first sight, must be wrong somewhere, since it commits us to absurdities. Thus, if we cannot have a presentation X but only the presentation of the difference between Y and Z, it would seem that in like manner we cannot have the presentation of Y or Z, nor therefore of their difference X, till we have had the presentation of A and B say, which differ by Y, and of C and D, which we may suppose differ by Z.

The lurking error in this doctrine, that all presentations are but differences, may perhaps emerge if we examine more closely what may be meant by difference. We may speak of (a) differences in intensity between sensations supposed to be qualitatively identical, as e.g. between the taste of strong and weak tea; or of (b) differences in quality between presentations of the same sense, as e.g. between red and green; or of (c) differences between presentations of distinct senses, as e.g. between blue and bitter. Now as regards (a) and (b), it will be found that the difference between two intensities of the same quality, or between two qualities of the same order, may be itself a distinct presentation, that is to say, in passing from a load of 10 ℔ to one of 20 ℔, for example, or from the sound of a note to that of its octave, it is possible to experience the change continuously, and to estimate it as one might the distance between two places on the same road. But nothing of this kind holds of (c).[22] In passing from the scent of a rose to the sound of a gong or a sting from a bee we have no such means of bringing the two into relation—scarcely more than we might have of measuring the length of a journey made partly on the common earth and partly through the looking-glass. In (c), then, we have only a diversity of presentations, but not a special presentation of difference; and we only have more than this in (a) or (b) provided the selected presentations occur together. We say that we know the difference between a sound and a taste; but what we mean is simply that we know what it is to pass from attending to the one to attending to the other. It is simply an experience of change. Change, however, implies continuity, and there is continuity here in the movement of attention and the affective state consequent on that, but not directly in the qualities themselves.

c. If red follows green we may be aware of a greater difference than we could if red followed orange; and we should ordinarily call a 10 ℔ load heavy after one of 5 ℔ and light after one of 20 ℔. Facts like these it is which make the differential theory of presentations plausible. On the strength of such facts Wundt has formulated a law of relativity, free, apparently, from the objections just urged against Bain’s doctrine. It runs thus: “Our sensations afford no absolute but only a relative measure of external impressions. The intensities of stimuli, the pitch of tones, the qualities of light, we apprehend (empfinden) in general only according to their mutual relation, not according to any unalterably fixed unit given along with or before the impression itself.”[23]

But if true this law would make it quite immaterial what the impressions themselves were: provided the relation continued the same, the sensation would be the same too, just as the ratio of 2 to 1 is the same whether our unit be miles or millimetres. In the case of intensities, e.g. there is a minimum sensible and a maximum sensible. The existence of such extremes is alone sufficient to turn the flank of the thoroughgoing relativists; but there are instances enough of intermediate intensities that are directly recognized. A letter-sorter, for example, who identifies an ounce or two ounces with remarkable exactness identifies each for itself and not the first as half the second; of an ounce and a half or of three ounces he may have a comparatively vague idea. And so generally within certain limits of error, indirectly ascertained, we can identify intensities, each for itself, neither referring to a common standard nor to one that varies from time to time—to any intensity, that is to say, that chances to be simultaneously presented; just as an enlisting sergeant will recognize a man fit for the Guards without a yard measure and whether the man’s comrades are tall or short. As regards the qualities of sensations the outlook of the relativists is, if anything, worse. In what is called Meyer’s experiment (described under Vision) what appears greenish on a red ground will appear of an orange tint on a ground of blue; but this contrast is only possible within certain very narrow limits. In fact, the phenomena of colour-contrast, so far from proving, distinctly disprove that we apprehend the qualities of light only according to their mutual relation. In the case of tones it is very questionable whether such contrasts exist at all. Summing up on the particular doctrine of relativity of which Wundt is the most distinguished adherent, the truth seems to be that, in some cases where two presentations whose difference is itself presentable occur in close connexion, this difference—as we indirectly learn—exerts a certain bias on the assimilation or identification of one or both of the presentations. There is no “unalterably fixed unit” certainly, but, on the other hand, “the mutual relations of impressions” are not everything.

15. The term “field of consciousness” has occurred sundry times in the course of this exposition: it is one of several Subconsciousness. employed in describing what have been incidentally referred to as “degrees or grades of consciousness”—a difficult and perplexing topic that we must now endeavour further to elucidate. Sailors steering by night are said to look at the pole-star, “the cynosure of every eye,” but this does not prevent them from seeing the rest of the starry vault. At a conversazione we may listen to some one speaker while still hearing the murmur of other voices, and while listening we may also see the speaker and thereby identify him the better. What in these instances is looked at or listened to has been called the “focus” of consciousness, the rest of what is heard or seen or otherwise presented being called the “field” within which attention is thus concentrated or brought to a point. Of these objects beyond the focus we have then only a lower degree of consciousness, and the more “distant” they are from the centre of interest the fainter and obscurer they are supposed to be or to become. Now, it is obvious that the continuity here implied, if strictly taken, logically commits us to a field of consciousness extending with ever diminishing intensity ad indefinitum. But we have next to notice certain new features that have led psychologists to give to the term field of consciousness a more restricted meaning. A meteor flashing across the sky would certainly divert the helmsman's attention, and for the nonce he would look at that and not at the star in the Little Bear's tail; a voice at our elbow accosting us, we should turn to the new speaker and listen to him, still hearing it may be, but no longer “following,” the discourse thus for us interrupted. In these cases a change in the field of consciousness brings about a non-voluntary change in the focus. But it only does so provided it is sufficiently intense and abrupt, and the more attention is already concentrated the less effective a given disturbance will be. A whole swarm of meteors might have streaked the sky unheeded while Ulysses, life in hand, steered between Scylla and Charybdis, just as all the din of the siege failed to distract Archimedes bent over his figures in the sand. On the other hand, we can voluntarily transfer the focus of consciousness to any object within the field, provided again this is sufficiently differentiated from the rest. But, more than that, we can not only of our own motion turn to lock at or to listen to what we have only seen or heard, but not noticed before; we can also look out or listen for something not as yet distinguishable, perhaps not as yet existing at all. And here again the concentration of attention may be maximal, as when a shipwrecked crew scan the horizon for a sail, or a beleaguered troop hearken for the oncoming of rescue. Now, such anticipated presentations as soon as they are clearly discernible have already a certain finite intensity, and so they are said to have passed over “the threshold”—to use Herbart's now classic phrase—and to have entered the field of consciousness. Afterwards any further increase in their intensity is certainly gradual; are we then to suppose that before this their intensity changed instantly from zero to a finite quantity and not rather that there was an ultraliminal or subliminal phrase where too it only changed continuously? The latter alternative constitutes the hypothesis of subconsciousness.

According to this hypothesis the total field with which we began is divided into two parts by what Fechner emphatically called “the fact of the threshold,” and the term field of consciousness is henceforth restricted to that part within which the focus of consciousness always lies, the outlying part being the region of subconsciousness. Difficulties now begin to be apparent. The intensity or vivacity of a presentation within the field of consciousness depends partly on what we may call its inherent or absolute intensity, partly on the attention that it receives; but this does not hold of presentations in subconsciousness. These sub-presentations, as we ought perhaps to call them, cannot be severally and selectively attended to, cannot be singled out as direct objects of experience. Many psychologists have accordingly maintained not only that they cannot with propriety be called presentations, but that they have no strictly psychical existence at all. This, however, is too extreme a view. If nothing of a presentational character can exist save in the field of consciousness as thus circumscribed by a definite boundary or threshold, a breach of continuity is implied such as we nowhere else experience: even the field of sight, from which the metaphor of a field of consciousness is derived has no such definite margin. The threshold then is not comparable to a mathematical line on opposite sides of which there is an intensive discontinuity. This has been amply proved by the psychophysical investigations of Fechner and others. We listen, say, to a certain sound as it steadily diminishes; at length we cease to hear it. Again, we listen for this same sound as it steadily increases and presently just barely hear it. In general it is found that its intensity in the former case is less than it is in the latter, and there is also in both cases a certain margin of doubt between clear presence and clear absence; the presentation seems to flicker in and out, now there and now gone. Further, in comparing differences in sensations—of weight, brightness, temperature, &c.—we may fail wholly to detect the difference between a and b, b and c, and yet the difference between a and c may be clearly perceived. We have thus to recognize the existence of a difference between sensations, although there is no so-called “sensation of difference.” But if this much continuity must be admitted we can hardly fail to admit more. If differences of presentation exist within the field of consciousness beyond the outermost verge of the “threshold of difference,” we cannot consistently deny the existence of any presentations at all beyond the threshold of consciousness. Since the field of consciousness varies greatly and often suddenly with the amount and distribution of attention, we must, as already said, either recognize such subconscious presentations or suppose that clearly differentiated presentations, presentations that is to say of finite intensity, pass abruptly into or out of existence with every such variation of the field.

The hypothesis of subconsciousness, then, is in the main nothing more than the application to the facts of presentation of the law of continuity, its introduction into psychology being due to Leibnitz, who first formulated that law. Half the difficulties in the way of its acceptance are due to our faulty terminology. With Leibnitz consciousness was not coextensive with all psychical life, but only with certain higher phases of it.[24] Of late, however, the tendency has been to make consciousness cover all stages of mental development, and all grades of presentation, so that a presentation of which there is no consciousness resolves itself into the manifest contradiction of an unpresented presentation—a contradiction not involved in Leibnitz's “unapperceived perception.” But such is not the meaning intended when it is said, for example, that a soldier in battle is often unconscious of his wounds or a scholar unconscious at any one time of most of the knowledge “hidden in the obscure recesses of his mind.” There would be no point in saying a subject is not conscious of what is not presented at all; but to say that what is presented lacks the intensity requisite in the given distribution of attention to change that distribution appreciably is pertinent enough. Subconscious presentations may tell on conscious life—as sunshine or mist tells on a landscape, or the underlying writing on a palimpsest—although lacking either the intensity or the individual distinctness requisite to make them definite features. Even if there were no facts to warrant this concept of an ultra-liminal presentation of impressions it might still claim an a priori justification.

The subconscious presentation of ideas as distinct from impressions calls, however, for some special consideration. As we Subconscious Ideas. can turn our attention to the sensory threshold and await the entrance of an expected impression, so we may await the emergence of a “memory-image”; and again the threshold turns out to be not a mathematically exact boundary but a region of varying depth.[25] What we are trying to recollect seems first to waver, now at the tip of our tongue and the next moment completely gone, then perhaps a moment afterwards rising into clear consciousness. Sometimes when asked, say, for the name of a certain college contemporary we reply: I cannot tell, but I should know the name if I heard it. We are aware that we could “recognize,” though we cannot “reproduce.” At other times we are confident that even recognition is no longer possible, and still if we met the man himself in the old scenes and heard his voice his name might yet recur. Nevertheless, it may be urged, it is surely incredible that all the incidents of a long lifetime and all the items of knowledge of a well-stored mind that may possibly recur—“the infinitely greater part of our spiritual treasures,” as Hamilton says—are severally retained and continuously presented in the form and order in which they were originally experienced or acquired. This, however, is not implied. Images in contrast to impressions have always a certain generality. The same image may figure in very various connexions, as may the same letter, for example, in many words, the same word in many sentences. We cannot measure the literature of language by its vocabulary, nor may we equate the extent of our “spiritual treasures” when these are successively unfolded with the psychical apparatus, so to say, in which they are subconsciously involved.[26] Take the first book of the Aeneid, which, as Macaulay would say, every schoolboy knows: as subconsciously involved, when the boy is not thinking of it, his knowledge is more comparable to a concordance than to the text itself, which nevertheless can be reproduced from it. In the text Aeneas occurs many times, in the concordance as a heading but once. But give him the cue Aeneas scopulum, and the boy reels off from the 180th line; or Praecipue pius Aeneas, and he starts with the 220th. But ask him for the 580th line; he is probably helpless, while a dunce with the book in his hand can read it off at once. Say instead Et pater Aeneas, and the boy can straightway complete the line while the dunce is now helpless. So though its explicit revival is successional, occurs, so to say, in single file, a whole scheme in which many ideas are involved may rise towards the threshold together. When our schoolboy, for example, turns from classics to geography, the mention of Atlas, which might then have recalled a Titan, now leads him to think only of his book of maps. And there is a like sudden shifting of the substratum of our thoughts, when, taking up the morning paper, we glance first at the foreign telegrams, then at the money market, and then at the doings of our political friends. Yet more remote than all, obscurer but more pervasive, like the clouds of cherubs or imps vaguely limned in medieval pictures, are the indefinite constituents of our emotional atmosphere, “gay motes that people the sunbeams” of our cheerfulness and make all couleur de rose, or “horrid shapes and sights unholy” that overcast the outlook when we “have the blues.” And as attention relaxes, these advance into the foreground and become more or less palpable hopes or fears.

Because of the manifold forms into which they may evolve, subconscious images, while still involved, are sometimes called “psychical” or more definitely “presentational dispositions.” The word disposition means primarily an arrangement, as when we talk of the disposition of troops in a battle or of cards in a game; the disposita, that is to say, are always something actual. Which of several potential dispositions they will actually assume will depend upon circumstances, but at least, as Leibnitz long ago maintained, “les puissances véritables ne sont jamais des simples possibilités.” What is requisite to the realization of a given potentiality is sometimes a condition to be added, sometimes it is one to be taken away. A locomotive with the fire out has no tendency to move, but with steam up it is only hindered from moving by the closure of the throttle-valve or the friction of the brake. Now presentational dispositions we assume to be of the latter sort. They are processes or functions more or less inhibited, and the inhibition is determined by their relation to other psychical processes or functions. The analysis and genesis of these presentational interactions will occupy us at length by and by; it may then be possible to explain the gradual involution of what was successively unfolded in explicit consciousness into those combinations which Herbart called “apperception-masses,” combinations devoid of the concrete hints of date and place which are essential to memory. Meanwhile the evidence adduced—decidedly cogent though admittedly indirect—together with the difficulties besetting the extreme view that beyond or below the threshold of consciousness there is nothing presentational, seems clearly to justify the hypothesis of subconsciousness. At the same time the principle of continuity, everywhere of fundamental importance when we are dealing with reality, forbids the attempt arbitrarily to assign any limits to the subconscious.

Many psychologists have proposed to explain subconscious retention by habit. But it is obvious that habit itself implies retention and is practically synonymous with disposition; it must therefore presuppose disposita if we are to escape the absurdities of puissances ou facultés nues, with which in this very connexion Leibnitz twitted Locke. Yet, obvious as all this may be, it is frequently ignored even by those who are fond of exposing the pretended explanations of the “faculty-psychologists” and quoting Molière to confute them. Thus we find J. S. Mill arguing: “I have the power to walk across the room though I am sitting in my chair; but we should hardly call this power a latent act of walking.”[27] Nor should we call it a power at all if Mill had been paralysed, or if, instead of sitting in his chair, he had been lying in his cradle. What we want is the simplest psychological description of the situation after the power has been acquired by practice and is still retained. In such a case we can be conscious of the “idea” of the movement without the movement actually ensuing; yet only in such wise that the idea is more apt to pass over into action the intenser it is, and often actually passes over in spite of us. Surely there must be some functional activity answering to this conscious presentation; why may not a much less amount of it be conceived possible in subconscious presentation?

Sensation, Movement and the External World.

16. On the view of experience here maintained, we are bound to challenge the description of sensations[28] as due to Definition of Sensation. physical stimuli—widely current though it is—as one that is psychologically inappropriate. The following definition, given by Bain, may be taken as a type: “By sensations, in the strict meaning, we understand the mental impressions, feelings or states of consciousness following on the action of external things on some part of the body, called on that account sensitive.”[29] It is true, no doubt, that what the psychologist calls sensibility has as its invariable concomitant what physiologists call sensibility, or what the more careful of them call irritability; and, true again, that this irritability is invariably preceded by a physical process called stimulation. But it may be urged, why not recognize a connexion that actually obtains, since otherwise sensation must remain unexplained? Well, in the first place, such “psychophysical” connexion is not a psychological explanation: it cannot be turned directly to account in psychology, either analytic or genetic. Next the psychological fact called sensation always is, and at bottom always must be, independently ascertained; for the physiological “neurosis” or irritation has not necessarily a concomitant “psychosis” or sensation and, strictly dealt with, affords no hint of such. Finally, this inexplicability of sensation is a psychological fact of the utmost moment: it answers to what we call reality in the primary sense of the term. The psychophysicist, in setting out to explain sensation, has—unawares to himself—left this fundamental reality behind him. For it belongs essentially to individual experience, and this—in assuming the physical standpoint—he has of course transcended. Nevertheless the mistake of method that here reveals itself was perhaps inevitable, for the facts of another's sense-organs and their physical excitants must have obtruded themselves on observation long before the reflective attitude was advanced enough to make strictly psychological analysis possible. The psycho physical standpoint, that is to say, was attained before the purely psychological; and the consequent bias is only now in process of correction. A series of physical processes, first without and then within the organism—ethereal or aerial vibrations, neural and cerebral excitations—was the starting point. What comes first, immediately, and alone, in the individual's experience, and is there simply and positively real, was then misinterpreted as subjective modification, mental impression, species sensibiles, or the like. For from the days of Democritus to our own the same crude metaphor has prevailed without essential variation. And here the saying holds: Vestigia nulla retrorsum. Into the man's head the whole world goes, including the head itself. Such thoroughgoing “introjection” affords no ground for subsequent “projection.” Thus the endeavour to explain sensation overreaches itself: the external object or thing that was supposed to cause sensations and to be therefore distinct from them, was in the end wholly resolved into these and regarded as built out of them by association (Mill) or by apperceptive synthesis (Kant). But no “mental chemistry,” no initial alchemy of “forms,” can generate objective reality from feelings or sense-impressions as psychophysically defined.[30] A's experience as it is for B is not real but inferential; and if the grounds of the inference, which are the only realities for B, are to be regarded as the causes of which A's experiences are merely the effects, then the two experiences are on a wholly different footing. When A treats B in the same fashion we get the world in duplicate: (1) as original and outside, i.e. as cause, and (2) as copied within each percipient's head, i.e. as effect. But when B interprets his own experience as he had interpreted A's we seem to have lost the real world altogether. In presence of this dilemma, the philosophers of our time, as already said, are feeling it needful to revise their psychology. The question of method is vital. If the psychophysical standpoint were the more fundamental, psychology would be based on physiology, and the old definition of sensation might stand. If, on the other hand, it is the exclusive business of psychology to analyse and trace the development of individual experience as it is for the experiencing individual, then—however much neurological evidence may be employed as a means of ascertaining psychological facts—the facts themselves must be scrupulously divested of all physical implications, the psychophysical method takes a secondary place, and the objective reality of “sensory” presentations stands unimpeached.

The duality of subject and object in experience compels us also to object to the description of sensations as “states of consciousness.” Since it is the subject, not the object that is conscious, the term state of consciousness implies strictly a subjective reference; and so it is only applicable to sensations, if they are regarded as subjective modifications, either affective or active. The former would identify sensation with feeling, and this—for reasons already given—we must disallow. But it is true that a sensation, like other presentations, implies the subjective activity we call attention; it is not, however, a modification or state of this activity, but the object of it. This relation is expressed in German by means of the distinction generally of Vorstellen and Vorstellung and in the present case of Empfinden and Empfindung; and German psychology has gained in clearness in consequence. The distinction of conception and concept (conceit) is to be found in older English writers and was revived by Sir W. Hamilton, who suggested also the analogous distinction of perception and percept. It would be a great gain if there were a corresponding pair of terms to distinguish between “the sensing act” and the object “sensed,” as some have been driven to say. Reception and recept at once occur and seem unexceptionable—apart, of course, from their novelty.[31] At any rate, if we are to rest content with our present untechnical terminology we must understand sensations to mean objective changes as they first break in upon the experience of our psychological individual; in this respect Locke's term “impression” has a certain appropriateness.

What we ordinarily call a single sensation has not only a characteristic quality but it is also quantitatively determined in respect of intensity, pro tensity (or duration) and extensity. A plurality of properties, it may be said, straightway implies Characteristics of Sensation. complexity of some sort. This is obvious and undeniable; psychological—as distinct from psychical[32]—analysis of simple sensations is possible, and the description just given is reached by means of it. Such analysis, however, presupposes the comparison of many sensations; but to the complexity it discloses there is no answering plurality discernible in the immediate experience of a single sensation. To make this clearer let us start from a case in which such plurality can be directly verified. In a handful of rose petals we are aware at once of a definite colour, a definite odour and a definite “feel.” Here there is a plurality (a+b+c), any part of which can be withdrawn from our immediate experience without prejudice to the rest, for we can close the eyes, hold the nose, or drop the petals on the table. Let us now turn to the colour alone; this we say has a certain quality, intensity, extensity, &c. But not only have we not one sense for quality, another for intensity, &c., but we cannot reduce the intensity to zero and yet have the quality remaining; nor can we suppress the quality and still retain the extensity. In this case then what we have is not a plurality of presentations (a+b+c), but a single presentation having a plurality of attributes (a b c) so related that the absence of any one annihilates the whole. But though, as already said, such single presentation gives, as it stands, no evidence of this plurality, yet it is to be remembered that in actual experience we do not deal with sensations in isolation; here, accordingly, we find evidence in plenty to justify our psychological analysis. In innumerable cases we experience varieties of intensity with little or no apparent change of quality, as happens, for example, when a sounding pitch-pipe is moved towards or away from the ear; and continuous changes of quality without any change of intensity, as happens when the pipe is shortened or lengthened without any alteration of position. We may have tactual or visual sensations which vary greatly in extensity without any striking change of quality, and we may have such sensations in every possible variety of quality without any changes of extensity.

The numerous and striking diversities among our present sensations are obviously not primordial; what account then can we give of their gradual differentiation? Some psychologists have assumed the existence of absolute “units of sensibility,” all identically the same, and explain the unlikenesses in our existing sensations as resulting “from unlike Differentiation of Sensation. modes of integration of these absolute units.”[33] The sole evidence on which they rely is physiological, the supposed existence of a single nerve shock or neural tremor. It is true that in an extirpated nerve what is known as the “negative variation” is approximately such an isolated event of uniform quality. But the same cannot be said of what happens during the stimulation of a nerve in situ with its peripheral and central connexions still intact. The only evidence apparently to which we can safely appeal in this inquiry is that furnished by biology. Protoplasm, the so-called “physical basis of life,” is amenable to stimulation by every form of physical agency—mechanical, chemical, thermal, photical, electrical—with the single exception of magnetism; and in keeping with this it is found that unicellular organisms respond, and respond in ways more or less peculiar, to each of these possible modes of excitation. Since, so far as is known, there is no morphological separation of function in these lowest forms of life, it is reasonably assumed that the single cell acts the part of “universal sense-organ,” and that the advance to such complete differentiation of sense-organs as we find among the higher vertebrates has been a gradual advance. Numerous facts can now be adduced of the occurrence of “transitional” or “alternating” sense-organs among the lower forms of multicellular animals; organs, that is to say, which are normally responsive to two or more kinds of stimulus, and thus hold an intermediate position between the universal sense-organ of the Protozoa and the special sense-organ of the Mammalia. For example, a group of cells which would behave towards all stimuli impartially were they independent unicellular organisms become, as an organ in a multicellular organism, amenable only to mechanical or only to chemical stimuli,—become, that is to say, an organ of touch and of hearing, or an organ of taste and also of smell; until, finally, when differentiation is sufficiently advanced, the group ends by becoming exclusively the organ of one specified sense, touch or hearing in the one case, taste or smell in the other.[34] Of course the imperfectly specialized sensations, say of the leech, and still more the wholly unspecialized sensations of the amoeba, cannot be regarded as blends of some or all of those which we are said to receive through our five senses. We must rather suppose that sensations at the outset corresponded very closely with the general vital action of stimuli as distinct from their action on specially differentiated sensory apparatus. Even now we are still aware of the general effects of light, heat, fresh air, food, &c., as invigorating or depressing quite apart from their specific qualities. Hence the frequent use of the term general or common sensibility (coenesthesis). But, though less definitely discriminated, the earlier, and what we call the lower, sensations are not any less concrete than the later and higher. They have been called general rather than specific, not because psychologically they lack any essential characteristic of sensation which those acquired later possess, but simply because physiologically they are not, like these, correlated to special sense-organs.

But, short of resolving such sensations into combinations of one primordial modification of consciousness, if we could Complexity of Sensations. conceive such, there are many interesting facts which point clearly to a complexity that we can seldom directly detect. Several of our supposed sensations of taste, e.g., are complicated with sensations of touch and smell: thus the pungency of pepper and the dryness of wine are tactual sensations, and their spicy flavours are really smells. How largely smells mingle with what we ordinarily take to be simply tastes is best brought home to us by a severe cold in the head, as this temporarily prevents the access of exhalations to the olfactory surfaces. The difference between the smooth feel of a polished surface and the roughness of one that is unpolished, though to direct introspection an irresolvable difference of quality, is probably due to the fact that several nerve terminations are excited in each case: where the sensation is one of smoothness all are stimulated equally; where it is one of roughness the ridges compress the nerve-ends more, and the hollows compress them less, than the level parts do. The most striking instance in point, however, is furnished by the differences in musical sounds, to which the name timbre is given. To the inattentive or uninstructed ear notes or “compound tones” appear to be only qualitatively diverse and not to be complexes of simple tones. Yet it is possible with attention and practice to distinguish these partial tones in a note produced on one instrument, a horn, say, and to recognize that they are different from those of the same note produced on a different instrument, for example, a violin.

In like manner many persons believe that they can discriminate in certain colours, hence called “mixed,” the elementary colours of which they are held to be composed; red and yellow, for example, in orange, or blue and red in violet. But in so thinking they appear to be misled, partly by the resemblance that certainly exists between orange and red, on the one hand, and orange and yellow on the other, the two colours between which in the colour spectrum it invariably stands; and partly by the knowledge that, as a pigment, orange is obtainable by the mixture of red and yellow pigments; and so in the other cases. As we shall see later, however (§ 39), in this particular case of sensory continua, resemblance is no proof of complexity. Were it otherwise we should have to conclude that a given tone, since this also resembles the two between which it is intermediate, ought to be a blend of both; whereas, in point of fact, the tone d—though as regards pitch it has a certain resemblance to c and e, its neighbours on either side—differs widely from the chord c-e, which is made up of these. In all cases in which the psychical complexity of a sensation is beyond dispute the partial sensations are distinguished by discernible differences of extensity, and usually of intensity as well. Thus, if the skin be touched by the point of a hot or cold bradawl the temperature sensation has not the punctual character of the touch but seems rather to surround this as a sort of penumbra. Similarly, the ground-tone of a clang-complex has not only a greater intensity but also a greater extensity than any of the over-tones.[35] There is also in such cases a certain rivalry or antagonism between the complex as an unanalysed whole and the complex as analysed, and even between the several partial sensations after such analysis. In the absence of such direct evidence it is unwarrantable to infer psychical complexity from complexity in the physical stimuli, even when this is really present. In the case of pigment mixture, however, there is no such physical complexity as is vulgarly supposed. And it is worth noting that white light is physically the most complex of all, whereas the answering sensation is not only simple but probably the most primitive of all visual sensations.

Every sensation within the fields of consciousness has sensibly some continuous duration and seems sensibly to admit Quantitative Continuity. of some continuous variation in intensity and extensity. But whether this quantitative continuity of presentational change is more than apparent has been questioned. Sensations of almost liminal intensity are found to fluctuate every few seconds, and, as already remarked, when the threshold of intensity is actually reached, they seem intermittently to appear and disappear, a fact which Hume long ago did not fail to notice. The results of numerous experiments, however, justify the conclusion that these variations are due primarily to oscillation of attention, and furnish so far no ground for the assumption that even the liminal sensation is discontinuous. But again we can only detect a difference of intensity when this is of finite amount and bears a certain constant ratio to the initial intensity with which it is compared—a fact commonly known as Weber's Law. But this imperfection in our power of discrimination is no proof that our sensations vary discontinuously; and not only is there no positive evidence in favour of such discontinuity, but it is altogether improbable on general grounds. Lastly, there is always more or less distinctness in the several nerve-endings as well as isolation of the nerve-fibres themselves. The skin, for example, when carefully explored, turns out to be a complex mosaic of so-called “spots,” severally responding to stimulation by sensations of pressure, heat, cold and pain. But from this to argue that the extensity of a sensation is really a mere aggregate without any continuity is on a par with calling a lake a collection of pools because it is fed by separate streams. If it could be shown that in the brain as a whole there is no functional continuity a formidable psychophysical problem would no doubt arise.

As regards the quality of sensations—the primitive sensation of sight appears to consist only of the single quality we call Quality of Sensations. “light,” a quality which ranges in intensity from a dazzling brightness that becomes painful and blinding down to a zero of complete darkness; a limit which, however, is never completely attained, since the retina is always more or less internally stimulated—hence what is called the eye's own light (Eigenlicht). The first responses to light-stimulation seem to be very much on a par with our own to diffused heat or cold; some organisms seek the light and others shun it. As little as our temperature-sense yields us a perception of form does the light-sense at this level yield any. Not until the stage of visual spatial perception is reached and some discrimination of form is possible, do black and white attain the meaning they now have for us. An object can be visually perceived only when its colour or shade differs from that of the surrounding field; so far black as a “secondary quality” is on a par with colour, that is to say, when we are talking of things it may be called a quality. But there is still an important difference; in a light field many colours or shades may be distinguished, but in a dark field none. Though it is correct to speak of perceiving a black object, must we not then maintain that—so far as it is really black[36]—the object yields us directly no sensation? Similarly, the piper is said to “feel” the holes in his whistle when actually he only touches the solid metal in which they are pierced; or the soldier is said to hear the tattoo, though he has no auditory sensation of the silence intervening between successive taps on the drum. And it has yet to be shown that there is any more justification for speaking of visual sensations without luminosity. Meanwhile we must maintain that in absolute darkness we do not see black, since we do not see at all. No doubt we are prone to identify the two concepts darkness and blackness, for what we may call their sensory content is the same, viz. the absence of visual sensation.

Whereas in nature the only diffused light we need consider is that emitted by the sun, the rays transmitted by the things about us vary in physical quality and in their effects upon protoplasm. As soon, therefore, as visual forms can be distinguished, a differentiation among light-sensations becomes obviously advantageous. The first colours to be differentiated were probably yellow and blue, or perhaps it would be truer to say “warm” colour and “cold” colour, upon which there followed a further differentiation of the warm colour into red and green.[37] It is interesting to note that all possible sensations of colour constitute a specific continuum. We may represent it by a sphere, in which (a) the maximum of luminosity is at one pole and the minimum at the other; (b) the series of colours proper (red to violet and through purple back to red), constituting a closed line, are located round the equator or in zones parallel to it, according to shade; and (c) the amount of saturation (or absence of white) for any given zone of illumination increases with distance from the axis.

In dealing with the quality of auditory sensations we have to distinguish between the simple sensations called tones and the sensation-complexes, either clangs or noises, which result from their combination. Simple tones also constitute a qualitative continuum, but it has only one dimension, their so-called “pitch”; this may be represented by a straight line ranging between two more or less indefinite extremes. If intensity, that is to say loudness, is taken into account, we have of course a continuum of two dimensions. The tone-continuum is also universally regarded as steadily diminishing in massiveness or extensity as the pitch rises. And, in fact, as we approach the lower limit, the so-called deep or grave tones become less “even,” till at length distinct, more or less pervasive, tremors are felt rather than heard as distinct impulses on the ear-drum. The so-called high or acute tones again, as we approach their limit, are accompanied by tactual, often more or less painful, sensations, as if the ear were pierced by a fine needle. This connexion of auditory with tactual sensations confirms the independent evidence of biology pointing to an original differentiation of sound from touch. The special characteristics of tone-complexes, whether clangs or noises, are due to the remarkable analytic power which belongs to the sense of hearing. Two colours cannot be simultaneously presented unless they are differently localized, but several tones may form one complex whole within which they, as “partial” tones, are distinguishable, though spatially undifferentiated.

Unlike the higher senses of sight and hearing, the lower senses of touch, taste, smell, &c., do not constitute qualitative continua. Temperatures may indeed be represented as ranging in opposite directions, i.e. through heat or through cold, between a zero of no sensation and the organic sensations due to the destructive action of both extremes, heat and cold alike. But the continuity in this case is intensive rather than qualitative. Tastes fall into the four isolated qualities known as sweet, sour, bitter, saline; but smells hardly admit of classification at all. Sensations of touch and sight have in a pre-eminent degree a certain peculiar continuity which differentiations of extensity entail, and which we shall have presently to consider further under the title of local signs. The various sensations classed together as organic, hunger, thirst, physical pain, &c., are left to the physiologist to describe.

Our motor presentations contrast with the sensory by their want of striking qualitative differences. They are divided Movements. into two groups: (a) motor presentations proper and (b) auxilio-motor of kinaesthetic presentations. The former answer to our “feelings of muscular effort” or “feelings of innervation.” The latter are those presentations due to the straining of tendons, stretching and flexing of the skin, and the like, by which the healthy man knows that his efforts to move are followed by movement, and so knows the position of his body and limbs. It is owing to the absence of these presentations that the anaesthetic patient cannot directly tell whether his efforts are effectual or not, nor in what position his limbs have been placed by movements from without. Thus under normal circumstances motor presentations are always accompanied by auxilio-motor; but in disease and in passive movements they are separated and their distinctness thus made manifest. Originally we may suppose kinaesthetic presentations to have formed one imperfectly differentiated continuum, but now, as with sensations, they have become a collection of special continua, viz. the groups of movements possible to each limb and certain combinations of these movements.

But whereas kinaesthetic presentations were commonly allowed to be purely sensory, the concomitants of centripetal excitations—hence the older name of “muscular or sixth sense” applied to them by Sir Charles Bell, Weber, Sir William Hamilton and others—concerning motor presentations proper, a very different view, first tentatively advanced by the great physiologist Johannes Müller, and adopted by Helmholtz, Wundt, and especially by Bain, long prevailed. It is, however, now generally discredited, if not completely overthrown.[38] According to this view, “the characteristic feeling of exerted force” must be regarded, Bain maintained, “not as arising from an inward transmission . . . but as the concomitant of the outgoing current by which the muscles are stimulated to act” (op. cit. p. 79). The necessity for this assumption has certainly not been established on physiological grounds, nor apparently did Bain rely primarily on these; for at the very outset of his discussion we find him saying “that action is a more intimate and inseparable property of our constitution than any of our sensations, and enters as a component part into every one of our senses” (op. cit. p. 59). But this important psychological truth is affirmed as strenuously by some, at any rate (e.g. Professor James) of Bain's opponents as it was by Bain himself. Unhappily many, under the same psychophysical bias and so induced, like the upholders of this innervation theory, to look for evidence of subjective activity in the wrong place, have been led to doubt or to deny the reality of this activity altogether. In fact, this theory, while it lasted, tended to sustain an undue separation of so-called “sensory” from so-called “motor” presentations, as if living experience were literally an alternation of two independent states, one wholly passive and the other wholly active, corresponding to the anatomical distinction of organs of sense and organs of movement. The subject of experience or Ego does not pass to and fro between a sensorium commune or intelligence department and a motorium commune or executive, is not in successive intervals receptive and active, still less always passive, but rather always actively en rapport with an active Non-Ego, commonly called the External World.

Perception.

17. In treating apart of the differentiation of our sensory and motor continua, as resulting merely in a number of Mental Synthesis or Integration. distinguishable sensations and movements, we have been compelled by the exigencies of exposition to leave out of sight another process which really advances pari passu with this differentiation, viz. the integration or synthesis of these proximately elementary presentations into those complex presentations which are called percepts, intuitions, sensori-motor reactions and the like. It is, of course, not to be supposed that in the evolution of mind any creature attained to such variety of distinct sensations and movements as a human being possesses without making even the first step towards building up this material into the most rudimentary knowledge and action. On the contrary, there is every reason to think, as has been said already incidentally, that further differentiation was helped by previous integration, that perception prepared the way for distincter sensations, and purposive action for more various movements. This process of synthesis, which is in the truest sense a psychical process, deserves some general consideration before we proceed to the several complexes that result from it. Most complexes, certainly the most important, are consequences of that principle of subjective selection whereby interesting sensations lead through the intervention of feeling to movements; and the movements that turn out to subserve such interest come to have a share in it. In this way—which we need not stay to examine more closely now—it happens that a certain sensation, comparatively intense, and a certain movement, definite enough to control that sensation, engage attention, to the more or less complete exclusion of the other less intense sensations and more diffused movements that accompany them. Apart from this intervention of controlling movements, the presentation-continuum, however much differentiated, would remain for all purposes of knowledge little better than the disconnected manifold for which Kant took it. At the same time it is to be remembered that the subject obtains command of particular movements out of all the mass involved in emotional expression only because such movements prove on occurrence adapted to control certain sensations. A long process, in which natural selection probably played the chief part at the outset—subjective selection becoming more prominent as the process advanced-must have been necessary to secure as much purposive movement—as even a worm displays. We must look to subjective interest to explain, so far as psychological explanation is possible, those syntheses of motor and sensory presentations which we call spatial perception and the intuitions of material things. For example, some of the earliest lessons of this kind seem to be acquired as we may presently see, in the process of exploring the body by means of the limbs,—a process for which grounds in subjective interest can obviously never be wanting.

Perception sometimes means only the recognition of a sensation or movement as distinct from its original presentation, Meaning of Perception. thus implying the more or less definite revival of certain residua of past experience which resembled the present. More frequently it is used as the equivalent of what has been otherwise called the “localization and projection” of sensations—that is to say, of sensations apprehended either as affections of some part of our own body regarded as extended or as states of some foreign body beyond it. According to a former usage, strictly taken, there might be perception without any spatial presentation at all; a sensation that had been attended to a few times might be perceived as familiar. According to the latter, an entirely new sensation, provided it were complicated with motor experiences in the way required for its localization or projection, would be perceived. But as a matter of fact actual perception probably invariably includes both cases: impressions which we recognize we also localize or project, and impressions which are localized or projected are never entirely new—they are, at least, perceived as sounds or colours or aches, &c. It will, however, frequently happen that we are specially concerned with only one side of the whole process, as is the case with a tea-taster or a colour-mixer on the one hand, or, on the other, with the patient who is perplexed to decide whether what he sees and hears is “subjective,” or whether it is “real.” But there is still a distinction called for: perception as we now know it involves not only recognition (or assimilation) and localization, or “spatial reference,” as it is not very happily termed, but it usually involves “objective reference” as well. We may perceive sound or light without any presentation of that which sounds or shines; but none the less we do not regard such sound or light as merely the object of our attention, as having only immanent existence, but as the quality or change or state of a thing, an object distinct not only from the subject attending but from all presentations whatever to which it attends. Here again the actual separation is impossible, because this factor in perception has been so intertwined throughout our mental development with the other two. Still a careful psychological analysis will show that such “reification,” as we might almost call it, has depended on special circumstances, which we can at any rate conceive absent. These special circumstances are briefly the constant conjunctions and successions of impressions, for which psychology can give no reason, and the constant movements to which they prompt. Thus we receive together, e.g. those impressions we now recognize as severally the scent, colour, and “feel” of the rose we pluck and handle. We might call each a “percept,” and the whole a “complex percept.” But there is more in such a complex than a sum of partial percepts; there is the apprehension or intuition of the rose as a thing having this scent, colour and texture. We have, then, under perception to consider (a) the recognition and (b) the localization of impressions, and (c) the intuition of things.

18. The range of the terms recognition or assimilation of impressions is wide: between the simplest mental process they Assimilation of Impressions. may be supposed to denote and the most complex there is a great difference. The penguin that watched unmoved the first landing of man upon its lonely rock becomes as wild and wary as more civilized fowl after two or three visits from its molester: it then recognizes that featherless biped. His friends at home also recognize him though altered by years of peril and exposure. In the latter case some trick of voice or manner, some “striking” feature, calls up and sustains a crowd of memories of the traveller in the past-events leading on to the present scene. The two recognitions are widely different, and it is from states of mind more like the latter than the former that psychologists have usually drawn their description of perception. At the outset, they say, we have a primary presentation or impression P, and after sundry repetitions there remains a mass or a series of P residua, ppp₃ . . . ; perception ensues when, sooner or later, Pn “calls up” and associates itself with these representations or ideas. Much of our later perception, and especially when we are at all interested, awakens, no doubt, both distinct memories and distinct expectations; but, since these imply previous perceptions, it is obvious that the earliest form of recognition, or, as we might better call it, assimilation, must be free from such complications, can have nothing in it answering to the overt judgment, Pn is a P. Assimilation involves retentiveness and differentiation, as we have seen, and prepares the way for re-presentation; but in itself there is no confronting the new with the old, no determination of likeness, and no subsequent classification. The pure sensation we may regard as a psychological myth; and the simple image, or such sensation revived, seems equally mythical, as we may see later on. The nth sensation is not like the first: it is a change in a presentation-continuum that has itself been changed by those preceding; and it cannot with any propriety be said to reproduce these past sensations, for they never had the individuality which such reproduction implies. Nor does it associate with images like itself, since where there is association there must first have been distinctness, and what can be associated can also, for some good time at least, be dissociated.

19. To treat of the localization of impressions is really to give an account of the steps by which the psychological individual comes to a knowledge of space. At the outset of such an inquiry it seems desirable first of all to make plain what lies within our purview, Localization of Impressions. and what does not, lest we disturb the peace of those who, confounding philosophy and psychology, are ever eager to fight for or against the a priori character of this element of knowledge. That space is a priori in the epistemological sense it is no concern of the psychologist either to assert or to deny. Psychologically a priori or original in such sense that it has been either actually or potentially an element in all presentation from the very beginning it certainly is not. It will help to make this matter clearer if we distinguish what philosophers frequently confuse, viz. the concrete spatial experiences, constituting actual localization for the individual, and the abstract concept of space, generalized from what is found to be common in such experiences. A gannet's mind “possessed of” a philosopher, if such a conceit may be allowed, would certainly afford its tenant very different spatial experiences from those he might share if he took up his quarters in a mole. So, any one who has revisited in after years a place from which he had been absent since childhood knows how largely a “personal equation,” as it were, enters into his spatial perceptions. Or the same truth may be brought home to him if, walking with a friend more athletic than himself, they come upon a ditch, which both know to be twelve feet wide, but which the one feels he can clear by a jump and the other feels he cannot. In the concrete “up” is much more than a different direction from “along.” The hen-harrier, which cannot soar, is indifferent to a quarry a hundred feet above it—to which the peregrine, built for soaring, would at once give chase—but is on the alert as soon as it descries prey of the same apparent magnitude, but upon the ground. Similarly, in the concrete, the body is the origin or datum to which all positions are referred, and such positions differ not merely quantitatively but qualitatively. Moreover, our various bodily movements and their combinations constitute a network of co-ordinates, qualitatively distinguishable but geometrically, so to put it, both redundant and incomplete. It is a long way from these facts of perception, which the brutes share with us, to that scientific concept of space as having three dimensions and no qualitative differences which we have elaborated by the aid of thought and language, and which reason may see to be the logical presupposition of what in the order of mental development has chronologically preceded it. That the experience of space is not psychologically original seems obvious—quite apart from any successful explanation of its origin from the mere consideration of its complexity. Thus we must have a plurality of objects—A out of B, B beside C, distant from D, and so on; and these relations of externality, juxtaposition, and size or distance imply further specialization; for with a mere plurality of objects we have not straightway spatial differences. Juxtaposition, e.g. is only possible when the related objects form a continuum; but, again, not any continuity is extensive. Now how has this complexity come about?

The first condition of spatial experience seems to lie in what has been noted above (§ 11) as the extensity of sensation. This much we may allow is original; for the longer we reflect the more clearly we see that no combination or association of sensations varying only in intensity and Extensity. quality, not even if motor presentations are added, will account for this space-element in our perceptions. A series of touches a, b, c, d may be combined with a series of movements m1, m2, m3, m4; both series may be reversed; and finally the touches may be presented simultaneously. In this way we can attain the knowledge of the coexistence of objects that have a certain quasi-distance between them, and such experience is an important element in our perception of space; but it is not the whole of it. For, as has been already remarked by critics of the associationist psychology, we have an experience very similar to this in singing and hearing musical notes or the chromatic scale. The most elaborate attempt to get extensity out of succession and coexistence is that of Herbert Spencer. He has done, perhaps, all that can be done, and only to make it the more plain that the entire procedure is a ὔστερον πρότερον. We do not first experience a succession of touches or of retinal excitations by means of movements, and then, when these impressions are simultaneously presented, regard them as extensive, because they are associated with or symbolize the original series of movements; but, before and apart from movement altogether, we experience that massiveness or extensity of impressions in which movements enable us to find positions, and also to measure.[39] But it will be objected, perhaps not without impatience, that this amounts to the monstrous absurdity of making the contents of consciousness extended. The edge of this objection will best be turned by rendering the concept of extensity more precise. Thus, suppose a postage stamp pasted on the back of the hand; we have in consequence a certain sensation. If another be added beside it, the new experience would not be adequately described by merely saying we have a greater quantity of sensation, for intensity involves quantity, and increased intensity is not what is meant. For a sensation of a certain intensity, say a sensation of red, cannot be changed into one having two qualities, red and blue, leaving the intensity unchanged; but with extensity this change is possible. For one of the postage stamps a piece of wet cloth of the same size might be substituted and the massiveness of the compound sensation remain very much the same. Intensity belongs to what may be called graded quantity: it admits of increment or decrement, but is not a sum of parts. Extensity, on the other hand, does imply plurality: we might call it latent or merged plurality or a “ground” of plurality, inasmuch as to say that a single presentation has massiveness is to say that a portion of the presentation-continuum at the moment undifferentiated is capable of differentiation.

Attributing this property of extensity to the presentation-continuum as a whole, we may call the relation of any particular sensation to this larger whole its local sign, and can see that, so long as the extensity of a presentation admits of diminution without the presentation becoming nil such presentation Local Signs. either has or may have two or more local signs—its parts, taken separately, though identical in quality and intensity, having a different relation to the whole. Such difference of relation must be regarded fundamentally as a ground or possibility of distinctness of sign—whether as being the ground or possibility of different complexes or otherwise—rather than as being from the beginning such an overt difference as the term “local sign,” when used by Lotze, is meant to imply.[40] From this point of view we may say that more partial presentations are concerned in the sensation corresponding to two stamps than in that corresponding to one. The fact that these partial presentations, though identical in quality and intensity, on the one hand are not wholly identical, and on the other are presented only as a quantity and not as a plurality, is explained by the distinctness along with the continuity of their local signs. Assuming that to every distinguishable part of the body there corresponds a local sign, we may allow that at any moment only a certain portion of this continuum is definitely within the field of consciousness; but no one will maintain that a part of one hand is ever felt as continuous with part of the other or with part of the face. Local signs have thus an invariable relation to each other: two continuous signs are not one day coincident and the next widely separate.[41] This last fact is only implied in the mere massiveness of a sensation in so far as this admits of differentiation into local signs. We have, then, when the differentiation is accomplished, a plurality of presentations constituting an extensive continuum, presented simultaneously, and having certain fixed and invariable relations to each other. Of such experience the typical case is that of passive touch, though the other senses exemplify it. It must be allowed that our concept of space in like manner involves a fixed continuity of positions; but then it involves, further, the possibility of movement. Now in the continuum of local signs there is nothing whatever of this; we might call this continuum an implicit plenum. It only becomes the presentation of occupied space after its several local signs are complicated in an orderly way with active touches, when in fact we have experienced the contrast of movements with contact and movements without, i.e. in vacuo. It is quite true that we cannot now think of this plenum except as a space, because we cannot divest ourselves of these motor experiences by which we have explored it. We can, however, form some idea of the difference between the perception of space and this one element in the perception by contrasting massive internal sensations with massive superficial ones, or the general sensation of the body as “an animated organism” with our perception of it as extended. Or we may express the difference by remarking that extension implies the distinction of here and there, while extensity rather suggests ubiquity.

It must seem strange, if this conception of extensity is essential to a psychological theory of space, that it has escaped notice so long. The reason may be that in investigations into the origin of our knowledge of space it was always the concept of space and not our concrete space percepts that came up for examination. Now in space as we conceive it one position is distinguishable from another solely by its co-ordinates, i.e. by the magnitude and signs of certain lines and angles, as referred to a certain datum, position or origin; and these elements our motor experiences seem fully to explain. But on reflection we ought, surely, to be puzzled by the question, how these coexistent positions could be known before those movements were made which constitute them different positions. The link we thus suspect to be missing is supplied by the more concrete experiences we obtain from our own body, in which two positions have a qualitative difference or “local colour” independently of movement. True, such positions would not be known as spatial without movement; but neither would the movement be known as spatial had those positions no other difference than such as arises from movement. In a balloon drifting steadily in a fog we should have no more experience of change of position than if it hung becalmed and still.

We may now consider the part which movement plays in elaborating the presentations of this dimensionless continuum Positional signs. into percepts of space. In so doing we must bear in mind that while this continuum implies the incopresentability of two impressions having the same local sign, it allows not only of the presentation of sensations of varying massiveness, but also of a sensation involving the whole continuum simultaneously, as in Bain's classic example of the warm bath. As regards the motor element itself, the first point of importance is the incopresentability and invariability of a successive series of auxiliomotor or kinaesthetic presentations, P₁, P₂, P₃, P₄. P₁ cannot be presented along with P₂, and from P₄ it is impossible to reach P₁ again save through P₃ and P₂. Such a series, taken alone, could afford us, it is evident, nothing but the knowledge of an invariable sequence of impressions which it was in our own power to produce. Calling the series of P's “positional signs,” the contrast between them and local signs is obvious. Both are invariable, but succession characterizes the one, simultaneity the other; the one yields potential position without place, the other potential place (τόπος) without position; hence we call them both merely signs.[42] But in the course of the movements necessary to the exploration of the body—probably our earliest lesson in spatial perception—these positional signs receive a new significance from the active and passive touches that accompany them, just as they impart to these last a significance they could never have alone.

It is only in the resulting complex that we have the presentations of actual position and of spatial magnitude. For space, though conceived as a coexistent continuum, excludes the notion of omnipresence or ubiquity; two positions ld and lg must coexist, but they are not strictly distinct positions so long as we conceive ourselves present in the same sense in both. But, if Fd and Fg are, e.g. two impressions produced by compass points touching two different spots as ld and lg on the hand or arm, and we place a finger upon ld and move it to lg, experiencing thereby the series P₁, P₂, P₃, P₄, this series constitutes ld and lg into positions and also invests Fd and Fg with a relation not of mere distinctness as τόποι but of definite distance. The resulting complex perhaps admits of symbolization as follows:—

. . . . . FaFbFcFdFeFfFgFhFk . . . . .

T t t t

Pppp

Here the first line represents a portion of the tactual continuum, Fd and Fg being distinct “feels,” if we may so say, or passive touches presented along with the fainter sensations of the continuum as a whole, which the general “body-sense” involves; T stands for the active touch of the exploring finger and P₁ for the corresponding kinaesthetic sensation regarded as “positional sign”; the rest of the succession, as not actually present at this stage but capable of revival from past explorations, is symbolized by the t t t and ppp₄.

When the series of movements is accompanied by active touches without passive there arises the distinction between one's own body and foreign bodies; when the initial movement of a series is accompanied by both active and passive touches, the final movement by active touches only, and the intermediate movements are unaccompanied by either, we get the further presentation of empty space lying between us and them—but only when by frequent experience of contacts along with those intermediate movements we have come to know all movement as not only succession but change of position. Thus active touches come at length to be projected, passive touches alone being localized in the stricter sense. But in actual fact, of course, the localization of one impression is not perfected before that of another is begun, and we must take care lest our necessarily meagre exposition give rise to the mistaken notion that localizing an impression consists wholly and solely in performing or imaging the particular movements necessary to add active touches to a group of passive impressions. That this cannot suffice is evident merely from the consideration that a single position out of relation to all other positions is a contradiction. Localization, though it depends on many special experiences of the kind described, is not like an artificial product which is completed a part at a time, but is essentially a growth, its several constituents advancing together in definiteness and interconnexion. So far has this development advanced that we do not even imagine the special movements which the localization of an impression implies, that is to say, they are no longer distinctly represented as they would be if we definitely intended to make them: the past experiences are “retained,” but too much blended in the mere perception to be appropriately spoken of as remembered or imaged.

À propos of this almost instinctive character of even our earliest spatial percepts it will be appropriate to animadvert on a misleading implication in the current use of such terms as “localization,” “projection,” “bodily reference,” “spatial reference” and the like. The implication is that external space, or the body as extended, is in some sort presented or supposed apart from the localization, projection or reference of impressions to such space. That it may be possible to put a book in its place on a shelf there must be (1) the book, and (2), distinct and apart from it, the place on the shelf. But in the evolution of our spatial experience impressions and positions are not thus presented apart. We can have, or at least we can suppose, an impression which is recognized without being localized as has been already said; but if it is localized this means that a more complex presentation is formed by the addition of new elements, not that a second distinct object is presented and some indescribable connexion established between the impression and it, still less that the impression is referred to something not strictly presented at all. The truth is that the body as extended is from the psychological point of view not perceived at all apart from localized impressions. In like manner impressions projected (or the absence of impressions projected) constitute all that is perceived as the occupied (or unoccupied) space beyond. It is not till a much later stage, after many varying experiences of different impressions similarly localized or projected, that even the mere materials are present for the formation of such an abstract concept of space as “spatial reference” implies.[43] Psychologists, being themselves at this later stage, are apt to commit the oversight of introducing it into the earlier stage which they have to expound.

20. In a complex percept, such as that of an orange or a piece of wax, may be distinguished the following points concerning Intuition of Things. which psychology may be expected to give an account. (a) the object's reality, (b) its solidity or occupation of space, (c) its unity and complexity, (d) its permanence, or rather its continuity in time and (e) its substantiality and the connexion of its attributes and powers. Though, in fact, these items are most intimately blended, our exposition will be clearer if we consider each for a moment apart.

a. The terms actuality and reality have each more than one meaning. Thus what is real, in the sense of material, is opposed Actuality or Reality. to what is mental; as the existent or actual it is opposed to the non-existent; and again, what is actual is distinguished from what is possible or necessary. But here both terms, with a certain shade of difference, in so far as actual is more appropriate to movements and events, are used, in antithesis to whatever is ideal or represented, for what is sense-given or presented. This seems at least their primary psychological meaning; and it is the one most in vogue in English philosophy at any rate, over-tinged as that is with psychology.[44] Any examination of this characteristic will be best deferred till we come to deal with ideation generally (see § 21 below). Meanwhile it may suffice to remark that reality or actuality is not a single distinct element added to the others which enter into the complex presentation we call a thing, as colour or solidity may be. Neither is it a special relation among these elements, like that of substance and attribute, for example. In these respects the real and the ideal, the actual and the possible, are alike; all the elements or qualities within the complex, and all the relations of those elements to each other, are the same in the rose represented as in the presented rose. The difference turns not upon what these elements are, regarded as qualities or relations presented or represented, but upon whatever it is that distinguishes the presentation from the representation of any given qualities or relations. Now this distinction, as we shall see, depends partly upon the relation of such complex presentation to other presentations in consciousness with it, partly upon its relation as a presentation to the subject whose presentation it is. In this respect we find a difference, not only between the simple qualities, such as cold, hard, red and sweet in strawberry ice, e.g. as presented and as represented, but also, though less conspicuously, in the spatial, and even the temporal, relations which enter into our intuition as distinct from our imagination of it. So then, reality or actuality is not strictly an item by itself, but a characteristic of all the items that follow.

b. In the so-called physical solidity or impenetrability of things our properly motor presentations or “feelings of effort Interpenetrability. or innervation” come specially into play. They are not entirely absent in those movements of exploration by which we attain a knowledge of space; but it is when these movements are definitely resisted, or are only possible by increased effort, that we reach the full meaning of body as that which occupies space. Heat and cold, light and sound, the natural man regards as real, and by and by perhaps as due to the powers of things known or unknown, but not as themselves things. At the outset things are all corporeal like his own body, the first and archetypal thing, that is to say: things are intuited only when touch is accompanied by pressure; and, though at a later stage passive touch without pressure may suffice, this is only because pressures depending on a subjective initiative, i.e. on voluntary muscular exertion, have been previously experienced. It is of more than psychological interest to remark how the primordial factor in materiality is thus due to the projection of a subjectively determined reaction to that action of a not-self of which sense-impressions consist—an action of the not-self which, of course, is not known as such till this projection of the subjective reaction has taken place. Still we must remember that accompanying sense-impressions are a condition of its projection: muscular effort without simultaneous sensations of contact would not yield the distinct presentation of something resistant occupying the space into which we have moved and would move again. Nay more, it is in the highest degree an essential circumstance in this experience that muscular effort, though subjectively initiated, is still only possible when there is contact with something that, as it seems, is making an effort the counterpart of our own. But this something is so far no more than thing-stuff; without the elements next to be considered our psychological individual would fall short of the complete intuition of distinct things.

c. The remaining important factors in the psychological constitution of things might be described in general terms as Unity and Complexity. the time-relations of their components. Such relations are themselves in no way psychologically determined; impressions recur with a certain order or want of order quite independently of the subject's interest or of any psychological principles of synthesis or association whatever. It is essential that impressions should recur, and recur as they have previously occurred, if knowledge is ever to begin; out of a continual chaos of sensation, all matter and no form, such as some philosophers describe, nothing but chaos could result. But a flux of impressions having this real or sense-given order will not suffice; there must be also attention to and retention of the order, and these indispensable processes at least are psychological.

But for its familiarity we should marvel at the fact that out of the variety of impressions simultaneously presented we do not instantly group together all the sounds and all the colours, all the touches and all the smells; but, dividing what is given together, single out a certain sound or smell as belonging together with a certain colour and feel, similarly singled out from the rest, to what we call one thing. We might wonder, too—those at least who have made so much of association by similarity ought to wonder—that, say, the white of snow calls up directly, not other shades of white or other colours, but the expectation of cold or of powdery softness. The first step in this process has been the simultaneous projection into the same occupied space of the several impressions which we thus come to regard as the qualities of the body filling it. Yet such simultaneous and coincident projection would avail but little unless the constituent impressions were again and again repeated in like order so as to prompt anew the same grouping, and unless, further, this constancy in the one group was present along with changes in other groups and in the general field. There is nothing in its first experience to tell the infant that the song of the bird does not inhere in the hawthorn whence the notes proceed, but that the fragrance of the mayflower does. It is only where a group, as a whole, has been found to change its position relatively to other groups, and—apart from casual relations—to be independent of changes of position among them, that such complexes can become distinct unities and yield a world of things. Again, because things are so often a world within themselves, their several parts or members not only having distinguishing qualities but moving and changing with more or less independence of the rest, it comes about that what is from one point of view one thing becomes from another point of view several—like a tree with its separable branches and fruits, for example. Wherein then, more precisely, does the unity of a thing consist? This question, so far as it here admits of answer, carries us over to temporal continuity.

d. Amidst all the change above described there is one thing comparatively fixed: our own body is both constant as a group Temporal Continuity. and a constant item in every field of groups; and not only so, but it is beyond all other things an object of continual and peculiar interest, inasmuch as our earliest pleasures and pains depend solely upon it and what affects it. The body becomes, in fact, the earliest form of self, the first datum for our later conceptions of permanence and individuality. A continuity like that of self is then transferred to other bodies which resemble our own, so far as our direct experience goes, in passing continuously from place to place and undergoing only partial and gradual changes of form and quality. As we have existed—or, more exactly, as the body has been continuously presented—during the interval between two encounters with some other recognized body, so this is regarded as having continuously existed during its absence from us. However permanent we suppose the conscious subject to be, it is hard to see how, without the continuous presentation to it of such a group as the bodily self, we should ever be prompted to resolve the discontinuous presentations of external things into a continuity of existence. It might be said: Since the second presentation of a particular group would, by the mere workings of psychical laws, coalesce with the image of the first, this coalescence would suffice to “generate” the concept of continued existence. But such assimilation is only the ground of an intellectual identification and furnishes no motive, one way or the other, for real identification: between a second presentation of A and the presentation at different times of two A's there is so far no difference. Real identity no more involves exact similarity than exact similarity involves sameness of things; on the contrary, we are wont to find the same thing alter with time, so that exact similarity after an interval, so far from suggesting one thing, is often the surest proof that there are two concerned. Of such real identity, then, it would seem we must have direct experience; and we have it in the continuous presentation of the bodily self; apart from this it could not be “generated” by association among changing presentations. Other bodies being in the first instance personified, that then is regarded as one thing—from whatever point of view we look at it, whether as part of a larger thing or as itself compounded of such parts—which has had one beginning in time. But what is it that has thus a beginning and, continues indefinitely? This leads to our last point.

e. So far we have been concerned only with the combination of sensory and motor presentations into groups and with the Substantiality. differentiation of group from group; the relations to each other of the constituents of each group still for the most part remain. To these relations in the main must be referred the correlative concepts of substance and attribute, the distinction in substances of qualities and powers, of primary qualities and secondary, and the like.[45]

Of all the constituents of things only one is universally present, that above described as physical solidity, which presents itself according to circumstances as impenetrability, resistance or weight. Things differing in temperature, colour, taste and smell agree in resisting compression, in filling space. Because of this quality we regard the wind as a thing, though it has neither shape nor colour, while a shadow, though it has both but not resistance, is the very type of nothingness. This constituent is invariable, while other qualities are either absent or change—form altering, colour disappearing with light, sound and smells intermitting. Many of the other qualities—colour, temperature, sound, smell—increase in intensity if we advance till we touch a body occupying space; with the same movement too its visual magnitude varies. At the moment of contact an unvarying tactual magnitude is ascertained, while the other qualities and the visual magnitude reach a fixed maximum; then first it becomes possible by effort to change or attempt to change the position and form of what we apprehend. This tangible plenum we thenceforth regard as the seat and source of all the qualities we project into it. In other words, that which occupies space is psychologically the substantial; the other real constituents are but its properties or attributes, the marks or manifestations which lead us to expect its presence.

Imagination or Ideation.[46]

21. Before the intuition of things has reached a stage so complete and definite as that just described, imagination or ideation Impressions and Ideas. as distinct from perception has well begun. In passing to the consideration of this higher form of mental life we must endeavour first of all analytically to distinguish the two as precisely as may be and then to trace the gradual development of the higher.

To begin, it is very questionable whether Hume was right in applying Locke's distinction of simple and complex to ideas in the narrower sense as well as to impressions. “That idea of red,” says Hume, “which we form in the dark and that impression which strikes our eyes in the sunshine differ only in degree, not in nature.”[47] But what he seems to have overlooked is that, whereas we may have a mere sensation red, we can only have an image or representation of a red thing or a red form, i.e. of red in some way ideally projected or intuited. In other words, there are no ideas—though there are concepts—answering to simple or isolated impressions. The synthesis which has taken place in the evolution of the percept can only partially fail in the idea, and never so far as to leave us with a chaotic “manifold” of mere sensational remnants. On the contrary, we find that in “constructive imagination” a new kind of effort is often requisite in order partially to dissociate these representational complexes as a preliminary to new combinations. But it is doubtful whether the results of such an analysis are ever the ultimate elements of the percept, that is, merely isolated impressions in a fainter form. We may now try to ascertain further the characteristic marks which distinguish what is imaged from what is perceived.

The most obvious difference is that which Hume called “the force or liveliness” of primary presentations as compared Characteristics of Ideas. with secondary presentations. But what exactly are we to understand by this somewhat figurative language? A simple difference of intensity cannot be all that is meant, for—though we may be momentarily confused—we can perfectly well distinguish the faintest impression from an image; moreover, we can reproduce such faintest impressions in idea. The whole subject of the intensity of representations awaits investigation. Between moonlight and sunlight or between midday and dawn we can discriminate many grades of intensity; but it does not appear that there is any corresponding variation of intensity between them when they are not seen but imagined. Many persons suppose they can imagine a waxing or a waning sound or the gradual abatement of an intense pain; but what really happens in such cases is probably not a rise and fall in the intensity of a single representation, but a change in the complex represented. In the primary presentation there has been a change of quality along with change of intensity, and not only so, but most frequently a change in the muscular adaptations of the sense-organs too, to say nothing of organic sensations accompanying these changes. A representation of some or all of these attendants is perhaps what takes place when variations of intensity are supposed to be reproduced. Again, hallucinations are often described as abnormally intense images which simply, by reason of their intensity, are mistaken for percepts. But such statement, though supported by very high authority, is almost certainly false, and would probably never have been made if physiological and epistemological considerations had been excluded as they ought to have been. Hallucinations, when carefully examined, seem just as much as percepts to contain among their constituents some primary presentation—either a so-called subjective sensation of sight and hearing or some organic sensation due to deranged circulation or secretion. Intensity alone, then, will not suffice to discriminate between impressions and images.

What we may call superior steadiness is perhaps a more constant and not less striking characteristic of percepts. Ideas are not only in a continual fiux, but even when we attempt forcibly to detain one it varies continually in clearness and completeness, reminding one of nothing so much as of the illuminated devices made of gas jets, common at fêtes, when the wind sweeps across them, momentarily obliterating one part and at the same time intensifying another. There is not this perpetual flow and flicker in what we perceive. The impressions entering consciousness at any one moment are psychologically independent of each other; they are equally independent of the impressions and images presented the moment before—independent, i.e. as regards their order and character, not, of course, as regards the share of attention they secure. Attention to be concentrated in one direction must be withdrawn from another, and images may absorb it to the exclusion of impressions as readily as a first impression to the exclusion of a second. But, when attention is secured, a faint impression has a fixity and definiteness lacking in the case of even vivid ideas. One ground for this definiteness and independence lies in the localization or projection which accompanies all perception. But why, if so, it might be asked, do we not confound percept and image when what we imagine is imagined as definitely localized and projected? Because we have a contrary percept to give the image the lie; where this fails, as in dreams, or where, as in hallucination, the image obtains in other ways the fixity characteristic of impressions, such confusion does in fact result. But in normal waking life we have the whole presentation-continuum, as it were, occupied and in operation: we are distinctly conscious of being embodied and having our senses about us.

But how is this contrariety between impression and image possible? With eyes wide open, and while clearly aware of the actual field of sight and its filling, one can recall or imagine a wholly different scene: lying warm in bed one can imagine oneself out walking in the cold. It is useless to say the times are different, that what is perceived is present and what is imaged is past or future.[48] The images, it is true, have certain temporal marks—of which more presently—by which they may be referred to what is past or future; but as imaged they are present, and, as we have just observed, are regarded as actual whenever there are no correcting impressions. We cannot at once see the sky red and blue; how is it we can imagine it the one while perceiving it to be the other? When we attempt to make the field of sight at once red and blue, as in looking through red glass with one eye and through blue glass with the other, either the colours merge and we see a purple sky or we see the sky first of the one colour and then of the other in irregular alternation. That this does not happen between impression and image shows that, whatever their connexion, images as a whole are distinct from the presentation-continuum and cannot with strict propriety be spoken of as revived or reproduced impressions. This difference is manifest in another respect, viz. when we compare the effects of diffusion in the two cases. An increase in the intensity of a sensation of touch entails an increase in the extensity; an increase of muscular innervation entails irradiation to adjacent muscles; but when a particular idea becomes clearer and more distinct, there rises into consciousness an associated idea qualitatively related probably to impressions of quite another class, as when the smell of tar calls up memories of the sea-beach and fishing-boats. Since images are thus distinct from impressions, and yet so far continuous with each other as to form a train in itself unbroken, we should be justified, if it were convenient, in speaking of images as changes in a new continuum; and later on we may see that this is convenient.

Impressions then—unlike ideas—have no associates to whose presence their own is accommodated and on whose intensity their own depends. Each bids independently for attention, so that often a state of distraction ensues, such as the train of ideas left to itself never occasions. The better to hear we listen; the better to see we look; to smell better we dilate the nostrils and sniff; and so with all the special senses: each sensory impression sets up nascent movements for its better reception.[49] In like manner there is also a characteristic adjustment for images which can be distinguished from sensory adjustments almost as readily as these are distinguished from each other. We become most aware of this as, mutatis mutandis, we do of them, when we voluntarily concentrate attention upon particular ideas instead of remaining mere passive spectators, as it were, of the general procession. To this ideational adjustment may be referred most of the strain and “head-splitting” connected with recollecting, reflecting and all that people call headwork; and the “absent look” of one intently thinking or absorbed in reverie seems directly due to the absence of sensory adjustment that accompanies the concentration of attention upon ideas.

22. But, distinct as they are, impressions and images are still closely connected. In the first place, there are two or three Connexion of Impressions and Images. well-marked intermediate stages, so that, though we cannot directly observe it, we seem justified in assuming a steady transition from the one to the other. As the first of such intermediate stages, it is usual to reckon what are often, and—so far as psychology goes—inaccurately, styled after-images. They would be better described as after-sensations, inasmuch as they are due either (1) to the persistence of the original peripheral excitation after the stimulus is withdrawn, or (2) to the effects of the exhaustion or the repair that immediately follows this excitation. In the former case they are qualitatively identical with the original sensation and are called “positive,” in the latter they are complementary to it and are called “negative” (see Vision). These last, then, of which we have clear instances only in connexion with sight, are obviously in no sort re-presentations of the original impression, but a sequent presentation of diametrically opposite quality; while positive after-sensations are, psychologically regarded, nothing but the original sensations in a state of evanescence. It is this continuance and gradual waning after the physical stimulus has completely ceased that give after-sensations their chief title to a place in the transition from impression to image. There is, however, another point: after-sensations are less affected by movement than impressions are. If we turn away our eyes we cease to see the flame at which we have been looking, but the after-image remains still projected before us and continues localized in the dark field of sight, even if we close our eyes altogether. This fact that movements do not suppress them, and the fact that yet we are distinctly aware of our sense-organs being concerned in their presentation, serve to mark off after-sensations as intermediate between primary and secondary presentations. The after-sensation is in reality more elementary than either the preceding percept or its image. In both these, in the case of sight, objects appear in space of three dimensions, i.e. with all the marks of solidity and perspective;[50] but the so-called after-image lacks all these.

Still further removed from normal sensations (i.e. sensations determined by the stimuli appropriate to the sense-organ) are the “recurrent sensations” often unnoticed but probably experienced more or less frequently by everybody—cases, that is, in which sights or sounds, usually such as at the time were engrossing and impressive, suddenly reappear several hours or even days after the physical stimuli, as well as their effects on the terminal sense-organ, seem entirely to have ceased. Thus workers with the microscope often see objects which they have examined during the day stand out.clearly before them in the dark; it was indeed precisely such an experience that led the anatomist Henle first to call attention to these facts. But he and others have wrongly referred them to what he called a “sense-memory”; all that we know is against the supposition that the eye or the ear has any power to retain and reproduce percepts. “Recurrent sensations” have all the marks of percepts which after-sensations lack; they only differ from what are more strictly called “hallucinations” in being independent of all subjective suggestion determined by emotion or mental derangement.

In what Fechner has called the “memory after-image” or the primary memory-image, as it is better termed, we have the image proper in its earliest form. As an instance of what is meant may be cited the familiar experience that a knock at the door, the hour struck on the clock, the face of a friend whom we have passed unnoticed, may sometimes be recognized a few minutes later by means of the persisting image, although—apparently—the actual impression was entirely disregarded. But in vision the primary memory-image can always be obtained, and is obtained to most advantage, by looking intently at some object for an instant and then closing the eyes or turning them away. The image of the object will appear for a moment very vividly and distinctly, and can be so recovered several times in succession by an effort of attention. Such reinstatement is materially helped by rapidly opening and closing the eyes, or by suddenly moving them in any way. In this respect a primary memory image resembles an after-sensation, which can be repeatedly revived in this manner when it would otherwise have disappeared. This seems to show that the primary memory-image in such cases owes its vivacity in part to a positive after-sensation; at any rate it proves that it is in some way still sense-sustained. But in other respects the two are very different: the after-sensation is necessarily presented if the intensity of the original excitation suffices for its production, and cannot be presented otherwise, however much we attend. Moreover, the after-sensation is only for a moment positive, and then passes into the negative or complementary phase, when, so far from even contributing towards the continuance of the original percept, it directly hinders it. Primary memory-images on the other hand, and indeed all images, depend mainly upon the attention given to the impression; provided that was sufficient, the faintest impression may be long retained, and without it very intense ones will soon leave no trace. The primary memory-image retains so much of its original definiteness and intensity as to make it possible with great accuracy to compare two physical phenomena, one of which is in this way “remembered” while the other is really present. For the most part this is indeed a more accurate procedure than that of dealing with both together, but it is only possible for a very short time. From Weber's experiments with weights and lines[51] it would appear that even after 10 seconds a considerable waning has taken place, and after 100 seconds all that is distinctive of the primary image has probably ceased.

On the whole, then, it appears that the ordinary memory image is a joint effect; it is not the mere residuum of changes in the presentation-continuum, but an effect of these only when there has been some concentration of attention upon them. It has the form of a percept, but is not constituted of revived impressions, for the essential marks of impressions are absent; there is no localization in, or projection into, external space, neither is there the motor adaptation, nor the tone of feeling, incident to the reception of impressions. Ideas do not reproduce the intensity of these original constituents, but only their quality and complication. What we call the vividness of an idea is of the nature of intensity, but it is an intensity very partially and indirectly determined by that of the original impression; it depends much more upon the state of what we shall call the memory-continuum and the attention the idea receives. The range of vividness in ideas is probably comparatively small; what are called variations in vividness are often really variations in distinctness and completeness.[52] Where we have great intensity, as in hallucinations, primary presentations may be reasonably supposed to enter into the complex.

It is manifest that the memory-continuum has been in someway formed out of or differentiated from the presentation continuum by the movements of attention, but the precise connexion of the two continua is still very difficult to determine. We see perhaps the first distinct step of this evolution in the primary memory-image: here there has been no cessation in presentation, and yet the characteristic marks of the impression are gone, so much so, indeed, that superposition without “fusion” with an exactly similar impression is possible. We have now to inquire into the genesis and development of ideation.

Genesis and Development of Ideation.

23. We find ourselves sometimes engrossed in present perceptions, as when tracing, for example, the meanderings of an ant; at other times we may be equally absorbed in reminiscences; or, again, in pure reverie and “castle-building.” Here are three well-marked forms of conscious life: the first being concerned with what is, the second with what has been, and the third with the merely possible. Again, the first involves definite spatial and temporal order, though the temporal order, as just said, is in the main restricted to the “sensible present ”; the second involves only definite time-order; and the last neither in a definite way. Thus, analytically regarded, perception, memory, imagination, show a steady advance. In infancy the first predominates, while senility lapses back to the second; in the third, where similarities suggest themselves and the contrast of actual and possible is explicit, we have at length the groundwork of logical comparison. Nevertheless, since imagination plays a conspicuous part in child life before much personal reminiscence appears, it would seem probable that ideas do not first arise as definite memory-images or reminiscences. On the other hand, in the so-called homing instincts of the lower animals we have evidence of isolated “memories” of a simpler form than ours.

The subject is as difficult as it is interesting and important, and we can hardly hope at present for a final solution. One chief obstacle, as is so often the case in psychology, lies in the unsettled connotation of such leading terms as memory, association and idea. Even what is most fundamental of all, that “plasticity” which we have analysed into retentiveness, differentiation and integration, is sometimes described as if it already involved memory-ideas and their association. Ideas, that is to say, are identified with mere “residua” of former “impressions,” and yet at the same time are spoken of as “copies” of these: which is much like saying the evening twilight is a replica of the noonday glare as well as its parting gleam. Again, the continuous differentiation and red integration of the presentational continuum which mark the progress of perceptual experience are resolved into an original multiplicity of presentational atoms which are associated by “adhesion” of the contiguous. Yet before the differentiation there was no plurality, and after the integration there is only a complex unity, comparable perhaps with another organic whole, but certainly not with a mosaic stuck together with cement. This mistaken identification by the Associationist psychology of later processes with simpler and earlier ones, by which they are only partially explained, has not only obscured the science with inappropriate concepts but has prevented the question on which we are entering—that concerning the genesis and development of ideas—from being ever effectually raised. The discussion of this question will incidentally yield the best refutation of those views.

Experience, we say, is the acquisition of practical acquaintance and efficiency as the result of repeated opportunity and effort. This means that strangeness on the cognitive side gives place to familiarity, and that on the active side clumsiness is superseded by skill. But though analytically distinct, the two sides are, as we have already insisted, actually inseparable: to the uninteresting we are indifferent, and what does not call for active response is ignored. If the original presentations whether sensory or motor, be A, B, C, we find then that they gradually acquire a. new character, become, let us say, Aγ, Bγ, Cγ, γ representing the eventual familiarity or facility, as the case may be. We find, again, a certain sameness in this character, however various the presentations to which it pertains, a sameness which points to the presence of subjective constituents, and to these we may assign the “feelings” that enter into accommodation and adjustment. This factor is important as evidence of a subjective co-operation which may enable us to dispense with the mutual “adhesions” and “attractions” among presentations, on which the Associationists rely. But it is obvious that there must be an objective factor as well; and it is this objective factor in the process giving rise to γ that now primarily concerns us. We have described that process as assimilation or immediate recognition: the older psychology described it as association of the completely similar, or as automatic association. That the two views have something in common is shown by the juxtaposition of “automatic” and “immediate,” “similarity” and “assimilation.” To prepare the way for further discussion, let us first ascertain these points of agreement. “When I look at the full moon,” said Bain, “I am instantly impressed with the state arising from all my former impressions of her disc added together.” This we may symbolize in the usual fashion as A + an ⋅ ⋅ ⋅ + a₃ + a₂ + a₁. Now, it will be granted (1) that the present occurrence (full moon) has been preceded by a series of like occurrences, enumerable as 1, 2, 3, ⋅ ⋅ ⋅ n; (2) that the present experience (Aγ) is what it is in consequence of the preceding experiences of these occurrences; and (3) that it “arises instantly” as the joint result of such preceding experiences. But it is denied (1) that this present experience is the mere sum, or even the mere “fusion,” of the experiences preceding it; (2) that they were qualitatively identical; (3) that they persist severally unaltered, in such wise that experience “drags at each remove a lengthening chain” of them. In the case of dexterities, where γ answers to facility, it is obvious that there is no such series of identicals (a₁, a₂, ⋅ ⋅ ⋅ an) at all. From the first rude beginning—say the schoolboy's pothooks—up to the finished performance of the adept there is continuous approximation: awkward and bungling attempts, passing gradually into the bold strokes of mastery. Nor is the case essentially different in cognition where γ answers to familiarity; if we attend, as it is plain we ought, not to the physical fact cognized, but to the individual's perception of it. This, too, is an acquisition, has entailed activity, and is marked by gradual approximation towards clearness and distinctness. The successive experiences of n identical occurrences does not then result in an accumulation of n identical residua. The ineptness of the atomistic psychology with its “physical” and “chemical” analysis is nowhere more apparent than here. Considering the intimate relation of life and mind, and the strong physiological bias shown by the Associationists from Hartley onwards, it is surely extraordinary how completely they have failed to appreciate the light-bearing significance of such concepts as function and development. Facility and faculty (or function) are much the same, both etymologically and actually. As the perfected structure is not so many rudimentary structures “added together,” but something that supersedes them completely, must we not say the same of the perfected function? The less fit is not embodied in the fittest that finally survives. Development implies change of form in a continuous whole: every growth into means an equal growth out of: thus one cannot find the caterpillar in the butterfly. Between organic development and mental development there is then more than an analogy.

But though assimilation cannot be analysed into a series of identical ideas (a₁, a₂, ⋅ ⋅ ⋅ an), either “added together” or “instantaneously fused,” yet it does result in an a which may provisionally be called an idea. Such idea is, however, neither a memory-idea in the proper sense nor an idea within the meaning of the term implied in imagination or ideation. For it is devoid of the temporal signs[53] indicated by the subscript numerals in a₁, a₂, ⋅ ⋅ ⋅, and it does not yet admit of reproduction as part of an ideational continuum, one, that is, divested of the characteristics belonging to the actual and sensibly present. It is, so to say, embryonic, something additional to the mere sensation assimilated, and yet something less than a “free or independent idea.” It is, as it has been happily called,[54] a tied (gebundene) or implicit idea. We have clear evidence of the sense-bound stage of this immature “idea” in the so-called “memory after-image” (cf. § 22). There is, however, nothing in this of memory, save as the term is loosely used for mere retentiveness; and after-percept would therefore be a less objectionable name for it. This after-percept is entirely sense-sustained and admits of no ideal recall, though—in minds sufficiently advanced—it may persist for a few moments, and so form the basis of such comparison with a second sensation, as we find in the experiments of Weber, Fechner and others.[55] At a still lower level, or in actual perception, we cannot assume even this amount of partial independence, though continuity clearly points to something beyond the bare sensation, which is a pure abstraction, as we may presently see.

It is saying too little to maintain, as some do, that this “something” is subconscious, on the ground that it is not discoverable by direct analysis. Yet it is saying too much, regardless of this defect, to describe a percept as a preventative-representative complex, if representation is to imply the presence of a free or independent idea. To call this “something” a tied or nascent idea on the ground of its possible later development into an independent representation seems, then, nearest the truth. The same meaning is sometimes expressed in a wholly different and designedly paradoxical way, by saying that all cognition (perception) is recognition. This statement has been met by elaborate expositions of the difference between knowing and knowing again, the irrelevance of which any lexicon would show; and, further, by the demand: How on such a view is a first cognition possible, or how is an indefinite regress of assimilation to be avoided? We may confidently reply that it cannot be avoided: an absolute beginning of experience, whether phylogenetically or ontogenetically, is beyond us. Assimilation means further assimilation; in this sense all cognition is further cognition, and a bare sensation is, as said, an abstraction representing a limit to which we can never regress.

We find evidence, again, of ideas in the making in what Lewes called preperception. Of this instances in plenty are furnished by everyday illusions, as when a scarecrow is hailed by the traveller who mistakes it for a husbandman, or when what is taken for an orange proves to be but an imitation in wax. In reality all complex percepts involve preperception; and, so far, it must be allowed that such percepts are directly analysable into preventative-representative complexes. Nevertheless, the representative element is not yet, and may never become, an idea proper. The sight of ice yields a forefeel of its coldness, the smell of baked meats a foretaste of their savour. Such pre-percepts differ from free ideas just as after-percepts do: they are still sense-bound and sense-sustained. Nor can this complication be with any propriety identified either with the association pertaining to memory or with that specially pertaining to ideation; though, no doubt, the two processes—complication and association—are genetically continuous, as are their respective constituents, nascent and free ideas.[56] The whole course of perceptual integration being determined and sustained by subjective interest, involves from the outset, as we have seen, concurrent conative impulses; and thus the same assimilation that results in familiarity and preperception on the subjective side results in facility and purpose on the conative. Knowing immediately what to do is here the best evidence of knowing what there is to do with; the moth that flies into the candle has assuredly no preperception of it, and does not act with purpose. Bearing this in mind, we may now see one way, and probably the earliest, in which tied ideas become free.

The contrast between the actual and the possible constitutes, as we have seen, the main difference between experience at the perceptual and experience at the ideational stage. A subject confined to the former level knows not yet this difference. Such knowledge is attained, not through any quasi-mechanical interaction of presentations, but usually through bitter experience. The chapter of accidents is the Bible of fools, it has been said; but we are all novices at first, and get wisdom chiefly by the method of trial and failure. Things are not always different in what to us are their essential properties, but they so differ from time to time. Resemblances are frequent enough to give us familiarity and confidence; yet uniformity is flecked by diversity, and thwarted intentions disclose possibilities for which we were not prepared. What was taken for sugar turns out to be salt; what was seized as booty proves to be bait. We catch many Tatars, and so learn wariness in a rough school. In such wise preperceptions displaced by the actual fact yield the “what” severed from the “that,” the “ideal” freed at length from the exclusive hold of the real. In a new situation after such adventures the attitude assumed—if, for brevity, we describe it in terms of our own still more advanced experience—is of this sort: “It may be a weasel, if so, I back; it may be a rabbit, if it is, I spring.” Instead of unquestioned preperception that “makes the mouth water,” we have the alternative possibilities present as “free ideas,” and action is in suspense, the alternative courses, that is to say, again present only in idea. It is easy to see how in such situations one free idea, a “what” sundered from its “that,” will tend to loosen the sensory ties of alternative, still implicit ideas. On the cognitive side, from immediate assimilation an advance is made towards mediate cognition, towards comparison; on the active side there is advance from impulsive action towards deliberate action.[57]

We conclude, then, that implicit ideas—the products of assimilation, and integrated as such in complex percepts and the motor co-ordinations to which they lead—are more likely to emerge as free ideas the more this perceptual complexity increases. Perception in the lower animals, who give no signs of either memory or ideation, has apparently no such complexity. A fish, for example, can feel, smell, taste, see, and even hear, but we cannot assume solely on that account that it has any percepts to which its five senses contribute, as they do to our percept, say, of an orange or a peppermint. Taking voluntary movements as the index of psychical life, it would seem that the fish's movements are instigated and guided by its senses, not collectively but separately. Thus a dog-fish, according to Steiner, seeks its food exclusively by scent; so that when its olfactory bulbs are severed, or the fore-brain, in which they end, is destroyed, it ceases to feed spontaneously. The carp, on the other hand, appears to search for its food wholly under the guidance of sight, and continues to do so just as well when the fore-brain is removed, the mid-brain, whence the optic nerves spring, seeming to be the chief seat of what intelligence it has.[58] Again, Bateson observes: “There can be no doubt that soles also perceive objects approaching them, for they bury themselves if a stroke at them is made with a landing-net; yet they have no recognition of a worm hanging by a thread immediately over their heads, and will not take it even if it touch them, but continue to feel for it aimlessly on the bottom of the tank, being aware of its presence by the sense of smell.”[59] To this inability to combine simple percepts into one complex percept of a single object or situation we may reasonably attribute the fish's lack of true ideas, and consequent lack of sagacity. The sagacity even of the higher animals does not amount to “general intelligence,” such as enables a child “to put two and two together,” as we say, whatever “two and two” may stand for. So far as life consists of a series of definite situations and definite acts, so far the things done or dealt with together, the contents of the several foci or concentrations of attention, form so many integrated and comparatively isolated wholes. Round the more complicated of these, and closely connected with them, free ideas arise as sporadic groups, making possible those “lucid intervals,” those fitful gleams of intelligence in the very heat of action, which occasionally interrupt the prevailing irrationality of the brutes. And as we cannot credit even the higher animals with general trains of ideas, just as little can we credit them with a continuous memory: indeed, it is questionable how far memory of the past, as past, belongs to them at all. For they live entirely in an up-stream, expectant attitude, and it is in this aspect that “free ideas” arise when they arise at all. We cannot imagine a dog regretting, like one of Punch's heroes, that he “did not have another slice of that mutton.”[60]

The free idea (a) then at its first emergence has neither an assignable position in a continuous memory-record, as a₁ or a₂, nor has it a definite relation as a “generic idea” to possible specializations such as a′ or a″. These further developments bring us to the general consideration of mental association.

Mental Association and the Memory-Continuum.

24. Great confusion has been occasioned, as we have seen incidentally, by the lax use of the term “association,” and this confusion has been increased by a further laxity in the use of the term “association by similarity.” In so far as the similarity amounts to identity, as in Association by Similarity not Fundamental. assimilation, we have a process which is more fundamental than association by contiguity, but then it is not a process of association. And when the reviving presentation is only partially similar to the presentation revived, the nature of the association does not appear to differ from that operative when one “contiguous” presentation revives another. In the one case we have, say, a b x recalling a b y, and in the other a b c recalling d e f. Now anybody who will reflect must surely see that the similarity between a b x and a b y, as distinct from the identity of their partial constituent a b, cannot be the means of recall; for this similarity is nothing but the state of mind—to be studied presently—which results when a b x and a b y, having been recalled are in consciousness together and then compared. But if a b, having concurred with y before and being now present in a b x, again revives y, the association, so far as that goes, is manifestly one of contiguity, albeit as soon as the revival is complete, the state of mind immediately incident may be what Bain loved to style “the flash of similarity.” So far as the mere revival itself goes, there is no more similarity in this case than there is when a b c revives d e f. For the very a b c that now operates as the reviving presentation was obviously never in time contiguous with the d e f that is revived; if all traces of previous experiences of a b c were obliterated there would be no revival. In other words, the a b c now present must be “automatically associated,” or, as we prefer to say, must be assimilated to those residua of a b c which were “contiguous” with d e f, before the representation of this can occur. And this, and nothing more than this, we have seen, is all the “similarity” that could be at work when a b x “brought up” a b y.

On the whole, then, we may assume that the only principle of association we have to examine is the so-called association by contiguity, which, as ordinarily formulated, runs: Any presentations whatever, which are in consciousness together or in close succession, cohere in such a Contiguity Inexplicable. way that when one recurs it tends to revive the rest, such tendency increasing with the frequency of the conjunction. It has been often contended that any investigation into the nature of association must be fruitless.[61] But, if association is thus a first principle, it ought at least to admit of such a statement as shall remove the necessity for inquiry. So long, however, as we are asked to conceive presentations originally distinct and isolated becoming eventually linked together, we shall naturally feel the need of some explanation of the process, for neither the isolation nor the links are clear—not the isolation, for we can only conceive two presentations separated by other presentations intervening; nor the links, unless these are also presentations, and then the difficulty recurs. But, if for contiguity we substitute continuity and regard the associated presentations as parts of a new continuum, the only important inquiry is how this new whole was first of all integrated.

To ascertain this point we must examine each of the two leading divisions of contiguous association—that of simultaneous presentations and that of presentations occurring in close succession. The last, being the clearer, may be taken first. In a series of associated presentations Formation of Memory-Continuum. A B C D E, such as the movements made in writing, the words of a poem learned by heart, or the simple letters of the alphabet themselves, we find that each member recalls its successor but not its predecessor. Familiar as this fact is, it is not perhaps easy to explain it satisfactorily. Since C is associated both with B and D, and apparently as intimately with the one as with the other, why does it revive the later only and not the earlier? B recalls C; why does not C recall B? We have seen that any reproduction at all of B, C or D depends primarily upon its having been the object of special attention, so as to occupy at least momentarily the focus of consciousness. Now we can in the first instance only surmise that the order in which they are reproduced is determined by the order in which they were thus attended to when first presented. The next question is whether the association of objects simultaneously presented can be resolved into an association of objects successively attended to. Whenever we try to recall a scene we saw but for a moment there are always a few traits that recur, the rest being blurred and vague, instead of the whole being revived in equal distinctness or indistinctness. On seeing the same scene a second time our attention is apt to be caught by something unnoticed before, as this has the advantage of novelty; and so on, till we have “lived ourselves into” the whole, which may then admit of simultaneous recall. Bain, who is rightly held to have given the best exposition of the laws of association, admits something very like this in saying that “coexistence is an artificial growth formed from a certain peculiar class of mental successions.” But, while it is easy to think of instances in which the associated objects were attended to successively, and we are all perfectly aware that the surest—not to say the only—way to fix the association of a number of objects is by thus concentrating attention on each in turn, it seems hardly possible to mention a case in which attention to the associated objects could not have been successive. In fact, an aggregate of objects on which attention could be focused at once would be already associated.

The exclusively successional character of contiguous association has recently been denied, and its exclusively simultaneous character maintained instead. It is at once obvious that this opposition of succession and simultaneity cannot be pressed so as to exclude duration altogether and reduce the whole process to an instantaneous event. Nor is there any ground for saying that there is a fixed and even distribution of attention to whatever is simultaneously presented: facts all point the other way. Still, though we cannot exclude the notion of process from consciousness, we may say that presentations attended to together become pro tanto a new whole, are synthesized or complicated. Such primary synthesis leads not to an association of ideas, but rather to the formation of one percept, which may become eventually a free idea. The disconcerted preperception which sets this free may likewise liberate a similar or contrasting idea, but it will not resolve either complex into the several “ideas” of its sensory or motor constituents, with which only the psychologist is familiar. The actual recurrence of some of these constituents may again reinstate the rest, not, however, as memories or as “thoughts,” but only as tied ideas in a renewed perception.

Again, it has become usual to distinguish the association of contiguous experiences and the so-called association of similars or opposites as respectively external and internal forms of association. The new terminology is illuminating: the substitution of forms for laws marks the abandonment of the old notion that association was by “adhesion” of the contiguous and “attraction” of the similar. We are thus left to find the cause of association in interested attention; and that, we may safely say, is an adequate, and apparently the sole adequate, cause for the two commonly recognized forms of external association, the so-called simultaneous and the successive. But these two are certainly not co-ordinate; and if our analysis be sound, the former—for which we would retain the Herbartian term complication—yields us not members of an association but a member for association. So far, then, we should have but one form of association, that of the successive contents of localized attention: and but one result, the representation or memory-continuum,[62] in contrast to the primary- or presentation-continuum, whence its constituents arise. Turning now to the distinction of external and internal, it at once strikes the unprejudiced mind that “internal association” is something of an anomaly, since the very notion of association implies externality. Also, on closer inspection what we find is not an association of similars or opposites as such, but—something quite distinct—a similarity or contrast of associates; of ideas, that is to say, which are contiguous members of the memory (or experience) continuum, or of ideas which have become contiguous through its reduplication.

The only case, then, that now remains to be considered is that—to take it in its simplest form—of two primary presentations A and X, parts of different special continua or distinct—i.e. non-adjacent—parts of the same, and occupying the focus of consciousness in immediate succession. This constitutes their integration; for the result of this occupation may be regarded as a new continuum in which A and X become adjacent parts. For it is characteristic of a continuum that an increase in the intensity of any part leads to the intenser presentation of adjacent parts; and in this sense A and X, which were not originally continuous, have come to be so. We have here, then, some justification for the term secondary- or memory-continuum when applied to this continuous series of representations to distinguish it from the primary or presentation-continuum from which its constituents are derived. The most important peculiarity of this continuum, therefore, is that it is a series of representations integrated by means of the movements of attention out of the differentiations of the primary or presentation-continuum, or rather out of so much of these differentiations as pertain to what we know as the primary memory-image. These movements of attention, if the phrase may be allowed, come in the end to depend mainly upon interest, but at first appear to be determined entirely by mere intensity.[63] To them it is proposed to look for that continuity which images lose in so far as they part with the local signs they had as percepts and cease to be either localized or projected. Inasmuch as it is assumed that these movements form the connexion between one representation and another in the memory-train, they may be called “temporal signs.”[64] The evidence for their existence can be more conveniently adduced presently; it must suffice to remark here that it consists almost wholly of facts connected with voluntary attention and the voluntary control of the flow of ideas, so that temporal signs, unlike local signs, are fundamentally motor and not sensory. And, unlike impressions, representations can have each but a single sign,[65] the continuum of which, in contrast to that of local signs, is not rounded and complete, but continuously advancing. But in saying this we are assuming for a moment that the memory-continuum forms a perfectly single and unbroken train. If it ever actually were such, then, in the absence of any repetition of old impressions and apart from voluntary interference with the train, consciousness, till it ceased entirely, would consist of a fixed and mechanical round of images. Some approximation to such a state is often found in uncultured persons who lead uneventful lives, and still more in idiots, who can scarcely think at all.

25. In actual fact, however, the memory-train is liable to change in two respects, which considerably modify its structure, viz. (1) through the evanescence of some parts, and (2) through the partial recurrence of like impressions, which produces reduplications of varying amount and extent in other parts. As regards the first, we may infer that the waning or sinking towards the threshold of consciousness which we can observe Formation of Ideational Continuum. in the primary mental image continues in subconsciousness after the threshold is past. For the longer the time that elapses before their revival the fainter, the less distinct, and the less complete are the images when revived, and the more slowly they rise. All the elements of a complex are not equally relivable, as we have seen already: tastes, smells and organic sensations, though powerful as impressions to revive other images, have little capacity for ideal reproduction themselves, while muscular movements, though perhaps of all presentations the most readily revived, do not so readily revive other presentations. Idiosyncrasies are, however, frequent; thus we find one person has an exceptional memory for sounds, another for colours, another for forms. Still it is in general true that the most intense, the most impressive, and the most interesting presentations persist the longest. But the evanescence, which is in all cases comparatively rapid at first, deepens sooner or later into real or apparent oblivion. In this manner it comes about that parts of the memory-continuum Obliviscence. lose all distinctness of feature and, being without recognizable content, shrivel up to a dim and meagre representation of life that has lapsed—a representation that just suffices, for example, to show us that “our earliest recollections” are not of our first experiences, or to save them from being not only isolated but discontinuous. Such discontinuity can, of course, never be absolute; we must have something represented even to mark the gap. Oblivion and the absence of all representation are thus the same, and the absence of all representation cannot psychologically constitute a break. The terms “evolution” and “involution” have in this respect been happily applied to the rising and falling of representations. When we recall a particular period of our past life, or what has long ceased to be a familiar scene, events and features gradually unfold and, as it were, spread out as we keep on attending. A precisely opposite process may then be supposed to take place when they are left in undisturbed forgetfulness; this process is called obliviscence.

More important changes are produced by the repetition of parts of the memory-train. The effect of this is not merely Repetition. to prevent the evanescence of the particular image or series of images, but by partial and more or less frequent reduplications of the memory-train or “thread” upon itself to convert it into a partially new continuum, which we might perhaps call the ideational continuum or “tissue.”[66] The reduplicated portions of the train are strengthened, while at the points of divergence it becomes comparatively weakened, and this apart from the effects of obliviscence. One who had seen the king but once would scarcely be likely to think of him without finding the attendant circumstances recur as well; this could not happen after seeing him in a hundred different scenes. The central representation of the whole complex would Generic Images. have become more distinct, whereas the several diverging lines would tend to dissipate attention and, by involving opposing representations, to neutralize each other, so that probably no definite background would be reinstated. Even this central representation would be more or less generalized. It has been often remarked that one's most familiar friends are apt to be mentally pictured less concretely and vividly than persons seen more seldom and then in similar attitudes and moods; in the former case a “generic image” has grown out of such more specific representations as the latter affords. Still further removed from memory-images are the images that result from such familiar percepts as those of horses, houses, trees, &c.

Thus as the joint effect of obliviscence and reduplication we are provided with trains of ideas distinct from the memory thread Trains of Ideas. and thereby with the material, already more or less organized, for intellectual and volitional manipulation. We do not experience the flow of ideas—save very momentarily and occasionally—altogether undisturbed; even in dreams and reverie it is continually interrupted and diverted. Nevertheless it is not difficult to ascertain that, so far as it is left to itself, it takes a very different course from that which we should have to retrace if bent on reminiscence and able to recollect perfectly. The readiness and steadiness of this flow are shown by the extremely small effort necessary in order to follow it. Nevertheless from its very nature it is liable, though not to positive breaches of continuity from its own working, yet to occasional blocks or impediments to the smooth succession of images at points where reduplications diverge, and either permanently or at the particular time neutralize each other.[67]

The flow of ideas is, however, exposed to positive interruptions from two distinct sides—by the intrusion of new presentations and Conflict of Presentations. of voluntary interference. The only result of such interruptions which we need here consider is the conflict of presentations that may ensue. Herbart and his followers have gone so far as to elaborate a complete system of psychical statics and dynamics, based on the conception of presentations as forces and on certain more or less improbable assumptions as to the modes in which such forces interact. Since our power of attention is limited, it continually happens that attention is drawn off by new presentations at the expense of old ones. But, even if we regard this non-voluntary redistribution of attention as implying a struggle between presentations, still such conflict to secure a place in consciousness is very different from a conflict between presentations that are already there. Either may be experienced to any degree possible without the other appearing at all; thus, absorbed in watching a starry sky, one might be unaware of the chilliness of the air, though recognizing at once, as soon as the cold is felt, that, so far from being incompatible, the clearness and the coldness are causally connected. This difference between a conflict of presentations to enter the field of consciousness—if we allow for a moment the propriety of the expression—and that opposition or incompatibility between presentations which is only possible when they are in consciousness has been strangely confused by the Herbartians. In the former the intensity of the presentation is primarily alone of account; in the latter, on the contrary, quality and content are mainly concerned. Only the last requires any notice here, since such opposition arises when the ideational continuum is interrupted in the ways just mentioned, and apparently arises in no other way. Certainly there is no such opposition between primary presentations: there we have the law of incopresentability preventing the presentation of opposites with the same local sign; and their presentation with different local signs involves, on this level at all events, no conflict. But what has never been presented could hardly be represented, if the ideational process were undisturbed: even in our dreams white negroes or round squares, for instance, never appear. In fact, absurd and bizarre as dream-imagery is, it never at any moment entails overt contradictions, though contradiction may be implicit.

But between ideas and percepts actual incompatibility is frequent. In the perplexity of Isaac, e.g.—“The voice is Jacob's voice, but the hands are the hands of Esau”—we have such a case in a familiar form. There is here not merely mental arrest but actual conflict: the voice perceived identifies Jacob, at the same time the hands identify Esau. The images of Esau and Jacob by themselves are different, but do not conflict; neither is there any strain, quite the contrary, in recognizing a person partly like Jacob and partly like Esau. For there is no direct incompatibility between smooth and rough, so long as one pertains only to voice and the other only to hands, but the same hands and voice cannot be both smooth and rough. Similar incompatibilities may arise without the intrusion of percepts, as when, in trying to guess a riddle or to solve a problem, or generally to eliminate intellectual differences, we have images which in themselves are only logically opposite, psychologically opposed, or in conflict, because each strives to enter the same complex. In all such conflicts alike we find, in fact, a relation of presentations the exact converse of that which constitutes similarity. In the latter we have two complete presentations, a b x and a b y, as similar, each including the common part a b; in the former we have two partial presentations, x and y, as contraries, each excluding the other from the incomplete a b—. And this a b, it is to be noted, is not more essential to the similarity than to the conflict. But in the one case it is a generic image (and can logically be predicated of two subjects); in the other it is a partially determined individual (and cannot be subject to opposing predicates). Except as thus supplementing a b, x and y do not conflict; black and white are not incompatible save as attributes of the same thing. The possibility of most of these conflicts—of all, indeed, that have any logical interest—lies in that reduplication of the memory-continuum which gives rise to these new complexes, generic images or general ideas.

Reminiscences and Expectation: Temporal Perception.

26. Having thus attempted to ascertain the formation of the ideational continuum out of the memory-train, the question arises: How now are we to distinguish between imagining and remembering, and again, between imagining and expecting? It is plainly absurd to make the difference depend on the presence of belief in memory and expectation and on its absence in mere imagination; for the belief itself depends on this difference instead of constituting it. One real and obvious distinction, however which Hume pointed out as regards memory, is the fixed order and position of the ideas of what is remembered or expected as Imagination and Memory. contrasted with “the liberty” of the imagination to transpose and change its ideas. This order and position in the case of memory are, of course, normally those of the original impressions, but it seems rather naïve of Hume to tell us that memory “is tied down to these without any power of variation,” while imagination has liberty to transpose as it pleases, as if the originals sat to memory for their portraits, while to imagination they were but studies. Such correspondence being out of the question—as Hume takes care to state as soon as it suits him—all we have, so far, is this fixity and definiteness as contrasted with the kaleidoscopic instability of ideation. In this respect what is remembered or expected resembles what is perceived: the grouping not only does not change capriciously and spontaneously, but resists any mental efforts to change it. But, provided these characteristics are there, we should be apt to believe that we are remembering, just as, mutatis mutandis, with like characteristics we might believe that we were perceiving: hallucination is possible in either case.

This fixity of order and position is, however, not sufficient to constitute a typical reminiscence where the term is exactly used. But remembering is often regarded as equivalent to knowing and recognizing, as when on revisiting some once familiar place one remarks, “How well I remember it!” What is meant is that the place is recognized, and that its recognition awakens memories. Memory includes recognition; recognition as such does not include memory. In human consciousness, as we directly observe it, there is, perhaps, no pure recognition: here the new presentation in not only assimilated to the old, but the former framing of circumstance is reinstated, and so perforce distinguished from the present. It may be there is no warrant for supposing that such red integration of a preceding field is ever absolutely nil, still we are justified in regarding it as extremely vague and meagre, both where mental evolution is but slightly advanced and where frequent repetition in varying and irrelevant circumstances has produced a blurred and neutral zone. The last is the case with a great part of our knowledge; the writer happens to know that bos is the Latin for “ox” and bufo the Latin for “toad,” and may be said to remember both items of knowledge, if “remember” is only to be synonymous with “retain.” But if he came across bos in reading he would think of an ox and nothing more; bufo would immediately call up not only “toad” but Virgil's Georgics, the only place in which he has seen the word, and which he never read but once. In the former there is so far nothing but recognition (which, however, of course rests upon retentiveness); in the latter there is also remembrance of the time and circumstances in which that piece of knowledge was acquired. Of course in so far as we are aware that we recognize we also think that remembrance is at any rate possible, since what we know we must previously have learned—recognition excluding novelty. But the point here urged is that there is an actual reminiscence only when the recognition is accompanied by a reinstatement of portions of the memory-train continuous with the previous presentation of what is now recognized. Summarily stated, we may say that between knowing and remembering on the one hand and imagining on the other the difference primarily turns on the fixity and completeness of the grouping in the former; in the latter there is a shifting play of images more or less “generic,” reminding one of “dissolving views.” Hence the first two approximate in character to perception, and are rightly called recognitions. Between them, again, the difference turns primarily on the presence or absence of temporal signs. In what is remembered these are still intact enough to ensure a localization in the past of what is recognized; in what is known merely such localization is prevented, either because of the obliviscence of temporal connexions or because the reduplications of the memory-train that have consolidated the central group have entailed their suppression. There is further the difference first mentioned, which is often only a difference of degree, viz. that reminiscences have more circumstantiality, so to say, than mere recognitions have: more of the collateral constituents of the original concrete field of consciousness are reinstated. But of the two characteristics of memory proper—(a) concreteness or circumstantiality, and (b) localization in the past—the latter is the more essential. It sometimes happens that we have the one with little or nothing of the other. For example, we may have but a faint and meagre representation of a scene, yet if it falls into and retains a fixed place in the memory train we have no doubt that some such experience was once actually ours. On the other hand, as in certain so-called illusions of memory, we may suddenly find ourselves reminded by what is happening at the moment of a preceding experience exactly like it—some even feel that they know from what is thus recalled what will happen next; and yet, because we are wholly unable to assign such representation a place in the past, instead of a belief that it happened, there arises a most distressing sense of bewilderment, as if one were haunted and had lost one's personal bearings.[68] It has been held by some psychologists[69] that memory proper includes the representation of one's past self as agent or patient in the event or situation recalled. And this is true as regards all but the earliest human experience, at any rate; still, whereas it is easy to see that memory is essential to any development of self consciousness, the converse is not at all clear, and would involve us in a needless circle.

27. Intimately connected with memory is expectation. We may as the result of reasoning conclude that a certain event Expectation. will happen; we may also, in like manner, conclude that a certain other event has happened. But as we should not call the latter memory, so it is desirable to distinguish such indirect anticipation as the former from that expectation which is directly due to the interaction of ideas. Any man knows that he will die, and may make a variety of arrangements in anticipation of death, but he cannot with propriety be said to be expecting it unless he has actually present to his mind a series of ideas ending in that of death, such series being due to previous associations, and unless, further, this series owes its representation at this moment to the actual recurrence of some experience to which that series succeeded before. And as familiarity with an object or event in very various settings may be a bar to recollection, so it may be to expectation: the average Englishman, e.g. is continually surprised without his umbrella, though only too familiar with rain, since in our climate one not specially attentive to the weather obtains no clear representation of its successive phases. But after a series of events A B C D E . . . has been once experienced we instinctively expect the recurrence of B C . . . on the recurrence of A, i.e. provided the memory-train continues so far intact. Such expectation, at first perhaps slight—a mere tendency easily overborn—becomes strengthened by every repetition of the series in the old order, till eventually, if often fulfilled and never falsified, it becomes certain and, as we commonly say, irresistible. To have a clear case of expectation, then, it is not necessary that we should distinctly remember any previous experience like it, but only that we should have actually present some earlier member of a series which has been firmly associated by such previous experiences, the remaining members, or at least the next, if they continue serial, being revived through that which is once again realized. This expectation may be instantly checked by reflection, just as it may, of course, be disappointed in fact; but these are matters which do not concern the inquiry as to the nature of expectation while expectation lasts.

We shall continue this inquiry to most advantage by widening it into an examination of the distinction of present, past and future. To a being whose presentations never passed through the transitions which ours undergo—first divested of the strength and vividness of impressions, again reinvested with Present, Past, and Future. them and brought back from the faint world of ideas—the sharp contrasts of “now ” and “ then, ” and all the manifold emotions they occasion, would be quite unknown. Even we, so far as we confine our activity and attention to ideas are almost without them. Time-order, succession, antecedence, and consequence, of course, there might be still, but in that sense of events as “past and gone for ever,” which is one of the melancholy factors in our life; and in the obligation to wait and work in hope or dread to what is “still to come” there is much more than time-order. It is to presentations in their primary stage, to impressions, that we owe what real difference we find between now and then, whether prospective or retrospective, as it is to them also that we directly owe our sense of the real, of what is and exists as opposed to the non-existent that is not. But the present alone and life in a succession of presents, or, in other words, continuous occupation with impressions, give us no knowledge of the present as present. This we first obtain when our present consciousness consists partly of memories or partly of expectations as well. An event expected differs from a like event remembered chiefly in two ways—in its relation to present impressions and images and in the active attitude to which it leads. The diverse feelings that accompany our intuitions of time and contribute so largely to their colouring are mainly consequences of these differences. Let us take a series of simple and familiar events A B C D E, representing ideas by small letters, and perceptions by capitals whenever it is necessary to distinguish them. Such series may be present in consciousness in such wise that a b c d are imaged while E is perceived anew, i.e. the whole symbolized as proposed would be a b c d E; such would be, e.g. the state of a dog that had just finished his daily meal. Again, there may be a fresh impression of A which revives b c d e; we should have then (1) A b c d e—the state of our dog when he next day gets sight of the dish in which his food is brought to him. A little later we may have (2) a b C d e. Here a b are either after-sensations or primary memory-images, or have at any rate the increased intensity due to recent impression; but this increased intensity will be rapidly on the wane even while C lasts, and a b will pale still further when C gives place to D, and we have (3) a b c D e. But, returning to (2), we should find d e to be increasing in intensity and definiteness, as compared with their state in (1), now that C, instead of A, is the present impression. For, when A occupied this position, not only was e raised less prominently above the threshold of consciousness by reason of its greater distance from A in the memory-continuum, but, owing to the re duplications of this continuum, more lines of possible revival were opened up, to be successively negatived as B succeeded to A and C to B; even dogs know that “there is many a slip 'twixt the cup and the lip.” But, where A B C D E is a series of percepts such as we have here supposed—and a series of simpler states would hardly afford much ground for the distinctions of past, present and future—there would be a varying amount of active adjustment of sense-organs and other movements supplementary to full sensation. In (2), the point at which we have a b C d e, for instance, such adjustments and movements as were appropriate to b would cease as B lapsed and be replaced by those appropriate to C. Again, as C succeeded to B, and d in consequence increased in intensity and definiteness, the movements adapted to the reception of D would become nascent, and so on. Thus, psychologically regarded, the distinction of past and future and what we might call the oneness of direction of time depend, as just described, (1) upon the continuous sinking of the primary memory-images on the one side, and the continuous rising of the ordinary images on the other side, of that member of a series of percepts then repeating which is actual at the moment; and (2) on the prevenient adjustments of attention, to which such words as “expect,” “await,” “anticipate,” all testify by their etymology. These conditions in turn will be found to depend upon all that is implied in the formation of the memory-train and upon that recurrence of like series of impressions which we attribute to the “uniformity of nature.” If we never had the same series of impressions twice, knowledge of time would be impossible, as indeed would knowledge of any sort.

28. Time is often figuratively represented as a line, and we may perhaps utilize this figure to make clear the relation of our Succession. perception of time to what we call time itself. The present, though conceived as a point or instant of time, is still such that we actually can and do in that moment attend to a plurality of presentations to which we might otherwise have attended to severally in successive moments. Granting this implication of simultaneity and succession, we may, if we represent succession as a line, represent simultaneity as a second line at right angles to the first; pure time—or time-length without time-breadth, we may say—is a mere abstraction. Now it is with the former line that we have to do in treating of time as it is (or as we conceive it), and with the latter in treating of our perception of time, where, just as in a perspective representation of distance, we are confined to lines in a plane at right angles to the actual line of depth. In a succession of events A B C D E . . . the presence of B means the absence of A and of C, but the presentation of this succession involves the simultaneous presence, in some mode or other, of two or more of the presentations A B C D. In our temporal perception, then, all that corresponds to the differences of past, present and future is presented simultaneously. To this fact the name of “specious present” or “psychical present” has been given. What we have is not a moving point or moment of objective time, but rather a moving line, the contents of which, continuously changing, simultaneously represent a portion of the line of objective succession, viz. the immediate past as still present in primary memory-images, and the immediate future as anticipated in prepercepts and nascent acts.[70] This truism—or paradox—that all we know of succession is but an interpretation of what is really simultaneous or coexistent, we may then concisely express by saying that we are aware of time only through time-perspective, and experience shows that it is a long step from a succession of presentations to such presentation of succession. The first condition of such presentation is that we should have represented together presentations that were in the first instance attended to successively, and this we have both in the persistence of primary memory-images and in the simultaneous reproduction of longer or shorter portions of the memory-train. In a series thus secured there may be time-marks, though no time, and by these marks the series will be distinguished from other simultaneous series. To ask which is first among a number of simultaneous presentations is unmeaning; one might be logically prior to another, but in time they are together and priority is excluded. Nevertheless after each distinct representation a, b, c, d there probably follows, as we have supposed, some trace of that movement of attention of which we are aware in passing from one presentation to another. In our present reminiscences we have, it must be allowed, little direct proof of this interposition, though there is strong indirect evidence of it in the tendency of the flow to follow the order in which the presentations were first attended to. With the movements themselves we are familiar enough, though the residua of such movements are not ordinarily conspicuous. These residua, then, are our temporal signs, and, together with the representations connected by them, constitute the memory-continuum. But temporal signs alone will not furnish all the pictorial exactness of the time-perspective. They give us only a fixed series; but the working of obliviscence, by insuring a progressive variation in intensity and distinctness as we pass from one member of the series to the other, yields the effect which we call time-distance. By themselves such variations would leave us liable to confound more vivid representations in the distance with fainter ones nearer the present, but from this mistake the temporal signs save us; and, as a matter of fact, where the memory-train is imperfect such mistakes continually occur. On the other hand, where these variations are slight and imperceptible, though the memory-continuum preserves the order of events intact, we have still no such distinct appreciation of comparative distance in time as we have nearer the present where these perspective effects are considerable.

29. When in retrospect we note that a particular presentation X has had a place in the field of consciousness, while certain Duration. other presentations, A B C D . . ., have succeeded each other, then we may be said in observing this relation of the two to perceive the duration of X. And it is in this way that we do subjectively estimate longer periods of time. But first, it is evident that we cannot apply this method to indefinitely short periods without passing beyond the region of distinct presentation; and, since the knowledge of duration implies a relation between distinguishable presentations A B C D and X, the case is one in which the hypothesis of subconsciousness can hardly help any but those who confound the fact of time with the knowledge of it. Secondly, if we are to compare different durations at all, it is not enough that one of them should last out a series A B C D, and another a series L M N O; we also want some sort of common measure of those series. Locke was awake to this point, though he expresses himself vaguely (Essay, ii. 14, §§ 9-12). He speaks of our ideas succeeding each other “at certain distances not much unlike the images in the inside of a lantern turned round by the heat of a candle,” and “guesses” that “this appearance of theirs in train varies not very much in a waking man.” Now what is this “distance” that separates A from B, B from C, and so on, and what means have we of knowing that it is tolerably constant in waking life? It is probably that the residuum of which we have called a temporal sign; or, in other words, it is the movement of attention from A to B. But we must endeavour here to get a more exact notion of this movement. Everybody knows what it is to be distracted by a rapid succession of varied impressions, and equally what it is to be wearied by the slow and monotonous recurrence of the same impressions. Now these “feelings” of distraction and tedium owe their characteristic qualities to movements of attention. In the first, attention is kept incessantly on the move; before it is accommodated to A, it is disturbed by the suddenness, intensity, or novelty of B; in the second, it is kept all but stationary by the repeated presentation of the same impression. Such excess and defect of surprises make one realize a fact which in ordinary life is so obscure as to escape notice. But recent experiments have set this fact in a more striking light, and made clear what Locke had dimly before his mind in talking of a certain distance between the presentations of a waking man. In estimating very short periods of time, of a second or less—indicated say by the beats of a metronome—it is found that there is a certain period for which the mean of a number of estimates is correct, while shorter periods are on the whole over-estimated, and longer periods under-estimated. This we may perhaps take to be evidence of the time occupied in accommodating or fixing attention. Whether the “point of indifference” is determined by the rate of usual bodily movement, as Spencer asserts and Wundt conjectures, or conversely, is a question we need not discuss just now. But, though the fixation of attention does of course really occupy time, it is probably not in the first instance perceived as time, i.e. as continuous “protensity,” to use a term of Hamilton's, but as intensity. Thus, if this supposition be true, there is an element in our concrete time-perception which has no place in our abstract conception of time. In time conceived as physical there is no trace of intensity; in time psychically experienced duration is primarily an intensive magnitude, witness the comparison of times when we are “bored” with others when we are amused. It must have struck every one as strange who has reflected upon it that a period of time which seems long in retrospect—such as an eventful excursion—should have appeared short in passing; while a period, on the contrary, which in memory has dwindled to a wretched span seemed everlasting till it was gone. But, if we consider that in retrospect length of time is represented primarily and chiefly by impressions that have survived, we have an explanation of one-half; and in the intensity of the movements of attention we shall perhaps find an explanation of the other. What tells in retrospect is the series a b c d e, &c.; what tells in the wearisome present is the intervening ttt₃, &c., or rather the original accommodation of which these temporal signs are the residuum. For, as we have seen elsewhere, the intensity of a presentation does not persist, so that in memory the residuum of the most intense feeling of tedium may only be so many t’s in a memory-continuum whose surviving members are few and uninteresting. But in the actual experience, say, of a wearisome sermon, when the expectation of release is continually balked and attention forced back upon a monotonous dribble of platitudes, the one impressive fact is the hearer's impatience. On the other hand, so long as we are entertained, attention is never involuntary, and there is no continually deferred expectation. Just as we are said to walk with least effort when our pace accords with the rate of swing of our legs regarded as pendulums, so in pastimes impressions succeed each other at the rate at which attention can be most easily accommodated, and are such that we attend willingly.[71] We are absorbed in the present without being unwillingly confined to it; not only is there no motive for retrospect or expectation, but there is no feeling that the present endures. Each impression lasts as long as it is interesting, but does not continue to monopolize the focus of consciousness till attention to it is fatiguing, because uninteresting. In such facts, then, we seem to have proof that our perception of duration rests ultimately upon quasi-motor acts of varying intensity, the duration of which we do not directly experience as duration at all. They do endure and their intensity is a function of their duration; but the intensity is all that we directly perceive. In other words, it is here contended that what Locke called an instant or moment—“the time of one idea in our minds without the succession of another, of one wherein therefore we perceive no succession at all”—is psychologically not “a part in duration” in that sense in which, as he says, “we cannot conceive any duration without succession” (Essay, ii. 16, 12).

But, if our experience of time depends primarily upon acts of attention to a succession of distinct objects, it would seem that Is time Discrete or Continuous? time, subjectively regarded, must be discrete and not continuous. This, which is the view steadily maintained by the psychologists of Herbart's school, was implied if not stated by Locke, Berkeley and Hume. Locke hopelessly confuses time as perceived and time as conceived, and can only save himself from pressing objections by the retort, “It is very common to observe intelligible discourses spoiled by too much subtlety in nice divisions.” But Berkeley and Hume, with the mathematical discoveries of Newton and Leibnitz before them, could only protest that there was nothing answering to mathematical continuity in our experience. And, whereas Locke had tried to combine with his general psychological account the inconsistent position that “none of the distinct ideas we have of either [space or time] is without all manner of composition,” Berkeley declares, “For my own part, whenever I attempt to frame a simple idea of time, abstracted from the succession of ideas in my mind, which flows uniformly and is participated by all beings, I am lost and embrangled in inextricable difficulties. I have no notion of it at all, only I hear others say it is infinitely divisible, and speak of it in such a manner as leads me to harbour odd thoughts of my existence. . . . Time therefore being nothing, abstracted from the succession of ideas in our minds, it follows that the duration of any finite spirit must be estimated by the number of ideas or actions succeeding each other in that same spirit or mind” (Principles of Knowledge, i. § 98), Hume, again, is at still greater pains to show that “the idea which we form of any finite quality is not infinitely divisible, but that by proper distinctions and separations we may run this idea up to inferior ones, which will be perfectly simple and indivisible . . . that the imagination reaches a minimum, and may raise up to itself an idea of which it cannot conceive any subdivision, and which cannot be diminished without a total annihilation” (Human Nature, pt. ii. § 1, Green's ed., pp. 334 seq.).

At first blush we are perhaps disposed to accept this account of our time-perception, as Wundt, e.g. does, and to regard the attribution of continuity as wholly the result of after-reflection.[72] But it may be doubted if this is really an exact analysis of the case. Granted that the impressions to which we chiefly attend are distinct and discontinuous in their occupation of the focus of consciousness, and that, so far, the most vivid element in our time-experience is discrete; granted further that in recollection and expectation such objects are still distinct—all which seems to imply that time is a mere plurality—yet there is more behind. The whole field of consciousness is not occupied by distinct objects, neither are the changes in this field discontinuous. The experimental facts above-mentioned illustrate the transition from a succession the members of which are distinctly attended to to one in which they are indistinctly attended to, i.e. are not discontinuous enough to be separately distinguished. Attention does not move by hops from one definite spot to another, but, as Wundt himself allows, by alternate diffusion and concentration, like the foot of a snail, which never leaves the surface it is traversing. We have a clear presentation discerned as A or B when attention is gathered up; and, when attention spreads out, we have confused presentations not admitting of recognition. But, though not recognizable, such confused presentations are represented, and so serve to bridge over the comparatively empty interval during which attention is unfocused. Thus our perception of a period of time is not comparable to so many terms in a series of finite units any more than it is to a series of infinitesimals. When attention is concentrated in expectation of some single impression, then, no doubt, it is brought to a very fine point (“zugespitzt,” as Herbart would say); and a succession of such impressions would be represented as relatively discrete compared with the representation of the scenery of a day-dream. But absolutely discrete it is not and cannot be. In this respect the truth is rather with Herbert Spencer, who, treating of this subject from another point of view, remarks, “When the facts are contemplated objectively, it becomes manifest that, though the changes constituting intelligence approach to a single succession, they do not absolutely form one” (Psychology, i. § 180).

On the whole, then, we may conclude that our concrete time-experiences are due to the simultaneous representation of a series of definite presentations both accompanied and separated by more or fewer indefinite presentations more or less confused; that, further, the definite presentations have certain marks or temporal signs due to the movements of attention; that the rate of these movements or accommodations is approximately constant; and that each movement itself is primarily experienced as an intensity.

Experimental Investigations concerning Memory and Association.

30. Of the vast mass of experimental work undertaken in recent years, that relating to memory and association is probably the most important. A brief account of some of it is therefore offered at this point, by way of illustrating the character of the “new psychology.”

The learning and retaining of a stanza of poetry, say, is obviously a function of many variables, such as the mode of presentation (whether the words are heard only, or heard and seen, or both heard, seen and spoken aloud), the length, familiarity with the words and ideas used, the number of repetitions, the attention given, &c. Familiarity of course implies previous learning and retaining; the first essential, therefore, in any attempt to study these processes from the beginning, is the exclusion of this factor. Accordingly Ebbinghaus, the pioneer in experiments of this kind,[73] devised the new material, which is now regularly employed, namely, closed monosyllables, not themselves words, and strung together promiscuously into lines of fixed length so as never to form words: bam, rit, por, sig, nef, gud, &c., is an instance of such “senseless verses.” With very slight attention most persons would be able to reproduce three or four such syllables on a single reading or hearing; and by greater concentration six or seven might be so reproduced. This maximum, called sometimes the “span of pretension,” has been repeatedly made the subject of special inquiry. In idiots it is found, as might be expected, remarkably low; in school children it increases rapidly between the ages of eight and fourteen, and then remains almost stationary, individual differences being small compared with the striking differences that appear when longer lines make repetitions necessary.[74] This comparatively constant span of prehension is doubtless closely connected with certain other psychical constants, such as the duration of the psychical present and of the primary memory-image, the tempo of movements of attention (§§ 28, 29), &c. There are isolated investigations of these several conditions, but the subject as a whole still awaits systematic treatment.[75] That it is not wanting in interest is evident when we consider that if our span of prehension were enlarged, a corresponding increase in the variety and range of metre and rhyme in poetry, of “phrase” in music, and of evolution in the dance would be possible. The limits at present imposed on these and like complexities find their ultimate explanation in the constants just mentioned.

With lines of greater length than seven syllables some repetition is requisite before they can be said correctly: the number of such repetitions was found by Ebbinghaus to increase very rapidly with the number of syllables to be learnt. In his own case, for lines of 12, 16, 24, 36 syllables the repetitions necessary were on the average 16.6, 30, 44, 55 respectively. Thus for a line exceeding in length that of the span of prehension only about five times, he required fifty-five times as many repetitions, if we may call the single presentation of the syllables a “repetition.” Substituting poetry for gibberish of equal amount, Ebbinghaus found that one-tenth the number of repetitions sufficed; the enormous saving thus effected showing how numerous and intimate are the ready-made associations that “rhyme and reason” involve. But at one and the same time to memorize five verses even of sense requires more than five times as many repetitions as the memorizing of one. Two or three lines of inquiry here present themselves, e.g. (1) as to the comparative value of successive repetitions when several are taken together; (2) as to retention after an interval, as (a) a function of the number of repetitions previously made, and as (b) a function of the time; (3) as to the respective effects of more or less cumulating, or more or less distributing, the repetitions, on the number of these required.

1. It is at once obvious that beyond a certain point exhaustion of attention renders further repetition for a time futile; thus Ebbinghaus found 64 repetitions at one sitting of six 16-syllable nonsense verses, a task lasting some three-quarters of an hour, “was apt to bring on asthenia, a sort of epileptic aura, and the like!” But keeping well within this heroic limit, a certain “law of diminishing return,” to use an economic analogy, discloses itself. Thus taking a line of 10 syllables, the number of syllables reproduced correctly and in their proper order, after 1, 3, 6, 9 and 12 “repetitions,” were 2.2, 2.5, 2.8, 3.4, 3.9 respectively, as the averages of a series of experiments with each of eight persons.[76] “The first repetition is undoubtedly the best,” assuming, of course, that the subjects start with their attention fully concentrated. Some persons naturally do this, many do not; the experimenter has therefore to take special precautions to secure uniformity in this respect.

2. (a) On relearning a line after an interval of twenty-four hours there was in Ebbinghaus's case an average saving of one repetition for every three made the day before. A line of 16 syllables, for example, required some 30 repetitions, and could then be said off correctly. If only 8 repetitions were taken at first, the line being “under learnt,” it probably appeared quite strange the next day, yet the proportional saving was no less; on the other hand, if an additional 30 repetitions followed immediately on the first, the line being “doubly learnt,” in spite of the familiarity next day apparent, the proportional saving was no greater. The absolute saving would, of course, be less. We are so far led to infer that the stronger associations effected by many repetitions at one time fall off more rapidly than weaker associations effected by fewer repetitions in the same way. Herbart in his “psychical dynamics”—influenced probably by physical analogies—conjectured that the “sinking” or “inhibition” of presentations generally was proportional to their intensity: the less there was to sink, the slower the sinking became. Recent experiments certainly point in this direction. (b) As to retention as a function of the time—we all know that memories fade with time, but not at what precise rate. Ebbinghaus, by a series of prolonged experiments, ascertained the rate to be proportional to the logarithm of the time—a result already implied in that connecting retention and intensity; albeit in inquiries of this kind independent confirmation is always of value.

3. Had the proportional saving just described held good indefinitely, some 100 repetitions of the 16 syllables at one time should have dispensed with any further repetition twenty-four hours afterwards; whereas, in fact, this result seemed never attainable. Beyond a certain degree of accumulation, an ever-diminishing return was manifest, and that apparently short of the stage at which exhaustion of attention began to be felt. But, contrariwise, when the repetitions were distributed over several days, an ever-increasing efficiency was then the result. Thus, for Ebbinghaus, 38 repetitions spread over three days were as effective as 68 taken together. The results of careful experiments by Jost with two different subjects, using G. E. Müller's “method of telling” (to be described later on), are still more conclusive. Comparing 8 repetitions on three successive days with 4 repetitions on six, and 2 on twelve, the efficiencies, tested twenty-four hours later, were respectively as 11.5, 35, and 54; and probably, as Jost surmises, the effect of the maximum distribution—single “repetition” on twenty-four successive days—would have been more advantageous still, securing in fact the superiority of a first impression (cf. 1, above) on every occasion. This result again, is in part explained by the law of sinking already found. For if the sinking were simply proportional to the time, or were independent of the intensity, there would so far be no reason why one mode of distributing a given number of repetitions should be more economical than another. There is, however, another reason for this superiority, less clearly implied, to which we shall come presently.

Invariably, and almost of necessity, a more or less complex rhythmical articulation becomes apparent as the syllables are repeated, even when—as in the improved methods of G. E. Müller and his collaborateurs—they are presented singly and at regular intervals. A series of twelve syllables, for example, would be connected into six trochees, with a caesura in the middle of the verse; while in each half of it the first and last accented syllables would be specially emphasized; thus:

bām fĭs | lūp tŏl | gēn kĕr || dūb năf | &c.

In trying to suppress this tendency and to repeat the syllables in a monotonous, staccato fashion, just as they were presented, the tempo, though really unchanged, seemed to be distinctly quickened, a consequence, doubtless, of the greater effort involved. Moreover, the attempt, which was seldom successful, about doubled the number of repetitions required for learning off, thereby showing how much is gained by this psychical organization of disconnected material. But the gain thus ensured was manifest in other ways. Each foot, Whether dissyllabic or trisyllabic, became a new complex unit, the elements to be connected by successive association being thereby reduced to a half or a third, and the whole line seemingly shortened. The varied intonation, again, helped to fix the place of each foot in the verse, thus further facilitating the mind's survey of the whole. Such a transformation can hardly be accounted for so long as retention and association are regarded as merely mechanical and passive processes.

Psychical rhythm, upon which we here touch, has also been experimentally investigated at great length, alike in its physiological psychological and aesthetical aspects. The topic is far too intricate and unsettled for discussion here, yet two or three points may be noted in passing. We are not specially concerned with objective rhythms, recurring series of impressions—that is to say, in which there are actually periodic variations of intensity, interval and the like. What is remarkable is that even a perfectly regular succession of sounds (or touches), qualitatively and quantitatively all alike, a series therefore devoid of all objective rhythm, is nevertheless apprehended as rhythmically grouped, provided the rate lies between the limits of about 0·8″ and 0·14″. The slower of these rates leads to simple groups of two, replaced by groups of four or eight as the rate increases; groups of three and six also occur, though less frequently. The average duration of the groups, whether these are large or small, is comparatively constant, measuring rather more than one second. The subject usually keeps time by taps, nods or other accompanying movements; the pulse and respiration are also implicated. These organic rhythms have even been regarded as the prime source of all psychical rhythm and of its manifold aesthetic effects. Some connexion there is unquestionably. As the decimal system corresponds to our possession of ten fingers, and our movements to the structure of our limbs, so here we may assume that physiological processes fix the limits within which psychical rhythm is possible, but yet may be as little an adequate cause of it or its developments as fingers are of arithmetic, or legs of an Irish jig. In motor rhythms, such as the last, the initiative is obviously psychical, and the respiratory and other periodic organic processes simply follow suit. And even sensory rhythms can often be varied at the subject’s own choice, or on the suggestion of another; and then again the breathing is altered in consequence. Familiar instances of such procedure are to be found in the “tunes” so readily attributed to the puff of a locomotive, to the churning of a steamer’s screw, and the like. Psychical rhythm, then, we may conclude, is due to attention or apperception, but the conditions determining it are many, and their relations very complex. If the presentations to be “rhythmized” (the rhythmizomenon, as the Germans say) succeed each other slowly, the length (or shall we say the breadth?) of the “psychical present” tells one way: the first impression is below the threshold when the third appears. If they arrive rapidly, their intensity and duration and the span of pretension tell another way; for it is essential that they retain their individual distinctness and only so many can be grasped at once. But if the series continue long enough, or be frequently experienced, sub-groups may be treated as individuals; and indeed till some facility is acquired, the subject attending is aware of no rhythm. In the act of attention itself there are phases, in so far as expectation involves readjustment to what is coming: usually the first members of a tact are predominant, and the rhythm tends to “fall”; several alternations of accent within a complex rhythmic whole are of course still compatible with this. But it is important to note that, whether simple or complex, the rhythm is an intuited unity as truly as a geometrical figure may be. Unlike a geometrical figure, however, it rarely or never has symmetry. We cannot reverse a tune and obtain an effect comparable with that obtained by reprinting the score backwards in line with the original. We now pass to a question in which the psychological bearing of this fact becomes apparent.[77]

But first a new method of dealing with memory-problems must be mentioned, in which the connexion between rhythmizing and memorizing has been turned to account by the Göttingen psychologists. The method of Ebbinghaus consisted in ascertaining the repetitions saved in consequence of previous repetitions, when the verse was relearnt some fixed time later. Hence this method is called the learning method or the method of saving. When, a given time after a certain number of repetitions (say) in trochaic measure, the subject is confronted with one of the accented syllables and asked to name the unaccented syllable that belongs to it, he will answer sometimes rightly, sometimes wrongly, and sometimes be unable to answer at all. This, the new, method is therefore named Treffer-methode, the method of “shots,” or, let us say, the telling method. It enables the experimenter to obtain far more insight into details than was possible before, for the “misses” as well as the “hits” are instructive. Moreover, by measuring the time of each answer (Trefferzeit) and comparing these times together, much can be learnt; in stronger or recent associations, for example, the answers being quicker than in weaker or older ones.

Does association work forwards only or backwards also, as the middle link of a chain, when lifted, raises the contiguous links on either side of it? This is certainly not the case when the forward direction makes sense, but with nonsense verses, if the mechanical analogy is a sound one, such reversal is to be expected. For here there are none of the “obstructing associations” which “rhyme and reason” imply. In learning a verse backwards Ebbinghaus found a saving of 12·4% of the time originally taken up in learning it forwards. A saving almost as great (10·4%) was effected by relearning a like verse forwards, but skipping one syllable: the order of syllables, that is to say, being 1, 3, 5, . . . 15, 2, 4, . . . 16. Even when learning backwards and skipping one syllable, Ebbinghaus found a saving of 5%. But the number of his experiments (four) was too few to give this result much value, as he fully admits. These experiments as a whole, then, might incline us to suppose that association does work in both directions, though the connexions backwards are considerably weaker. But if so the associations both ways should be alike at least in form—continuous, that is to say, backwards, d c b a, as well as forwards, a b c d. The facts at present available are, however, against this. In two or three hundred experiments by Müller and Pilzecker, verses of twelve syllables were repeated a set number of times in anapestic measure—accented, that is to say, on the 3rd, 6th, 9th and 12th. After a fixed interval the subject, confronted with one of the accented syllables, mentioned any of the other syllables which he called to mind. Now the cases in which the syllable immediately preceding was revived were only about half as frequent as those in which the syllable next but one preceding was revived; the time of telling (Trefferzeit) for the latter was also shorter. This result is incompatible with the theory of continuous backward association, but it is readily explained by the fact that the group of three syllables had become one complex whole, and it shows that the tendency to reinstate the initial member of the group is stronger than that to reinstate the middle. The saving effected in Ebbinghaus’s experiment is also thus explained.[78]

A somewhat paradoxical situation is brought to light when the method of saving and the method of telling are used together. In the experiments by ]ost, mentioned above, the series of verses were repeated thirty times; after an interval of twenty-four hours one series was tested by the first method and the other by the second. Two new series were then taken: the first repeated four times, and after an interval of a minute tested by the first method; the other was then repeated in like manner, and tested after the same interval by the second method. The old series was found (by the method of saving) to require on an average 5·85 repetitions for relearning, and the new 9·6; yet on the method of telling, the new series yielded 2·7 “hits,” with an average time of about 13/4 second for each, while the old yielded only ·9 “hits,” with an average time of 41/2 seconds for each. Thus one may be able to reproduce relatively little of a given subject-matter, and yet require only a few repetitions in order to learn it off anew; on the other hand, one may know relatively much, and still find many more repetitions requisite for such complete learning. The “age” of the associations is then important. Other things being equal, we may conclude that each fresh repetition effects more for old associations than for recent ones. It might be supposed that the strength of the old associations was more uniform and on the average greater than the strength of the new; so that while none of the old were far below the threshold, few, if any, were above it; whereas more of the new might be above the threshold though the majority had lapsed entirely. And the latter would certainly be the case if the subject of experiment tried to make sure of a few “hits,” and paid no attention to the rest of the series. Due care was, however, taken that the ends of the experiment should not in this way be defeated. Also, there is ample evidence to show that the supposed greater uniformity in strength of old associations is not, in fact, the rule. We seem left, then, to conjecture that the difference is the effect of the process of assimilation working subconsciously—that psychical aspect of nervous growth which Professor James has aptly characterized by saying that “we learn to skate in summer and to swim in winter.” It continually happens that we can recognize connexions that we are quite unable to reproduce. To the diminished “strength” of an association, as tested by the method of telling, there may then quite well be an equivalent set-off in more developed assimilation. As a seed germinates it has less latent energy, but this is replaced by growth in root and stem: similar relations may obtain when an old association is said merely to lose “strength.” On the other hand—within the range of the primary memory-image—we can often reproduce what after a longer interval we should fail to recognize. We seem warranted, then, in concluding that this conception of “association-strength,” so freely used by G. E. Müller and his co-workers, requires more analysis than it has yet received. The two factors which their methods disclose in it appear to confirm the distinction we have already made between impressions and free ideas. They help us also to understand, further, the superiority of distributed over cumulated repetition, of “inwardly digesting” over “cram.”

Feeling.

31. Such summary survey as these limits allow of the more elementary facts of cognition is here at an end; so far the most conspicuous factors at work have been those of what might be termed the ideational mechanism. In the higher processes of thought we have to take more account of mental activity and of the part played by language. But it seems preferable, before entering upon this, to explore also the emotional and active constituents of mind in their more elementary phases.

In our preliminary survey we have seen that psychical life consists in the main of a continuous alternation of predominantly receptive and predominantly reactive consciousness. In its earliest form experience is simply an interplay of alternations of sensation and movement. At a later stage we find that in the receptive phase ideation is added to sensation; and that in the active phase thought and fancy, or the voluntary manipulation and control of the ideational trains, are added to the voluntary manipulation and control of the muscles. At this higher level also it is possible that either form of receptive consciousness may lead to either form of active: sensations may lead to thought rather than to action in the restricted sense, and ideas apart from sensations may prompt to muscular exertion. There is a further complication still: not only may either sensations or ideas lead to either muscular or mental movements, but movements themselves, whether of mind or limb, may as mere presentations determine other movements of either kind. In this respect, however, movements and thoughts either in themselves or through their sensational and ideational accompaniments may be regarded as pertaining to the receptive side of consciousness. With these provisos, then, the, broad generalization may hold that receptive states lead through feeling to active states, and that presentations that give neither pleasure nor pain meet with no responsive action. But first the objection must be met that presentations that are in themselves purely indifferent lead continually to very energetic action, often the promptest and most definite action. To this there are two answers. First, on the higher levels of psychical life presentations in themselves indifferent are often indirectly interesting as signs of, or as means to, other presentations that are more directly interesting. It is enough for the present, therefore, if it be admitted that all such indifferent presentations are without effect as often as they are not instrumental in furthering the realization of some desirable end. Secondly, a large class of movements, such as those called sensori-motor and ideo-motor, are initiated by presentations that are frequently, it must be allowed, neither pleasurable nor painful. In all such cases, however, there is probably only an apparent exception to the principle of subjective selection. They may all be regarded as instances of another important psychological principle which we shall have to deal with more fully by and by, viz. that voluntary actions, and especially those that either only avert pain or are merely subsidiary to pleasure giving actions, tend at length, as the effect of habit in the individual and of heredity in the race, to become “secondarily automatic,” as it has been called. Such mechanical or instinctive dexterities make possible a more efficient use of present energies in securing pleasurable and interesting experiences, and, like the rings of former growths in a tree, afford a basis for further advance, as old interests pall and new ones present themselves. Here, again, it suffices for our present purpose if it be granted that there is a fair presumption in favour of supposing all such movements to have been originally initiated by feeling, as certainly very many of them were.

Of the feeling itself that intervenes between these sensory and motor presentations there is but little to be said. The chief points have been already insisted upon, viz. that it is not itself a presentation, but a purely subjective state, at once the effect of a change in receptive consciousness and the cause of a change in motor consciousness; hence its continual confusion either with the movements, whether ideational or muscular, that are its expression, or with the sensations or ideas that are its cause. For feeling as such is, so to put it, matter of being rather than of direct knowledge; and all that we know about it we know from its antecedents or consequents in presentation.

Pure feeling, then, ranging solely between the opposite extremes of pleasure and pain, we are naturally led to inquire Causes of Feeling. whether there is any corresponding contrast in the causes of feeling on the one hand, and on the other in its manifestations and effects. To begin with the first question, which we may thus formulate: What, if any, are the invariable differences characteristic of the presentations or states of mind we respectively like and dislike; or, taking account of the diverse sources of feeling—sensuous, aesthetic, intellectual, active—is there anything that we can predicate alike of all that are pleasurable and deny of all that are painful, and vice versa? It is at once evident that at least in presentations objectively regarded no such common characters will be found; if we find them anywhere it must be in some relation to the conscious subject i.e. in the fact of presentation itself. There is one important truth concerning pleasures and pains that may occur at once as an answer to our inquiry, and that is often advanced as such, viz. that whatever is pleasurable tends to further and perfect life, and whatever is painful to disturb or destroy it. The many seeming exceptions to this law of self-conservation, as it has been called, probably all admit of explanation in conformity with it, so as to leave its substantial truth unimpeached.[79] But this law, however stated, is too teleological to serve as a purely psychological principle, and, as generally formulated and illustrated, it takes account of matters quite outside the psychologist's ken. We are not now concerned to know why a bitter taste e.g. is painful or the gratification of an appetite pleasant, but what marks distinctive of all painful presentations the one has and the other lacks. From a biological standpoint it may be true enough that the final cause of sexual and parental feelings is the perpetuation of the species; but this does not help us to ascertain what common character they have as actual sources of feeling for the individual. From the biological standpoint again, even the senile decadence and death of the individual may be shown to be advantageous to the race; but it would certainly be odd to describe this as advantageous to the individual; so different are the two points of view. What we are in search of, although a generalization, has reference to something much more concrete than concepts like race or life, and does not require us to go beyond the consciousness of the moment to such ulterior facts as they imply.

Were it possible it would be quite unnecessary to examine in detail every variety of pleasurable and painful consciousness in connexion with a general inquiry of this sort. It will be best to enumerate at the outset the only cases that specially call for investigation. Feeling may arise mainly from (a) single sensations or movements, including in these what recent psychologists call their tone; or it may be chiefly determined by (b) some combination or arrangement of these primary presentations—hence what might be styled the lower aesthetic feelings. We have thus among primary presentations a more material and a more formal cause or ground of feeling. The mere representation of these sources of feeling involves nothing of moment: the idea of a bright colour or a bitter taste has not definiteness or intensity enough to produce feeling; and the ideal presentation of a harmonious arrangement of sounds or colours does not in itself differ essentially as regards the feeling it occasions from the actual presentation. When we advance to the level at which there occur ideas more complex and more highly representative—or re-representative, as Mr Spencer would say—than any we have yet considered we can again distinguish between material and formal grounds of feeling. To the first we might refer, e.g. (c) the egoistic, sympathetic, and religious feelings; this class will probably require but brief notice. The second, consisting of (d) the intellectual and (e) the higher aesthetic feelings, is psychologically more important. There is a special class of feelings, which might be distinguished from all the preceding as reflex, since they arise from the memory or expectation of feelings but in fact these are largely involved in all the higher feelings, and this brief reference to them will suffice: of such hope, fear, regret are examples.

a. The quality and intensity as well as the duration and frequency of a sensation or movement all have to do with Sensations and Movements. determining to what feeling it gives rise. It will be best to leave the last two out of account for a time. Apart from these, the pleasantness or painfulness of a movement appears to depend solely upon its intensity, that is to say, upon the amount of effort necessary to effect it, in such wise that a certain amount of exertion is agreeable and any excess disagreeable. Some sensations also, such as those of light and sound, are agreeable if not too intense, their pleasantness increasing with their intensity up to a certain point, on nearing which the feeling rapidly changes and becomes disagreeable or even painful. Other sensations, as bitter tastes, e.g. are naturally unpleasant, however faint—though we must allow the possibility of an acquired liking for moderately bitter or pungent flavours. But in every case such sensations produce unmistakable manifestations of disgust, if at all intense. Sweet tastes, on the other hand, however intense, are pleasant to an unspoiled palate, though apt before long to become mawkish, like “sweetest honey, loathsome in his own deliciousness,” as confectioners' apprentices are said soon to find. The painfulness of all painful sensations or movements increases with their intensity without any assignable maximum being reached.

A comparison of examples of this kind, which it would be tedious to describe more fully and which are indeed too familiar to need much description, seems to show (1) that, so far as feeling is determined by the intensity of a presentation, there is pleasure so long as attention can be adapted or accommodated to the presentation, and pain so soon as the intensity is too great for this; and (2) that, so far as feeling is determined by the quality of a presentation, those that are pleasurable enlarge the field of consciousness and introduce or agreeably increase in intensity certain organic sensations, while those that are painful contract the field of consciousness and introduce or disagreeably increase in intensity certain organic sensations. There are certain other hedonic effects due to quality, the examination of which we must for the present defer. Meanwhile as to the first point it may be suggested, as at any rate a working hypothesis, that in itself any and every simple sensation or movement is pleasurable if there is attention forthcoming adequate to its intensity. In the earliest and simplest phases of life, in which the presentation-continuum is but little differentiated, it is reasonable to suppose that variation in the intensity of presentation preponderates over changes in the quality of presentation, and that to the same extent feeling is determined by the former and not by the latter. And, whereas this dependence on intensity is invariable, there is no ground for supposing the quality of any primary presentation, when not of excessive intensity, to be invariably disagreeable; the changes above-mentioned in the hedonic effects of bitter tastes, sweet tastes, or the like tend rather to prove the contrary. This brings us to the second point, and it requires some elucidation. We need here to call to mind the continuity of our presentations and especially the existence of a background of organic sensations or somatic consciousness, as it is variously termed. By the time that qualitatively distinct presentations have been differentiated from this common basis it becomes possible for any of these, without having the intensity requisite to affect feeling directly, to change it indirectly by means of the systemic sensations accompanying them, or, in other words, by their tone. The physiological concomitants of these changes of somatic tone are largely reflex movements or equivalents of movements, such as alterations in circulatory, respiratory and excretory processes. Such movements are psychologically movements no longer, and are rightly regarded as pertaining wholly to the sensory division of presentations. But originally it may have been otherwise. To us now, these organic reflexes seem but part and parcel of the special sensation whose tone they form, and which they accompany even when that sensation, so far as its mere intensity goes, might be deemed indifferent. But perhaps at first the special qualities that are now throughout unpleasant may have been always presented with an excessive intensity that would be painful on this score alone, and the reflexes that at present pertain to them may then have been psychologically the expression of this pain.[80] At any rate it is manifestly unfair to refuse either to seek out the primitive effects of the sensations in question and allow for the workings of heredity, or to reckon this accompanying systemic feeling as part of them. The latter seems the readier and perhaps, too, the preferable course. A word will now suffice to explain what is meant by enlarging and contracting the field of consciousness and agreeably increasing or decreasing certain elements therein.

The difference in point is manifest on comparing the flow of spirits, buoyancy and animation which result from a certain duration of pleasurable sensations with the lowness or depression of spirits, the gloom and heaviness of heart, apt to ensue from prolonged physical pain. Common language, in fact, leaves us no choice but to describe these contrasted states by figures which clearly imply that they differ in the range and variety of the presentations that make up consciousness, and in the quickness with which these succeed each other.[81] It is not merely that in hilarity as contrasted with dejection the train of ideas takes a wider sweep and shows greater liveliness, but as it were at the back of this, on the lower level of purely sensory experience, certain organic sensations which are ordinarily indifferent acquire a gentle intensity, which seems by flowing over to quicken and expand the ideational stream as we see, for instance, in the effects of mountain air and sunshine. Or, on the other hand, these sensations become so violently intense as to drain off and ingulf all available energy in one monotonous corroding care, an oppressive weight which leaves no place for free movement, no life or leisure to respond to what are wont to be pleasurable solicitations?[82]

As regards the duration and the frequency of presentation, it is in general true that the hedonic effect soon attains its maximum, and then, if pleasant, rapidly declines, or even changes to its opposite. Pains in like manner decline, but more slowly and without in the same sense changing to pleasures. The like holds of too frequent repetition. Physiological explanation of these facts, good as far as it goes, is, of course, at once forthcoming: sensibility is blunted, time is required for restoration, and so forth; but at least we want the psychological equivalent of all this. In one respect we find nothing materially new; so far as continued presentation entails diminished intensity we have nothing but diminished feeling as a consequence; so far as its continued presentation entails satiety the train of agreeable accompaniments ceases in which the pleasurable tone consisted. But in another way long duration and frequent repetition produce indirectly certain characteristic effects on feeling in consequence of habituation and accommodation. We may get used to a painful presentation in such wise that we cease to be conscious of it as positively disagreeable, though its cessation is at once a source of pleasure; in like manner we come to require things simply because it is painful to be without them, although their possession has long ceased to be a ground of positive enjoyment. This loss (or gain) consequent on accommodation[83] has a most important effect in changing the sources of feeling: it helps to transfer attention from mere sensations to what we may distinguish as interests.

b. Certain sensations or movements not separately unpleasant become so when presented together or in immediate succession; Combinations of Sensations and of Movements. and contrariwise, some combinations of sensations or of movements may be such as to afford pleasure distinct from, and often greater than, any that they separately yield. Here again we find that in some cases the effect seems mainly to depend on intensity, in others mainly on quality. (i.) As instances of the former may be mentioned the pleasurableness of a rhythmic succession of sounds or movements, of symmetrical forms and curved outlines, of gentle crescendos and diminuendos in sound, and of gradual variations of shade in colour, and the painfulness of flickering lights, “beats” in musical notes, false time, false steps, false quantities, and the like. In all these, whenever the result is pleasurable, attention can be readily accommodated—is, so to say, economically meted out; and, whenever the result is painful, attention is surprised, balked, wasted. Thus we can make more movements and with less expenditure of energy when they are rhythmic than when they are not, as the performances of a ball-room or of troops marching to music amply testify. Of this economy we have also a striking proof in the ease with which rhythmic language is retained. (ii.) As instances of the latter may be cited those arrangements of musical tones and of colours that are called harmonious or the opposite. Harmony, however, must be taken to have a different meaning in the two cases. When two or three tones harmonize there results, as is well known, a distinct pleasure over and above any pleasure due to the tones themselves. On the other hand, tones that are discordant are unpleasant in spite of any pleasantness they may have singly. Besides the negative condition of absence of beats, a musical interval to be pleasant must fulfil certain positive conditions, sufficiently expressed for our purpose by saying that two tones are pleasant when they give rise to few combination-tones, and when among these there are several that coincide, and that they are unpleasant when they give rise to many combination-tones, and when among these there are few or none that coincide. Too many tones together prevent any from being distinct. But where tones coincide the number of tones actually present is less than the number of possible tones, and there is a proportionate simplification, so to put it: more is commanded and with less effort. An ingenious writer[84] on harmony, in fact, compares the confusion of a discord to that of “trying to reckon up a sum in one's head and failing because the numbers are too high.” A different explanation must be given of the so-called harmonies of colour. The pleasurable effect of graduations of colour or shade—to which, as Ruskin tells us, the rose owes its victorious beauty when compared with other flowers—has been already mentioned: it is rather a quantitative than a qualitative effect. What we are now concerned with are the pleasurable or painful combinations of different ungraduated colours. A comparison of these seems to justify the general statement that those colours yield good combinations that are far apart in the colour circle, while those near together are apt to be discordant. The explanation given, viz. that the one arrangement secures and the other prevents perfect retinal activity, seems on the whole satisfactory—especially if we acknowledge the tendency of all recent investigations and distinguish sensibility to colour and sensibility to mere light as both psychologically and physiologically two separate facts. Thus, when red and green are juxtaposed, the red increases the saturation of the green and the green that of the red, so that both colours are heightened in brilliance. But such an effect is only pleasing to the child and the savage; for civilized men the contrast is excessive, and colours less completely opposed, as red and blue, are preferred, each being a rest from the other, so that as the eye wanders to and fro over their border different elements are active by turns. Red and orange, again, are bad, in that both exhaust in a similar manner and leave the remaining factors out of play.

c. The more or less spontaneous workings of imagination, as well as that direct control of this working necessary to thinking Ideation and Intellection. in the stricter sense, are always productive of pain or pleasure in varying degrees. Though the exposition of the higher intellectual processes has not yet been reached, there will be no inconvenience in at once taking account of their effects on feeling, since these are fairly obvious and largely independent of any analysis of the processes themselves. It will also be convenient to include under the one term “intellectual feelings,” not only the feelings connected with certainty, doubt, perplexity, comprehension, and so forth, but also what the Herbartian psychologists—whose work in this department of psychology is classical—have called par excellence the formal feelings—that is to say, feelings which they regard as entirely determined by the form of the flow of ideas, and not by the ideas themselves. Thus, be the ideas what they may, when their onward movement is checked by divergent or obstructing lines of association, and especially when in this manner we are hindered, say, from recollecting a name or a quotation (as if, e.g. the names of Archimedes, Anaximenes and Anaximander each arrested the clear revival of the other), we are conscious of a certain strain and oppressiveness, which give way to momentary relief when at length what is wanted rises into distinct consciousness and our ideas resume their flow. Here again, too, as in muscular movements, we have the contrast of exertion and facility, when “thoughts refuse to flow” and we work “invita Minerva,” or when the appropriate ideas seem to unfold and display themselves before us like a vision before one inspired. To be confronted with propositions we cannot reconcile—i.e. with what is or appears inconsistent, false, contradictory—is apt to be painful; the recognition of truth or logical coherence, on the other hand, is pleasurable. The feeling in either case is, no doubt, greater the greater our interest in the subject-matter; but the mere conflict of ideas as such is in itself depressing, while the discernment of agreement, of the one in the many, is a distinct satisfaction. Now in the one case we are conscious of futile efforts to comprehend as one ideas which the more distinctly we apprehend them for the purpose only prove to be the more completely and diametrically opposed: we can only affirm and mentally envisage the one by denying and suppressing the representation of the other; and yet we have to strive to predicate both and to embody them together in the same mental image. Attention is like a house divided against itself: there is effort but it is not effective, for the field of consciousness is narrowed and the flow of ideas arrested. When, on the other hand, we discern a common principle among diverse and apparently disconnected particulars, instead of all the attention we can command being taxed in the separate apprehension of these “disjecta membra,” they become as one, and we seem at once to have at our disposal resources for the command of an enlarged field and the detection of new resemblances.

d. Closely related to these formal intellectual feelings are certain of the higher aesthetic feelings. A reference to some Higher Aesthetic Feelings. of the commonplaces of aesthetical writers may be sufficient briefly to exhibit the leading characteristics of these feelings. There is a wide agreement among men in general as to what is beautiful and what is not, and it is the business of a treatise on empirical aesthetics from an analysis of these matters of fact to generalize the principles of taste—to do, in fact, for one source of pleasure and pain what we are here attempting in a meagre fashion for all. And these principles are the more important in their bearing upon the larger psychological question, because among aesthetic effects are reckoned only such as are pleasing or otherwise in themselves, apart from all recognition of utility, of possession, or of ulterior gratification of any kind whatever. Thus, if it should be objected that the intellectual satisfaction of consistency is really due to its utility, to the fact that what is incompatible and incomprehensible is of no avail for practical guidance, at least this objection will not hold against the aesthetic principle of unity in variety. In accordance with this primary maxim of art criticism, at the one extreme art productions are condemned for monotony, as incapable of sustaining interest because “empty,” “bald” and “poor”; at the other extreme they are condemned as too incoherent and disconnected to furnish a centre of interest. And those are held as so far praiseworthy in which a variety of elements, be they movements, forms, colours or incidents, instead of conflicting, all unite to enhance each other and to form not merely a mass but a whole. Another principle that serves to throw light on our inquiry is that which has been called the principle of economy,[85] viz. that an effect is pleasing in proportion as it is attained by little effort and simple means. The brothers Weber in their classic work on human locomotion discovered that those movements that are aesthetically beautiful are also physiologically correct; grace and ease, in fact, are well-nigh synonymous, as Herbert Spencer points out, and illustrates by apt instances of graceful attitudes, motions and forms. The same writer,[86] again, in seeking for a more general law underlying the current maxims of writers on composition and rhetoric is led to a special formulation of this principle as applied to style, viz. that “economy of the recipient's attention is the secret of effect.”

Perhaps of all aesthetical principles the most wide-reaching, as well as practically the most important, is that which explains aesthetic effects by association. Thus, to take one example where so many are possible, the croaking of frogs and the monotonous ditty of the cuckoo owe their pleasantness, not directly to what they are in themselves, but entirely to their intimate association with spring-time and its gladness. At first it might seem, therefore, that in this principle there is nothing fresh that is relevant to our present inquiry, since a pleasure that is only due to association at once carries back the question to its sources; so that in asking why the spring, for example, is pleasant we should be returning to old ground. But this is not altogether true; aesthetic effects call up not merely ideas but ideals. A great work of art improves upon the real in two respects: it intensifies and it transfigures. It is for art to gather into one focus, cleared from dross and commonplace, the genial memories of a lifetime, the instinctive memories of a race; and, where theory can only classify and arrange what it receives, art—in a measure free from “the literal unities of time and place”—creates and glorifies. Still art eschews the abstract and speculative; however plastic in its hands, the material wrought is always that of sense. We have already noticed more than once the power which primary presentations have to sustain vivid re-presentations, and the bearing of this on the aesthetic effects of works of art must be straightway obvious. The notes and colours, rhymes and rhythms, forms and movements, which produce the lower aesthetic feelings also serve as the means of bringing into view, and maintaining at a higher level of vividness, a wider range and flow of pleasing ideas than we can ordinarily command.

When we reach the level at which there is distinct self-consciousness (cf. § 44), we have an important class of Egoistic and Socialistic Feelings. feelings determined by the relation of the presentation of self to the other contents of consciousness. And as the knowledge of other selves advances pari passu with that of one's own self, so along with the egoistic feelings appear certain social or altruistic feelings. The two have much in common; in pride and shame, for example, account is taken of the estimate other persons form of us and of our regard for them; while, on the other hand, when we admire or despise, congratulate or pity another, we have always present to our mind a more or less definite conception of self in like circumstances. It will therefore amply serve all the ends of our present inquiry if we briefly survey the leading characteristics of some contrasted egoistic feelings, such as self-complacency and disappointment. When a man is pleased with himself, his achievements, possessions or circumstances, such pleasure is the result of a comparison of his present position in this respect with some former position or with the position of someone else. Without descending to details, we may say that two prospects are before him, and the larger and fairer is recognized as his own. Under disappointment or reverse the same two pictures may be present to his mind, but accompanied by the certainty that the better is not his or is his no more. So far, then, it might be said the contents of his consciousness are in each case the same, the whole difference lying in the different relationship to self. But this makes all the difference even to the contents of his consciousness, as we shall at once see if we consider its active side. Even the idlest and most thoughtless mind teems with intentions and expectations, and in its prosperity, like the fool in the parable, thinks to pull down its barns and build greater, to take its ease, eat, drink and be merry. The support of all this pleasing show and these far-reaching aims is, not the bare knowledge of what abundance will do, but the reflection—These many goods are mine. In mind alone final causes have a place, and the end can produce the beginning; the prospect of a summer makes the present into spring. But action is paralysed or impossible when the means evade us. In so far as a man's life consists in the abundance of the things he possesseth, we see then why it dwindles with these. The like holds where self-complacency or displicency rests on a sense of personal worth or on the honour or affection of others.

32. We are now at the end of our survey of certain typical pleasurable and painful states. The answer to our inquiry Summary and Result. which it seems to suggest is that there is pleasure in proportion as a maximum of attention is effectively exercised, and pain in proportion as such effective attention is frustrated by distractions, shocks, or incomplete and faulty adaptations, or fails of exercise, owing to the narrowness of the field of consciousness and the slowness and smallness of its changes. Something must be said in explication of this formula, and certain objections that might be made to it must be considered. First of all it implies that feeling is determined partly by quantitative, or, as we might say, material conditions, and partly by conditions that are formal or qualitative. As regards the former, both the intensity or concentration of attention and its diffusion or the extent of the field of consciousness have to be taken into account. Attention, whatever else it is, is a limited quantity—

Pluribus intentus minor est ad singula sensus—

to quote Hamilton's pet adage. Moreover, as we have seen, attention requires time. If, then, attention be distributed over too wide a field, there is a corresponding loss of intensity, and so of distinctness: we tend towards a succession of indistinguishables—indistinguishable, therefore, from no succession. We must not have more presentations in the field of consciousness than will allow of some concentration of attention: a maximum diffusion will not do. A maximum concentration, in like manner—even if there were no other objection to it—would seem to conflict with the general conditions of consciousness, inasmuch as a single simple presentation, however intense, would admit of no differentiation, and any complex presentation is in some sort a plurality. The most effective attention, then, as regards its quantitative conditions, must lie somewhere between the two zeros of complete indifference and complete absorption. If there be an excess of diffusion, effective attention will increase up to a certain point as concentration increases, but beyond that point will decrease if this intensification continues to increase; and vice versa, if there be an excess of concentration. But, inasmuch as these quantitative conditions involve a plurality of distinguishable presentations or changes in consciousness, the way is open for formal conditions as well. Since different presentations consort differently when above the threshold of consciousness together, one field may be wider and yet as intense as another, or intenser and yet as wide, owing to a more advantageous arrangement of its constituents.[87]

The doctrine here developed, viz., that feeling depends on efficiency, is in the main as old as Aristotle; all that has been Negative Pleasures. done is to give it a more accurately psychological expression, and to free it from the implications of the faculty theory, in which form it was expounded by Hamilton. Of possible objections there are at least two that we must anticipate, and the consideration of which will help to make the general view clearer. First, it may be urged that, according to this view, it ought to be one continuous pain to fall asleep, since in this state consciousness is rapidly restricted both as to intensity and range. This statement is entirely true as regards the intensity and substantially true as regards the range, at least of the higher consciousness: certain massive and agreeable organic sensations pertain to falling asleep, but the variety of presentations at all events grows less. But then the capacity to attend is also rapidly declining; even a slight intruding sensation entails an acute sense of strain in one sense, in place of the massive pleasure of repose throughout; and any voluntary concentration either in order to move or to think involves a like organic conflict, futile effort, and arrest of balmy ease. There is as regards the more definite constituents of the field of consciousness a close resemblance between natural sleepiness and the state of monotonous humdrum we call tedium or ennui; and yet the very same excitement that would relieve the one by dissipating the weariness of inaction would disturb the other by renewing the weariness of action: the one is commensurate with the resources of the moment, the other is not. Thus the maximum of effective attention in question is, as Aristotle would say, a maximum “relative to us.” It is possible, therefore, that a change from a wider to a narrower field of consciousness may be a pleasurable change, if attention is more effectively engaged. Strictly speaking, however, the so-called negative pleasures of rest do not consist in a mere narrowing of the field of consciousness so much as in a change in the amount of concentration. Massive organic sensations connected with restoration take the place of the comparatively acute sensations of jaded powers forced to work. We have, then, in all cases to bear in mind this subjective relativity of all pleasurable or painful states of consciousness.

33. But there is still another and more serious difficulty to face. It has long been a burning question with theoretical Do Pleasures Differ Qualitatively? moralists whether pleasures differ only quantitatively or differ qualitatively as well, whether psychological analysis will justify the common distinction of higher and lower pleasures or force us to recognize nothing but differences of degree, of duration, and so forth—as expounded, e.g. by Bentham, whose cynical mot, “pushpin is as good as poetry provided it be as pleasant,” was long a stumbling block in the way of utilitarianism. The entire issue here is confused by an ambiguity in terms that has been already noticed: pleasure and pleasures have not the same connotation. By a pleasure or pleasures we mean some assignable presentation or presentations experienced as pleasant—i.e. as affording pleasure; by pleasure simply is meant this subjective state of feeling itself. The former, like other objects of knowledge, admit of classification and comparison: we may distinguish them as coarse or as noble, or, if we will, as cheap and wholesome. But while the causes of feeling are manifold, the feeling itself is a subjective state, varying only in intensity and duration. The best evidence of this lies in the general character of the actions that ensue through feeling—the matter which has next to engage us. Whatever be the variety in the sources of pleasure, whatever be the moral or conventional estimate of their worthiness, if a given state of consciousness is pleasant we seek so far to retain it, if painful to be rid of it: we prefer greater pleasure before less, less pain before greater. This is, in fact, the whole meaning of preference as a psychological term. Wisdom and folly each prefer the course which the other rejects. Both courses cannot, indeed, be objectively preferable; that, however, is not a matter for psychology. But as soon as reflection begins, exceptions to this primary principle of action seem to arise continually, even though we regard the individual as a law to himself. Such exceptions, however, we may presently find to be apparent only. At any rate the principle is obviously true before reflection begins—true so long as we are dealing with actually present sources of feeling, and not with their re-presentations. But to admit this is psychologically to admit everything, at least if experience is to be genetically explained. Assuming then that we start with only quantitative variations of feeling, we have to attempt to explain the development of formal and qualitative differences in the character given to the grounds of feeling. But, if aversions and pursuits result from incommensurable states of pain and pleasure, there seems no other way of saving the unity and continuity of the subject except by speculative assumption—the doctrine known as the freedom of the will in its extremest form. The one position involves the other, and the more scientific course is to avoid both as far as we can.

The question, then, is: How, if action depends in the last resort on a merely quantitative difference, could it ever come about that what we call the higher sources of feeling should supersede the lower? If it is only quantity that turns the scales, where does quality come in, for we cannot say, e.g. that the astronomer experiences a greater thrill of delight when a new planet rewards his search than the hungry savage in finding a clump of pig-nuts? Tempora mutantur nos et mutamur in illis contains the answer in brief. We shall understand this answer better if we look at a parallel case, or what is really our own from another point of view. We distinguish between higher and lower forms of life: we might say there is more life in a large oyster than in a small one, other things being equal, but we should regard a crab as possessing not necessarily more life—as measured by waste of tissue—but certainly as manifesting life in a higher form. How, in the evolution of the animal kingdom, do we suppose this advance to have been made? The tendency at any one moment is simply towards more life, simply towards growth; but this process of self-conservation imperceptibly but steadily modifies the self that is conserved. The creature is bent only on filling its skin; but in doing this as easily as may be it gets a better skin to fill, and accordingly seeks to fill it differently. Though cabbage and honey are what they were before, they have changed relatively to the grub now it has become a butterfly. So, while we are all along preferring a more pleasurable state of consciousness before a less, the content of our consciousness is continually changing; the greater pleasure still outweighs the less, but the pleasures to be weighed are either wholly different, or at least are the same for us no more. What we require then, is not that the higher pleasures shall always afford greater pleasure than the lower did, but that to advance to the level of life on which pleasure is derived from higher objects shall on the whole be more pleasurable and less painful than to remain behind. And this condition seems provided in the fact of accommodation above referred to and in the important fact that attention can be more effectively expended by what we may therefore call improvements in the form of the field of consciousness. But when all is said and done a certain repugnance is apt to arise against any association of the differences between the higher and lower feelings with differences of quantity. Yet such repugnance is but another outcome of the common mistake of supposing that the real is obtained by pulling to pieces rather than by building up. No logical analysis—nay, further, no logical synthesis—is adequate to the fullness of things. For the rest, such aversion is wholly emotional, and has no more an intellectual element in it than has the disgust we feel on first witnessing anatomical dissections.[88]

Emotion and Emotional Expression.

34. We now pass from the causes of feeling to its effects. We have assumed (§ 7) that the simplest and earliest of these Effects of Feeling. effects are to be found in the various bodily movements commonly described as the expression or manifestation of emotion. But in a notorious article, entitled “What is an Emotion?” Professor James[89] attempted to turn this, the common-sense position, upside down. Before proceeding we must, therefore, examine his alternative theory: “Common sense says: we lose our fortune, are sorry and weep; we meet a bear, are frightened and run; we are insulted by a rival, are angry and strike.” But, Professor James continues, “the hypothesis here to be defended says that this order of sequence is incorrect: that the one mental state is not immediately induced by the other, that the bodily manifestations must first be interposed between, and that the more rational statement is that we feel sorry because we cry, angry because we strike, afraid because we tremble, and not that we cry, strike or tremble because we are sorry, angry or fearful, as the case may be.” In a word, whereas it is commonly supposed that the emotion precedes and produces the expression, it seems here to be maintained that the expression precedes and produces the emotion. But the sequence denied in the first case is a psychological sequence, the sequence maintained in the second is a physiological sequence. The subject's experiences of the bodily expressions is here the emotion, and these are physically, not psychically, determined. “They are sensational processes,” says Professor James; “processes due to inward currents set up by physical happenings.”

The new theory is, then, in part psychological, in part psychophysical. As to the first part, which the author calls “the vital point of the whole theory,” it consists mainly in exposing the ambiguity of the phrase “bodily expression of an emotion”—a phrase which is liable to mislead us into fancying that emotion, like thought, may be antecedent to, or independent of, any expression or utterance. My fear or anger may chance to be expressive to another, but they are of necessity impressive to me. “A disembodied human emotion is a sheer nonentity.” In so far as I have a certain emotion, in so far I have “the feelings of its bodily symptoms.” This is true, not to say trite; but how do these symptoms arise? With this question we pass to the psychophysical side of the theory, and here it becomes perplexing, and is itself perplexed; for to this question it is driven to return two distinct and divergent answers. First, we are told that it is not the emotion that gives rise to the bodily expression, but that, on the contrary, “the bodily changes follow directly the perception of the existing fact,” it being beyond doubt “that objects do excite bodily changes by a reorganised mechanism.” Again: “Each emotion is,” for Professor James, “a resultant of a sum of elements, and each element is caused by a physiological process of a sort already well known. The elements are all organic changes, and each of them is the reflex effect of the existing object.” The old attempts at classification and description being contemptuously dismissed as belonging only to “the lowest stage of science,” we are informed that now we step from a superficial to a deep order of inquiry. “The questions now are causal: ‘Just what changes does this object and what changes does that object excite?’ and ‘How come they to excite these particular changes, and not others?’ ” But we have not had to wait for the James-Lange theory to raise these questions, and surely there are none that bring out its defects more glaringly. “Objects” that determine bodily changes by means of reorganized mechanism and without psychical interposition might fairly be taken to be physical objects; and indeed the whole process is expressly described as reflex. But only very slovenly physiologists talk of “objects” exciting reflexes: it is inexact even to say that sensations do so. All that reflex action requires is a stimulus. “The essence of a reflex action,” says Foster, “consists in the transmutation, by means of the irritable protoplasm of a nerve-cell, of afferent into efferent impulses.” Let Professor James be confronted first by a chained bear and next by a bear at large: to the one object he presents a bun, and to the other a clean pair of heels; or let him first be thrilled by a Beethoven symphony and then by a Raphael Madonna. Will he now undertake to account, in terms of stimuli and their reflex effects, for the very different results of the similar “causes” in the one case, or for the similar results of the very different “causes” in the other? Such a challenge would certainly be declined, and Professor James would remind us that in his nomenclature “it is the total situation on which the reaction of the subject is made.”[90] But there is just a world of difference between “object” = stimulus transformed by reorganised mechanism into an efferent discharge, and “object” = total situation to which the subject reacts. The attempt to explain emotion causally on the lines of the former meaning lands us in the conscious automaton theory, with which we must deal presently: this Professor James rejects. The latter meaning, on the other hand, involves the recognition of the subject's attitude as essential to the reaction, and of this as determined by pleasure, pain or by some “interest” resting ultimately on these. Such, with scarcely an exception, has always been, and still remains, the analysis of emotion in vogue among psychologists. It brings to the fore a new category, that of worth or value, one wholly extraneous to the physiologist's domain, and repugnant to the mechanical analogies which are there in place. No doubt such a concept is attained only by reflexion, but the experiences from which it is drawn, the affective states and the conative tendencies of the subject experiencing, must have preceded. From this central standpoint alone the objective situation has a worth which explains the subject's attitude, and here alone can we ind the clue which will enable us to answer the questions of cause that Professor James propounds.

The experimental investigations of Mosso, Féré, Lehmann, and others have shown that the vaso-motor and such like bodily changes as are prominent in emotional excitement are present also to some extent in all forms of conscious activity. The more unwonted and interesting the situation, the more diffused movements predominate over movements that are purposive; the further assimilation, both on the cognitive and the reactive side, has advanced, the more diffusion is replaced by restriction and adaptation. But we are not warranted in separating these factors of voluntary activity into distinct processes, as the physiologist, for example, separates the functions of striped and unstriped muscle. Unless we are prepared to treat all activity as reflex—as the physiologist may quite well do, if he keep strictly to his own point of view—it does not seem possible to regard emotional expression as so much organic sensation with which purposive movement has nothing to do. No doubt this connexion of vegetal and animal functions remains one of the obscurest in all psycho-biology, though its teleological fitness, is obvious enough.

Nevertheless, Professor James's main position is that an emotion is but a sum of organic sensations; and in order to establish this he is led to the second and very different statement which we have now to examine. Here, so far from suggesting inquiries as to the “objects” that excite emotion, his point is to maintain that in so far as the bodily cause is set up, be the means what they may, in so far the emotion is present.[91] And here, at length, the contention is explicit: Emotions are a certain complex of organic sensations, and such complexes are emotions: the two are not merely coexistent, they are identical. The exciting object is thus, after all, physiological; that is to say, it is whatever stimulus sets up the sensations. It cannot be psychological, “the total situation for the reacting subject,” for in this sense the emotion, it is maintained, may be “objectless.” In support of his position Professor James first of all cites pathological cases of such objectless emotion. He next follows up these with accounts of other cases in which emotional apathy seemed to keep pace with sensory anaesthesia, arguing that, according to his theory, a subject absolutely anaesthetic should also be incapable of emotion, although “emotion-inspiring objects might evoke the usual bodily expression from him.” Whether any testimony from lunatics, hypnotics and other minds diseased could suffice to establish this novel doctrine is questionable: that the evidence so far adduced is insufficient, Professor James himself seems to allow. There are some four or five of the apathetic cases altogether: three of them are regarded by the mental pathologists who describe them as adverse to Professor James's theory.[92] Of the fourth case, reported by a pathologist on Professor James's side, the latter himself candidly observes, “We must remember that the patient's inemotivity may have been a co-ordinate result with the anaesthesia of his neural lesions, and not the anaesthesia's mere effect.” This missing link in the argument is supplied by the experiments of Professor Sherrington,[93] and these show conclusively that normal emotional states are possible along with complete visceral anaesthesia. As to emotional excitement induced by intoxication or disease, and so far groundless, the most that can safely be said is that the object may be vague, ill-defined and shifting, but not that it is absent altogether. States of physical exaltation, depression or irritability readily arouse by association appropriate troupes of imagery; only when they fail of this are we entitled to say that there is no object, and then we must add that there is also no emotion.

Emotional and Conative Action.

35. As in dealing with the causes of feeling, so we may now in like manner proceed to inquire whether in its manifestations or effects there is any contrast corresponding to the opposing extremes of pleasure and pain. We have already seen reasons for dismissing reflex movements or movements not determined by feeling as psychologically secondary, the effects of habit and heredity, and for regarding those diffusive movements that are immediately expressive of feeling as primordial-such movements as are strictly purposive being gradually selected or elaborated from them. But some distinction is called for among the various movements expressive of emotion; for there is more in these than the direct effect of feeling regarded as merely pleasure or pain. It has been usual with psychologists to confound emotions with feeling, because intense feeling is essential to emotion. But, strictly speaking, a state of emotion is a complete state of mind, a psychosis, and not a psychical element, if we may so say. Thus in anger we have over and above pain a more or less definite object as its cause, and a certain characteristic reactive display—frowns, compressed lips, erect head, clenched fists, in a word, the combative attitude—as its effect, and similarly of other emotions; so that generally in the particular movements indicative of particular emotions the primary and primitive effects of feeling are overlaid by what Darwin has called serviceable associated habits. The purposive actions of an earlier stage of development become, though somewhat atrophied as it were, the emotive outlet of a later stage: in the circumstances in which our ancestors worried their enemies we only show our teeth. We must, therefore, leave aside the more complex emotional manifestations and look only to the simplest effects of pleasure and of pain, if we are to discover any fundamental contrast between them.[94]

Joy finds expression in dancing, clapping the hands and meaningless laughter, and these actions are not only pleasurable Emotional Expression. in themselves but such as increase the existing pleasure. Attention is not drafted off or diverted; but rather the available resources seem reinforced, so that the old expenditure is supported as well as the new. To the pleasure on the receptive side is added pleasure on the active side. The violent contortions due to pain, on the other hand, are painful in themselves, though less intense than the pains from which they withdraw attention; they are but counter-irritants that arrest or inhibit still more painful thoughts or sensations. Thus, according to Darwin, “sailors who are to be flogged sometimes take a piece of lead into their mouths in order to bite it with their utmost force, and thus to bear the pain.” When in this way we take account of the immediate effects as well as of the causes of feeling, we find it still more strikingly true that only in pleasurable states is there an efficient expenditure of attention. It is needless now to dwell upon this point, although any earlier mention of it would hardly have been in place. But we should fail to realize the contrast between the motor effects of pleasure and of pain if we merely regarded them as cases of diffusion. The intenser the feeling the intenser the reaction, no doubt, whether it be smiles or tears, jumping for joy, or writhing in agony; but in the movements consequent on pleasure the diffusion is the result of mere exuberance, an overflow of good spirits, as we sometimes say, and these movements, as already remarked, are always comparatively purposeless or playful. Even the earliest expressions of pain, on the contrary, seem but so many efforts to escape from the cause of it; in them there is at least the blind purpose to flee from a definite ill, but in pleasure only the enjoyment of present fortune.

From Plato downwards psychologists and moralists have been fond of discussing the relation of pleasure and pain. It has been maintained that pain is the first and more fundamental fact, and pleasure nothing but relief from pain; and, again, on the other side, that pleasure is prior and positive, and pain only the negation of pleasure. So far as the mere change goes, it is obviously true that the diminution of pain is pro tanto pleasant, and the diminution of pleasure pro tanto unpleasant; and if relativity had the unlimited range sometimes assigned to it this would be all we could say. But we must sooner or later recognize the existence of a comparatively fixed neutral state, deviations from which, of comparatively short duration and of sufficient intensity, constitute distinct states of pleasure or pain. Such states, if not of liminal intensity, may then be further diminished without reversing their pleasurable or painful character. The turning-point here implied may, of course, gradually change too—as a result, in fact, of the law of accommodation. Thus a long run of pleasure would raise “the hedonistic zero,” while—to the small extent to which accommodation to pain is possible—a continuance of pain would lower it, But such admission makes no material difference where the actual feeling of the moment is alone concerned and retrospect out of the question. On the whole it seems, therefore, most reasonable to regard pleasure and pain as emerging out of a neutral state, which is prior to and distinct from both—not a state of absolute indifference, but of simple contentment, marked by no special active display. But it is by reference to such state of equilibrium or ἀπαθία that we see most clearly the superior volitional efficacy of gain upon which pessimists love to descant. “Nobody,” says Von Hartmann, “who had to choose between no taste at all for ten minutes or five minutes of a pleasant taste and then five minutes of an unpleasant taste, would prefer the last.” Most men and all the lower animals are content “to let well alone.”

To ascertain the origin and progress of purposive action it seems, then, that we must look to the effects of pain rather Purposive Action. than to those of pleasure. It is true that psychologists not infrequently describe the earliest purposive movements as appetitive, or at least they treat appetitive and aversive movements as co-ordinate and equally primitive, pleasures being supposed to lead to actions for their continuance as much as pains to actions for their removal. No doubt, as soon as the connexion between a pleasurable sensation and the appropriate action is completely established, as in the case of imbibing food, the whole process is then self-sustaining till satiety begins. But the point is that such facility was first acquired under the teaching of pain—the pain of unsatisfied hunger. The term “appetite” is apt both by its etymology and its later associations to be misleading. What are properly called the “instinctive” appetites are—when regarded from their active side—movements determined by some existing uneasy sensation. So far as their earliest manifestation in a particular individual is concerned, this urgency seems almost entirely of the nature of a vis a tergo; and the movements are only more definite than those simply expressive of pain because of inherited pre-adaptation, on, which account, of course, they are called “instinctive.” But what one inherits another must have acquired, and we have agreed here to leave heredity on one side and consider only the original evolution.

But if none but psychological causes were at work this evolution would be very long and in its early stages very uncertain. At first, when only random movements ensue, we may fairly suppose both that the chance of at once making a happy hit would be small and that the number of chances, the space for repentance, would also be small. Under such circumstances natural selection would have to do almost everything and subjective selection almost nothing. So far as natural selection worked, we should have, not the individual subject making a series of tries and perfecting itself by practice, as in learning to dance or swim, but we should have those individuals whose structure happened to vary for the better surviving, increasing and displacing the rest. How much natural selection, apparently unaided, can accomplish in the way of complicated adjustment we see in the adaptation of the form and colour of plants and animals to their environment. Both factors, in reality, operate at once, and it would be hard to fix a limit to either, though to our minds natural selection seems to lose in comparative importance as we advance towards the higher stages of life.

But psychologically we have primarily to consider subjective selection, i.e. first of all, the association of particular movements with particular sensations through the mediation of feeling. The sensations here concerned are mainly painful excitations from the environment, the recurring pains of innutrition, weariness, &c., and pleasurable sensations due to the satisfaction of these organic wants—pleasures which, although not a mere “filling-up,” as Plato at one time contended, are still preceded by pain, but imply over and above the removal of this a certain surplus of positive good. There seem only a few points to notice. (a) When the movements that ensue through pleasure are themselves pleasurable there is ordinarily no ground for singling out any one; such movements simply enhance the general enjoyment, which is complete in itself and so far contains no hint of anything beyond. (b) Should one of these spontaneous movements of pleasure chance to cause pain, no doubt such movement is speedily arrested. Probably the most immediate connexion possible between feeling and purposive action is that in which a painful movement leads through pain to its own suppression. But such connexion is not very fruitful of consequences, inasmuch as it only secures what we may call internal training and does little to extend the relation of the individual to its environment. (c) Out of the irregular, often conflicting movements which indirectly relieve pain some one may chance to remove the cause of it altogether. Upon this movement, the last of a tentative series, attention, released from the pain, is concentrated; and in this way the evil and the remedy become so far associated that on a recurrence of the former the many diffused movements become less, and the one purposive movement more, pronounced; the one effectual way is atlength established and the others, which were but palliatives, disappear. (d) When things have advanced so far that some one definite movement is definitely represented along with the painful sensation it remedies, it is not long before a still further advance is possible and we have preventive movements. Thanks to the orderliness of things, dangers have their premonitions. After a time, therefore, the occurrence of some signal sensation revives the image of the harm that has previously followed in its wake, and a movement—either like the first, or another that has to be selected from the random tries of fear—occurs in time to avert the impending ill. (e) In like manner, provided the cravings of appetite are felt, any signs of the presence of pleasurable objects prompt to movements for their enjoyment or appropriation. In these last cases we have action determined by percepts. The cases in which the subject is incited to action by ideas as distinct from percept require a more detailed consideration; such are the facts mainly covered by the term “desire.”

By the time that ideas are sufficiently self-sustaining to form trains that are not wholly shaped by the circumstances of the Desire. present, entirely new possibilities of action are opened up. We can desire to live again through experiences of which there is nothing actually present to remind us, and we can desire a new experience which as yet we only imagine. We often, no doubt, apply the term to the simpler states mentioned under (e) in the last paragraph: the fox in the fable is said to have desired the grapes he vilified because out of his reach. Again, at the other extreme it is usual to speak of a desire for honour, or for wealth, and the like; but such are not so much single states of mind as inclinations or habitual desires. Moreover, abstractions of this kind belong to a more advanced stage of development than that at which desire begins, and of necessity imply more complicated grounds of action than we can at present examine. The essential characteristics of desire will be more apparent if we suppose a case somewhere between these extremes. A busy man reads a novel at the close of the day, and finds himself led off by a reference to angling or tropical scenery to picture himself with his rods packed en route for Scotland, or booked by the next steamer for the fairyland of the West Indies. Presently, while the ideas of Jamaica or fishing are at least as vividly imagined as before, the fancied preparations receive a rude shock as the thought of his work recurs. Some such case we may take as typical and attempt to analyse it.

First of all it is obviously true, at least of such more concrete desires, that what awakens desire at one time fails to do so at another, and that we are often so absorbed or content with the present as not to be amenable to (new) desires at all. A given x or y cannot, then, be called desirable per se, it is only desirable by relation to the contents of consciousness at the moment. Of what nature is this relation? (1) At the level of psychical life that we have now reached very close and complete connexions have been formed between ideas and the movements necessary for their realization, so that when the idea is vividly present these movements are apt to be nascent. This association is the result of subjective selection—i.e. of feeling—but being once established, it persists like other associations independently of it. (2) Those movements are especially apt to become nascent which have not been recently executed, which are therefore fresh and accompanied by the organic sensations of freshness, but also those which are frequently executed, and so from habit readily aroused. The latter fact, which chiefly concerns habitual desires, may be left aside for a time. (3) At times, then, when there is a lack of present interests, or when these have begun to wane, or when there is positive pain, attention is ready to fasten on any new suggestion that calls for more activity, requires a change of active attitude, or promises relief. Such spontaneous concentration of attention ensures greater vividness to the new idea, whatever it be, and to its belongings. In some cases this greater vividness may suffice. This is most likely to happen when the new idea affords intellectual occupation, and this is at the time congenial, or with indolent and imaginative persons who prefer dreaming to doing. (4) But when the new idea does not lead off the pent-up stream of action by opening out fresh channels, when, instead of this, it is one that keeps them intent upon itself in an attitude comparable to expectation, then we have desire. In such a state the intensity of the re-presentation is not adequate to the intensity of the incipient actions it has aroused. This is most obvious when the latter are directed towards sensations or percepts, and the former remains only an idea. If it were possible by concentrating attention to convert ideas into percepts, there would be an end of most desires: “if wishes were horses beggars would ride.” (5) But our voluntary power over movements is in general of this kind: here the fiat may become fact. When we cannot hear we can at least listen, and, though there be nothing to fill them, we can at least hold out our hands. It would seem, then, that the source of desire lies essentially in this excess of the active reaction above the intensity of the re-presentation (the one constituting the “impulse,” the other the “object” of desire, or the desideratum), and that this disparity rests ultimately on the fact that movements have, and sensations have not, a subjective initiative. (6) The impulse or striving to act will, as already hinted, be stronger the greater the available energy, the fewer the present outlets, and, habits apart, the fresher the new opening for activity. (7) Finally, it is to be noted that, when such inchoate action can be at once consummated, desire ends where it begins: to constitute a definite state of desire there must be not only an obstacle to the realization of the desideratum—if this were all we should rather call the state one of wishing—but an obstacle to its realization by means of the actions its representation has aroused.

However the desire may have been called forth, its intensity is primarily identical with the strength of this impulse to action, Relation of Desire to Feeling. and has no definite or constant relation to the amount of pleasure that may result from its satisfaction. The feeling directly consequent on desire as a state of want and restraint is one of pain, and the reaction which this pain sets up may either suppress the desire or prompt to efforts to avoid or overcome the obstacles in its way. To inquire into these alternatives would lead us into the higher phases of voluntary action; but we must first consider the relation of desire to feeling more closely.

Instances are by no means wanting of very imperious desires accompanied by the clear knowledge that their gratification will be positively distasteful.[95] On the other hand it is possible to recollect or picture circumstances known or believed to be intensely pleasurable without any desire for them being awakened at all: we can regret or admire without desiring. Yet there are many psychologists who maintain that desire is excited only by the prospect of the pleasure that may arise through its gratification, and that the strength of the desire is proportional to the intensity of the pleasure thus anticipated. Quidquid petitur petitur sub specie boni is their main formula. The plausibility of this doctrine rests partly upon a seemingly imperfect analysis of what strictly pertains to desire and partly on the fact that it is substantially true both of what we may call “presentation-prompted” action, which belongs to an earlier stage than desire, and of the more or less rational action that comes later. In the very moment of enjoyment it may be fairly supposed that action is sustained solely by the pleasure received and is proportional to the intensity of that pleasure. But there is here no re-presentation and no seeking; the conditions essential to desire, therefore, do not apply. Again, in rational action, where both are present, it may be true—to quote the words of an able advocate of the view here controverted—that “our character as rational beings is to desire everything exactly according to its pleasure value.”[96] But consider what such conceptions as the good, pleasure value and rational action involve. Here we have foresight and calculation, regard for self as an object of permanent interest—Butler's cool self-love; but desire as such is blind, without either the present certainty of sense or the assured prevision of reason. Pleasure in the past, no doubt, has usually brought about the association between the representation of the desired object and the movement for its realization; but neither the recollection of this pleasure nor its anticipation is necessary to desire, and even when present they do not determine what urgency it will have. The best proof of this lies in certain habitual desires. Pleasures are diminished by repetition, whilst habits are strengthened by it; if the intensity of desire, therefore, were proportioned to the “pleasure value” of its gratification, the desire for renewed gratification should diminish as this pleasure grows less; but, if the present pain of restraint from action determines the intensity of desire, this should increase as the action becomes habitual. And observation seems to show that, unless prudence suggests the forcible suppression of such belated desires or the active energies themselves fail, they do in fact become more imperious, although less productive of positive pleasure, as time goes on.

In this there is, of course, no exception to the general principle that action is consequent on feeling—a greater pleasure being preferred before a less, a less pain before a greater; for, though the feeling that follows upon its satisfaction be less or even change entirely, still the pain of the unsatisfied desire increases as the desire hardens into habit. It is also a point in favour of the position here taken that appetites, which may be compared to inherited desires, certainly prompt to action by present pain rather than by prospective pleasure.

Intellection.

36. Desire naturally prompts to the search for the means to its satisfaction and frequently to a mental rehearsal of various possible courses of action, their advantages and disadvantages. Thus, by the time the ideational continuum has become—mainly by the comparatively passive working of association—sufficiently developed to furnish free ideas as thinking material, motives are forthcoming for thinking to begin. It is obviously impossible to assign any precise time for this advance; like all others, it is gradual. Fitfully, in strange circumstances and under strong excitement, the lower animals give unmistakable signs that they can understand and reason. But thought as a permanent activity may be fairly said to originate in and even to depend upon the acquisition of speech. This indispensable instrument, which more than anything else enables our pyschological individual to advance to the distinctly human or rational stage, consists of gestures and vocal utterances, which were originally—and, indeed, are still to a large extent—emotional expressions.[97] Our space will only allow us to note in what way language when it already exists, is instrumental in the development as distinct from the communication of thought. But first of all, what in general is thinking, of which language is the instrument?

In entering upon this inquiry we are really passing one of the hardest and fastest lines of the old psychology—that between sense Distinction between Sense and Understanding. and understanding. So long as it was the fashion to assume a multiplicity of faculties the need was less felt for a clear exposition of their connexion. A man had senses and intellect much as he had eyes and ears; the heterogeneity in the one case was no more puzzling than in the other. But for psychologists who do not cut the knot in this fashion it is confessedly a hard matter to explain the relation of the two. The contrast of receptivity and activity hardly avails, for all presentation involves activity and essentially the same activity, that of attention. Nor can we well maintain that the presentations attended to differ in kind, albeit such a view has been held from Plato downwards. Nihil est in intellectu quod non fuerit prius in sensu: the blind and deaf are necessarily without some concepts that we possess. If pure being is pure nothing, pure thought is equally empty. Thought consists of a certain elaboration of sensory and motor presentations and has no content apart from these. We cannot even say that the forms of this elaboration are psychologically a priori; on the contrary, what is epistemologically the most fundamental is the last to be psychologically realized. This is not only true as a fact; it is also true of necessity, in so far as the formation of more concrete concepts is an essential preliminary to the formation of others more abstract—those most abstract, like the Kantian categories, &c., being thus the last of all to be thought out or understood. And though this formative work is substantially voluntary, yet, if we enter upon it, the form at each step is determined by the so-called matter, and not by us; in this respect “the spontaneity of thought” is not really freer than the receptivity of sense.[98] It is sometimes said that thought is synthetic, and this is true; but imagination is synthetic also; and the processes which yield the ideational train are the only processes at work in intellectual synthesis. Moreover, it would be arbitrary to say at what point the mere generic image ceases and the true concept begins—so continuous are the two. No wonder, therefore, that English psychology has been prone to regard thought as only a special kind of perception—perceiving the agreement or disagreement of ideas—and the ideas themselves as mainly the products of association. Yet this is much like confounding observation with experiment or invention—the act of a cave-man in betaking himself to a drifting tree with that of Noah in building himself an ark. In reverie, and even in understanding the communications of others, we are comparatively passive spectators of ideational movements, non-voluntarily determined. But in thinking or “intellection,” as it has been conveniently termed, there is always a search for something more or less vaguely conceived, for a clue which will be known when it occurs by seeming to satisfy certain conditions. Thinking may be broadly described as solving a problem—finding an AX that is B. In so doing we start from a comparatively fixed central idea or intuition and work along the several diverging lines of ideas associated with it—hence far the aptest and in fact the oldest description of thought is that it is discursive. Emotional excitement—and at the outset the natural man does not think much in cold blood—quickens the flow of ideas: what seems relevant is at once contemplated more closely, while what seems irrelevant awakens little interest and receives little attention. At first the control acquired is but very imperfect; the actual course of thought of even a disciplined mind falls far short of the clearness, distinctness, and coherence of the logician's ideal. Familiar associations are apt to hurry attention away from the proper topic, so that thought becomes not only discursive but wandering; in place of concepts of fixed and crystalline completeness, such as logic describes, we may find a congeries of ideas but imperfectly compacted into one generic idea, subject to continual transformation and implicating much that is irrelevant and confusing.

Thus, while it is possible for thought to begin without language, just as arts may begin without tools, yet language enables us Thought and Language. to carry the same process enormously farther. In the first place it gives us an increased command of even such comparatively concrete generic images as can be formed without it. The name of a thing or action becomes, for one who knows the name, as much an objective mark or attribute as-any quality whatever can be. The form and colour of what we call an “orange” are perhaps even more intimately combined with the sound and utterance of this word than with the taste and fragance which we regard as strictly essential to the thing. But, whereas its essential attributes often evade us, we can always command its nominal attribute, in so far as this depends upon movements of articulation. By uttering the name (or hearing it uttered) we have secured to us, in a greater or less degree, that superior vividness and definiteness that pertain to images reinstated by impressions: our idea approximates to the fixity and independence of a percept (cf. § 21 above). With young children and uncultured minds—who, by the way, not uncommonly “think aloud”—the gain in this respect is probably more striking than those not confined to their mother-tongue or those used to an analytical handling of language at all realize.[99] When things are thus made ours by receiving names from us and we can freely manipulate them in idea, it becomes easier mentally to bring together facts that logically belong together, and so to classify and generalize. For names set us free from the cumbersome tangibility and particularity of perception, which is confined to just what is presented here and now. But as ideas increase in generality they diminish in definiteness and unity; they not only become less pictorial and more schematic, but they become vague and unsteady as well, because formed from a number of concrete images only related as regards one or two constituents, and not assimilated as the several images of the same thing may be. The mental picture answering to the word “horse” has, so to say, body enough to remain a steady object when under attention from time to time; but that answering to the word “animal” is perhaps scarcely twice alike. The relations of things could thus never be readily recalled or steadily controlled if the names of those relations, which as words always remain concrete, did not give us a definite hold upon them—make them comprehensible. Once these “airy nothings” have a name, we reap again the advantages a concrete constituent affords: by its means that which is relevant becomes more closely associated, and that which is irrelevant—abstracted from—falls off. When what answers to the logical connotation or meaning of a concept is in this way linked with the name, it is no longer necessary that such “matter or content” should be distinctly present in consciousness. It takes time for an image to raise its associates above the threshold; and, when all are there, there is more demand upon attention in proportion. There is thus a manifest economy in what Leibnitz happily styled “symbolic,” in contrast to “intuitive” thinking. Our power of efficient attention is limited, and with words for counters we can, as Leibnitz remarks, readily perform operations involving very complex presentations, and wait till these operations are concluded before realizing and spreading out the net result in sterling coin.

But this simile must not mislead us. In actual thinking there never is any complete separation between the symbol Thought and Ideation. and the ideas symbolized: the movements of the one are never entirely suspended till those of the other are complete. “Thus,” says Hume, “if, instead of saying, that in war the weaker have always recourse to negotiation, we should say, that they have always recourse to conquest, the custom which we have acquired of attributing certain relations to ideas still follows the words and makes us immediately perceive the absurdity of that proposition.”[100] How intimately the two are connected is shown by the surprises that give what point there is to puns, and by the small confusion that results from the existence of homonymous terms. The question thus arises—What are the properly ideational elements concerned in thought? Over this question psychologists long waged fight as either nominalists or conceptualists. The former maintain that what is imaged in connexion with a general concept, such as triangle, is some individual triangle “taken in a certain light,”[101] while the latter maintain that an “abstract idea” is formed embodying such constituents of the several particulars as the concept connotes, but dissociated from the specific or accidental variations that distinguish one particular from another. As often happens in such controversies, each side saw the weak point in the other. The nominalists easily showed that there was no distinct abstract idea representable apart from particulars; and the conceptualists could as easily show that a particular presentation “considered in a certain light” is no longer merely a particular presentation nor yet a mere crowd of presentations. The very thing to ascertain is what this consideration in a certain light implies. Perhaps a speedier end might have been put to this controversy if either party had been driven to define more exactly what was to be understood by image or idea. Such ideas as are possible to us apart from abstraction are, as we have seen, revived percepts, not revived sensations, are complex total re-presentations made up of partial re-presentations, which may figure in other totals (cf. § 21). Reproductive imagination is so far but a faint rehearsal of actual percepts, and constructive imagination but a faint anticipation of possible percepts. In either case we are busied with elementary presentations complicated or synthesized to what are tantamount to intuitions, in so far as the forms of intuition remain in the idea, though the fact, as tested by movement, &c., is absent. The several partial re-presentations, however, which make up an idea might also be called ideas, not merely in the wide sense in which every mental object may be so called, but also in the narrower sense as secondary presentations, i.e. as distinguished from primary presentations or impressions. But such isolated images of an impression, even if possible, would no more be intuitions than the mere impression itself would be one: taken alone the one would be as free of space and time as is the other. Till it is settled, therefore, whether the ideational elements concerned in conception are intuitive complexes or something answering to the ultimate elements of these, nothing further can be done.

In the case of what are specially called “concrete” as distinct from “abstract” concepts—if this rough-and-ready, but unscientific, distinction may be allowed—the idea answering to the concept differs little from an intuition, and we have already remarked that the generic image (Gemeinbild of German psychologists) constitutes the connecting link between imagination and conception. But even concerning these it is useless to ask what does one imagine in thinking, e.g. of triangle or man or colour. We never—except for the sake of this very inquiry—attempt to fix our minds in this manner upon some isolated concept; in actual thinking ideas are not in consciousness alone and disjointedly, but as part of a context. When the idea “man” is present, it is present in some proposition or question, as—Man is the paragon of animals; In man there is nothing great but mind; and so on. It is quite clear that in understanding or mentally verifying such statements very different constituents out of the whole complex “man” are prominent in each. Further, what is present to consciousness when a general term is understood will differ, not only with a different context, but also the longer we dwell upon it: we may either analyse its connotation or muster its denotation, as the context or the cast of our minds may determine. Thus what is relevant is alone prominent, and the more summary the attention we bestow the less the full extent and intent of the concept are displayed. To the nominalist's objection, that it is impossible to imagine a man without imagining him as either tall or short, young or old, dark or light, and so forth, the conceptualise might reply that at all events percepts may be clear without being distinct, that we can recognize a tree without recognizing what kind of tree it is, and that, moreover, the objection proves too much: for, if our image is to answer exactly to fact, we must represent not only a tall or a short man, but a man of definite stature—one not merely either light or dark, but of a certain precise complexion. But the true answer rather is that in conceiving as such we do not necessarily imagine a man or a tree at all, any more than—if such an illustration may serve—in writing the equation to the parabola we necessarily draw a parabola as well.

The individuality of a concept is thus not to be confounded with the sensible concreteness of an intuition either distinct or indistinct, and “the pains and skill” which Locke felt were required in order to frame what he called an abstract idea are not comparable to the pains and skill that may be necessary to discriminate or decipher what is faint or fleeting. The material “framed” consists no doubt of ideas, if by this is meant that in thinking we work ultimately with the ideational continuum, but what results is never a mere intuitive complex nor yet a mere group of such. The concept or “abstract idea” only emerges when acertain intelligible relation is established among the members of such a group; and the very same intuition may furnish the material for different concepts as often as a different geistiges Band is drawn between them. The stuff of this bond, as we have seen, is the word, and this brings into the foreground of consciousness when necessary those elements—whether they form an intuition or not—which are relevant to the concept. Conception, then, is not identical with imagination, although the two terms are still often, and were once generally, regarded as synonymous. The same ultimate materials occur in each; but in the one they start with and retain a sensible form, in the other they are elaborated into the form which is called “intelligible.”

37. The distinctive character of this intellectual synthesis lies, we have seen, in the fact that it is determined entirely General Character and Growth of Intellection. by what is synthesized, whether that be the elementary constituents of intuitions or general relations of whatever kind among these. It differs, therefore, in being selective from the synthesis of association, which rests upon contiguity and unites together whatever occurs together. It differs also from any synthesis, though equally voluntary in its initiation, which is determined by a purely subjective preference, since intellection depends upon objective relations alone. Owing to the influence of logic, which has long been in a much more forward state than psychology, it has been usual to resolve intellect ion into comparison, abstraction, and classification, after this fashion: ABCM and ABCN are compared, their differences M and N left out of sight, and the class notion ABC formed including both; the same process repeated with ABC and ABD yields a higher class notion AB; and so on. But our ideational continuum is not a mere string of ideas of concrete things, least of all such concrete things as this view implies. Not till our daily life resembles that of a museum porter receiving specimens will our higher mental activity be comparable to that of the savant who sorts such specimens into cases and compartments. What we perceive is a world of things in continual motion, waxing, waning, the centres of manifold changes, affecting us and apparently affected by each other, amenable to our action and, as it seems, continually interacting among themselves. Even the individual thing, as our analysis of perception has attempted to show, is not a mere sum of properties which can be taken to pieces and distributed like type, but a whole combined of parts very variously related. To understand intellection we must look at its actual development under the impetus of practical needs, rather than to logical ideals of what it ought to be. Like other forms of purposive activity, thinking is primarily undertaken as a means to an end, and especially the end of economy. It is often easier and always quicker to manipulate ideas than to manipulate real things; to the common mind the thoughtful man is one who “uses his head to save his heels.” In all the arts of life, in the growth of language and institutions, in scientific explanation, and even in the speculations of philosophy, we may remark a steady simplification in the steps to a given end or conclusion, or—what is for our present inquiry the same thing—the attainment of better results with the same means. The earliest machines are the most cumbrous and clumsy, the earliest speculations the most fanciful and anthropomorphic. Gradually imitation yields to invention, the natural fallacy of post hoc, ergo propter hoc to methodical induction, till what is essential and effective is realized and appreciated and what is accidental and inert is discarded and falls out of sight. In this way man advances in the construction of a complete mental clue or master key to the intricacies of the real world, but this key is still the counterpart of the world it enables us to control and explain.

To describe the process by which such insight is attained as a mere matter of abstraction deserves the stigma of “soulless blunder” which Hegel applied to it. Of course if attention is concentrated on X it must pro tanto be abstracted from Y, and such command of attention may require “some pains and skill.” But to see in this invariable accompaniment of thinking its essential feature is much like the schoolboy's saying that engraving consists in cutting fine shavings out of a hard block. The great thing is to find out what are the light-bearing and fruit-bearing combinations. Moreover, thinking does not begin with a conscious abstraction of attention from recognized differences in the way logicians describe. The actual process of generalization, for the most part at all events, is much simpler. The same name is applied to different things or events because only their more salient features are perceived at all. Their differences, so far from being consciously and with effort left out of account, often cannot be observed when attention is directed to them: to the inexperienced all is gold that glitters. Thus, and as an instance of the principle of progressive differentiation already noted (§ 6), we find genera recognized before species, and the species obtained by adding on differences, not the genus by abstracting from them. Of course such vague and indefinite concepts are not at first logically general: they only become so when certain common elements are consciously noted as pertaining to presentations in other respects qualitatively different, as well as numerically distinct. But actually thinking starts from such more potential generality as is secured by the association of a generic image with a name. So far the material of thought is always general—is freed, that is, from the local and temporal and other defining marks of percepts.

38. The process of thinking itself is psychologically much better described as (1) an analysis and (2) a re-synthesis of Thought as Analysis. this material already furnished by the ideational trains. The logical resolution of thought into hierarchies of concepts arranged like Porphyry's tree, into judgments uniting such concepts by means of a logical copula, &c., is the outcome of later reflection—mainly for technical purposes—upon thought as a completed product, and entirely presupposes all that psychology has to explain. The logical theory of the formation of concepts by generalization (or abstraction) and by determination (or concretion)—i.e. by the removal or addition of defining marks—assumes the previous existence of the very things to be formed, for these marks or attributes—X's and Y's, A's and B's—are themselves already concepts. Moreover, the act of generalizing or determining is really an act of judgment, so that the logician's account of conception presupposes judgment, while at the same time his account of judgment presupposes conception. But this is no evil; for logic does not essay to exhibit the actual genesis of thought but only an ideal for future thinking. Psychologically, however—that is to say, chronologically—the judgment is first. The growing mind, we may suppose, passes beyond simple perception when some striking peculiarity in what is at the moment perceived is a bar to its recognition. The stalking hunter is not instantly recognized as the destroying biped, because he crawls on all fours; or the scarecrow looks like him, and yet not like him, for, though it stands on two legs, it never moves. There is thus no immediate assimilation; recognition under such circumstances is in itself a judgment, involving an analysis more or less explicit. But of more account is the further judgment to which it leads, that which connects the new fact with the generic idea. Though actually complex, generic images are not explicitly known as complexes when they first enter into judgments; as the subjects of such judgments they are but starting-points for predication—It crawls; It does not move; and the like. Such impersonal judgments, according to most philologists, are in fact the earliest; and we may reasonably suppose that by means of them our generic images have been partially analysed, and have attained to something of the distinctness and constancy of logical concepts. But the analysis is rarely complete: a certain confused and fluctuating residuum remains behind. The psychological concept merges at sundry points into those cognate with it—in other words, the continuity of the underlying memory-train still operates; only the ideal concept of logic is in all respects totus, teres, atque rotundus. Evidence of this, if it seem to any to require proof, is obtainable on all sides, and, if we could recover the first vestiges of thinking, would doubtless be more abundant still.

But, if we agree that it is through acts of judgment which successively resolve composite presentations into elements that Logical Bias in Psychology. concepts first arise, it is still very necessary to inquire more carefully what these elements are. On the one side we have seen logicians comparing them to so many letters, and on the other psychologists enumerating the several sensible properties of gold or wax—their colour, weight, texture, &c.—as instances of such elements. In this way formal logic and sensationalist psychology have been but blind leaders of the blind. Language, which has enabled thought to advance to the level at which reflection about thought can begin, is now an obstacle in the way of a thorough analysis of it. A child or savage would speak only of “red” and “hot,” but we of “redness” and “heat.” They would probably say, “Swallows come when the days are lengthening and snipe when they are shortening”; we say, “Swallows are spring and snipe are winter migrants.” Instead of “The sun shines and plants grow,” we should say, “Sunlight is the cause of vegetation.” In short, there is a tendency to resolve all concepts into substantive concepts; and the reason of this is not far to seek. Whether the subject or starting-point of our discursive thinking be actually what we perceive as a thing, or whether it be a quality, an action, an effectuation (i.e. a transitive action), a concrete spatial or temporal relation, or finally, a resemblance or difference in these or in other respects, it becomes by the very fact of being the central object of thought pro tanto a unity, and all that can be affirmed concerning it may so far be regarded as its property or attribute. It is, as we have seen, the characteristic of every completed concept to be a fixed and independent whole, as it were, crystallized out of the still-fluent matrix of ideas. Moreover, the earliest objects of thought and the earliest concepts must naturally be those of the things that live and move about us; hence, then—to seek no deeper reason for the present—this natural tendency, which language by providing distinct names powerfully seconds, to reify or personify not only things but every element and relation of things which we can single out, or, in other words, to concrete our abstracts.[102] It is when things have reached this stage that logic begins. But ordinary, so-called formal, logic, which intends to concern itself not with thinking but only with the most general structure of thought, is debarred from recognizing any difference between concepts that does not affect their relations as terms in a proposition. As a consequence it drifts inevitably into that compartmental logic or logic of extension which knows nothing of categories or predicables, but only of the one relation of whole and part qualitatively considered. It thus pushes this reduction to a common denomination to the utmost: its terms, grammatically regarded, are always names and symbolize classes or compartments of things. From this point of view all disparity among concepts, save that of contradictory exclusion, and all connexion, save that of partial coincidence, are at an end.

Of a piece with this are the logical formula for a simple judgment, X is Y, and the corresponding definitions of judgment as the comparison of two concepts and the recognition of their agreement or disagreement.[103] It certainly is possible to represent every judgment as a comparison, although the term is strictly adequate only to judgments of one kind and affords but a very artificial description of others. But for a logic mainly concerned with inference—i.e. with explicating what is implicated in any given statements concerning classes—there is nothing more to be done than to ascertain agreements or disagreements; and the existence of these, if not necessarily, is at least most evidently represented by spatial relations. Such representation obviously implies a single ground of comparison only and therefore leaves no room for differences of category. The resolution of all concepts into class concepts and that of all judgments into comparisons thus go together. On this view if a concept is complex it can only be so as a class combination; and, if the mode of its synthesis could be taken account of at all, this could only be by treating it too as an element in the combination like the rest: iron is a substance, &c., virtue a quality, &c., distance a relation, &c., and so on. There is much of directly psychological interest in this thoroughgoing reduction of thought to a form which makes its consistency and logical concatenation conspicuously evident. But of the so-called matter of thought it tells us nothing. And, as said, there are many forms in that matter of at least equal moment, both for psychology and for epistemology: these formal logic has tended to keep out of sight.

It has generally been under the bias of such a formal or computational logic that psychologists, and especially English psychologists, have entered upon the study of mind. They have brought with them an analytic scheme which affords a ready place for sensations or “simple ideas” as the elements of thought, but none for any differences in the combinations of these elements. Sensations being in their very nature concrete, all generality becomes an affair of names; and, as Sigwart has acutely remarked, sensationalism and nominalism always go together. History would have borne him out if he had added that a purely formal logic tends in like manner to be nominalistic.

If we are still to speak of the elements of thought, we must extend this term so as to include not only the sensory elements Forms of Synthesis. we are said to receive but three distinct ways in which this pure matter is combined: (1) the forms of intuition—Time and Space[104]; (2) the real categories—Substance, Attribute, State, Act, Effect, End or Purpose, &c.—the exact determination of which is not here in place; and (3) certain formal (logical and mathematical) categories—as Unity, Difference, Identity, Likeness. These cannot be obtained by such a process of abstraction and generalization as logicians and psychologists alike have been wont to describe. They are not primarily concepts more general than all others in the sense in which animal is more general than man, but rather distinct methods of relating or synthesizing presentations. Kant, though he accepted almost unquestioned the logic and psychology current in his day, has yet been the occasion, in spite of himself, of materially advancing both, and chiefly by the distinction he was led to make between formal and transcendental logic. In his exposition of the latter he brings to light the difference between the “functions of the understanding” in synthesizing—or, as we might say, organizing—percepts into concepts and the merely analytic subsumption of abc and abd under aba, b, c and d being what they may. Unlike other concepts, categories as such do not in the first instance signify objects of thought, however general, but these functions of the understanding in constituting objects. In fine, they all imply some special process, and the general characteristic of the resulting products is what we have first of all to note

Objects of Higher Order: their Analysis and Genesis.

39. By transposing a tune from one key to another we may obtain two entirely diverse aggregates of notes, and yet the melody may remain unchanged. On the other hand, by varying the order of the notes two distinct tunes may result from the same collection of tones. Sense furnishes merely the parts: whence, then, this identity of the whole in spite of their diversity, this diversity of the whole in spite of their identity? From the sameness or difference of the several “intervals,” it is replied. But the answer is insufficient; for the tune is a unity, not a mere series, and, further, with every interval the same problem recurs. For the interval, too, is a whole, though a simpler one: it does not necessarily change with a change of its constituents, nor remain the same as long as their distance is unaltered. Feelings and “associations,” again, cannot account for the result, inasmuch as such accompaniments are not invariably present: moreover, they obviously presuppose the melody instead of producing it. Of such complex wholes or combinations—as distinct from mere aggregates or collections—there are many forms; as, for example, geometrical figures and patterns, motions and other changes, numbers, logical connexions, &c. In view of this variety it seems to strike the unprejudiced as wild to expect that “the progress of psychophysics” may disclose an explanation of such combinations conforming to the old scholastic maxim, Nihil est in intellectu quad non fuerit prius in sensu. Yet hopes of such a generatio aequivoca are entertained![105] Meanwhile the “old psychology,” at any rate, is content to regard such complex wholes as new presentations, the products, that is to say, not of a quasi-mechanical interaction of their constituents, but of intellectual synthesis.

What is here said of the combinations whereby the items of an aggregate are construed as parts of a whole holds equally of the comparisons whereby such items are related, as like or unlike, compatible or incompatible. Before either combination or comparison is possible, such items or particulars must be “given.” But it is conceivable that they should be given and no intellectual synthesis ensue; such a consciousness has been happily named anoetic.[106] Whether or no it actually exists is another matter: it is a conceivable limit, and has the theoretical usefulness of limiting conceptions generally. But relative anoesis suffices here. Suppose, then, we have: (a) item, a sound; item, ditto; item, ditto; or (b) item, blue; item, green. The sensationalist, from Hume onwards, has complained that he does not find in the one case a further item: total three; nor in the other a further item: unlikeness. After vainly seeking the living whole among the dead particulars, he next surmises that they generate it by their conjoint action! But whence this notion of “action”; and how, if such disjecta membra suffice, do they so often fail of their effect, so that we cannot “see the wood for the trees”? Combinations and comparisons then, we conclude, are not given, but “grounded” on what is given, and is thus their fundamentum. Hence Meinong, who has studied the psychology of intellection with especial care, has called the new presentations, due to this process of “grounding” (Fundiren), “objects of a higher order,” or ideal objects.[107] They have validity in respect of the particulars on which they are grounded, but not reality as data existing for perception alongside of such particulars.

The reader will here be reminded of Hume's distinction between knowledge and probability. His four philosophical relations, “which, depending solely upon ideas, can be the objects of knowledge and certainty—resemblance, continuity, degrees in quality an proportions in quantity or number”—are objects of higher order and ideal. “The other three, which depend not upon the idea, and may be absent or present even while that remains the same”—namely, identity, the situations in time and place, and causation—are thus obviously not the result of grounding or noesis merely, are not ideal but empirical, and have, that is to say, existential import. In fact, the second of these, the situations, though they imply synthesis in the wider sense in which all complex perception does, do not involve intellectual synthesis at all: are neither ideal combinations nor ideal relations. And since such temporal and spatial situations enter into both the other two—numerical identity and causation—the mixed, a posteriori character of these is obvious. Whatever be the defects of Hume's psychology, his classification of relations is so far sound, and its epistemological importance can hardly be overrated. It is accordingly to be regretted that the one vague term “relation” does not allow us to make these distinctions more precise. The German language, with the two terms Verhältniss and Beziehung, can do more.

It will be convenient at this point to digress somewhat for a moment to consider a question of some psychological interest. When we say that two “contents” are similar, and when too they admit of analysis, we can, if need be, enumerate certain elements as the ground of their partial likeness, and certain others as the ground of their partial diversity. We may further say that, abstracting from these last, we can regard the points of resemblance as constituting a general class to which the two contents belong as specific instances. But how is either comparison or abstraction possible when the two resembling contents appear as simple, and so far unanalysable? Instances, of course, are familiar to every one: thus we call red and orange colours, and say they resemble each other more than do red and blue. In presence of this question logicians and psychologists are apt to be at loggerheads. The logician maintains that abstraction and resemblance (as distinct from qualitative identity) imply complexity; and surely here he cannot be gainsaid. Yet there are the facts: reds and blues of sorts and a whole scale of degrees of likeness and unlikeness; but no constituent parts, no assignable marks of identity or diversity, are forthcoming, such as we find when we class sugar and salt together as solid or soluble, and pronounce them like in colour and unlike in taste. Here the logician's symbols a+b+c, a+b+d, have their counterparts: there—for the percipient's consciousness at all events—they have not. We cannot “consider and attend to either the sameness or the differences in” red and blue, as we can to the like or the unlike properties in salt and sugar. None the less it would be hasty to conclude that colours or any given sensations are simple. We are often struck by the likeness of complex wholes—two faces, say—long before we can discern the exact points of resemblance. Still, so long as there is no perceptible complexity in the individual presentations there can be no analysis of them, and, therefore, neither abstraction nor comparison based upon it. Can we find elsewhere the complexity that generalization and comparison invariably imply? Though colour may be regarded as a general term applicable alike to red, green and blue, just as animal is a general term applicable alike to bird, beast and fish, it is a mistake to infer that the processes are the same because of this similarity in their products. We seem bound to distinguish between consciously logical or “noetic” processes and processes that are unconsciously logical or “hyponoetic,” as we may perhaps call them. In the former the subjective aspect is left aside; in the latter it cannot be. The only common mark we can psychologically assign to colours is that they are all seen, and to tones—as the element of notes and noises—that they are all heard. So often as we talk of tasting tastes, smelling smells, feeling touches, language leads us to bear witness to this fact. When the sunset red changes to the twilight grey, I still see; but when the thunder follows the lightning there is a double change, though not an absolute one: from seeing I pass to hearing, but I am sentient still. And if progressive differentiation be the order of experience then the “universal” sentience precedes the differentiations seeing, hearing, &c., and, again, the “universal” colour the differentiations, red, green, blue, &c. Such “first universals,” then, are not reached by abstraction, but are given in the fundamental continuity of experience, and their subsequent differentiation admits neither of definition nor the classification applicable to discrete complexes, which are the material of logical comparison only. When red is pronounced liker or nearer to yellow than it is to green, this is because a smaller change is experienced in the transition from red to yellow than in that from red to green, and because in the latter yellow is reached and passed before green appears.[108] Proximity and resemblance are, then, so far one and the same; also both are equally relative, admit of the same indefinite gradation, and have the same limit in zero, regarded either as coincidence or identity. The conception of “distance between” answers, then, to what we have called a hyponoetic relation, and this is plainly distinct from the analysis of discrete complexes, with which, as said, noetic comparison is alone concerned: the one implies and the other excludes the notion of continuity and change—a fact which helps still further to distinguish the two.

Categories.

40. We come now to deal with the categories in more detail. To begin with what are par excellence formal categories, Formal Categories: Unity. and among these with that which is the most fundamental and formal of all—How do we come by the conception of unity? “Amongst all the ideas we have,” says Locke, “as there is none suggested to the mind by more ways, so there is none more simple than that of unity, or one. It has no shadow of variety or composition in it; every object our senses are employed about, every idea in our understandings, every thought of our minds, brings this idea along with it.”[109] But to assign a sensible origin to unity is certainly a mistake—one of a class of mistakes already more than once referred to, which consist in transferring to the data of sense all that is implied in the language necessarily used in speaking of them. The term “a sensation” no doubt carries along with it the idea of unity, but the bare sensation as received brings along with it nothing but itself. And, if we consider sensory consciousness merely, we do not receive a sensation, and then another sensation, and so on seriatim; but we have always a continuous diversity of sensations even when these are qualitatively sharply differentiated. Moreover, if unity were an impression of sense and passively received, it would, in common with other impressions, be unamenable to change. We cannot see red as blue, but we can resolve many (parts) into one (whole), and vice versa.[110] Unity, then, is the result of an act the occasions for which, no doubt, are at first non-voluntarily determined; but the act is still as distinct from them as is attention from the objects attended to. It is to that movement of attention already described in dealing with ideation (§ 24) that we must look as the source of this category. This same movement, in like manner, yields us temporal signs; and the complex unity formed by a combination of these is what we call number. When there is little or no difference between the field and the focus of attention, unifying is an impossibility, whatever the impressions received may be. On the other hand, as voluntary acts of concentration become more frequent and distinct the variegated continuum of sense is shaped into intuitions of definite things and events. Also, as soon as words facilitate the control of ideas, it becomes possible to single out special aspects and relations of things as the subjects or starting-points of our discursive thinking. Thus the forms of unity are manifold: every act of intuition or thought, whatever else it is, is an act of unifying.

It is obvious that the whole field of consciousness at any moment can never be actually embraced as one. What is unified becomes thereby the focus of consciousness and so leaves an outlying field; so far unity may be held to imply plurality. But it cannot with propriety be said that in a simple act of attention the field of consciousness is analysed into two distinct parts, i.e. two unities—this (now attended to) and the other or the rest (abstracted from). For the not-this is but the rest of a continuum and not itself a whole; it is left out but not determined, as the bounding space is left out when a figure is drawn. To know two unities we must connect both together; and herein comes to light the difference between the unity which is the form of the concept or subject of discourse and the unity of a judgment. The latter is of necessity complex; the former may or may not be. But in any case the complexity of the two is different. If the subject of thought is not only clear but distinct—i.e. not merely defined as a whole but having its constituents likewise more or less defined—such distinctness is due to previous judgments. At any future time these may of course be repeated; such are the analytical or explicative judgments of logic. As the mere subject of discourse it is, however, a single unity simultaneously apprehended; the relation ascertained between it and its predicate constitutes the unity of judgment, a unity which is comprehended only when its parts are successively apprehended.

But, though a judgment is always a complex unity, the extent of this complexity seems at first sight to vary as the form of Law of Dichotomy or Duality. synthesis varies. Formal logic, as we have seen, by throwing the form of synthesis into the predicate has no difficulty in reducing every judgment to an S is P. But, if we at all regard the matter thought, it is certain, for example, that “It is an explosion” is less complex than “The enemy explodes the mine.” The first answers one question; the second answers three. But as regards more complex judgment both the process of ascertaining the fact and the language in which it is expressed show that the three elements concerned in it are not synthesized at once. Suppose we start from the explosion—and changes or movements are not only apt to attract attention first, but, when recognized as events and not as abstracts personified, they call for some supplementing beyond themselves—then in this case we may search for the agent at work or for the object affected, but not for both at once. Moreover, if we find either, a complete judgment at once ensues: “The enemy explodes,” or “The mine is exploded.” The original judgment is really due to a synthesis of these two. But, when the results of former judgments are in this manner taken up into a new judgment, a certain “condensation of thought” ensues. Of this condensation the grammatical structure of language is evidence, though logical manipulation—with great pains—obliterates it. Thus our more complex judgment would take the form—“The enemy is now mine-exploding” or “The mine is enemy-exploded,” according as one or other of the simpler judgments was made first. An examination of other cases would in like manner tend to show that intellectual synthesis is always—in itself and apart from implications—a binary synthesis. Wundt, to whom belongs the merit of first explicitly stating this “law of dichotomy or duality”[111] as the cardinal principle of discursive thinking, contrasts it with synthesis by mere association. This, as running on continuously, he represents thus—A — B — C — D — . . . ; the synthesis of thought, on the other hand, he symbolizes by forms such as the following:—

AB AB CD AB   C
DE;
 &c.

Thus, Socrates is a philosopher; the philosopher Socrates discovered a method; the philosopher Socrates discovered the dialectical method; &c. The point is that the one thing attended to in an intellective act is the synthesis of two ideas, and of two ideas only, because, as only one movement of attention is possible at a time, only two ideas at a time can be synthesized. In that merely associative synthesis by which the memory continuum is produced attention moves from A to B and thence to C without any relation between A and B being attended to at all, although they must have relations, that of sequence e.g. at least.

“Difference,” says Hume, “I consider rather as a negation of relation than anything real or positive. Difference is of two Difference and Likeness. kinds, as opposed either to identity or resemblance. The first is called a difference of number, the other of kind.” The truth seems rather to be that difference in Hume's sense of numerical difference[112] is so far an element in all relations as all imply distinct correlatives. To this extent even identity—or at least the recognition of it—rests on difference, that form of difference, viz. which is essential to plurality. But absolute difference (i.e. diversity) of kind may be considered tantamount not, indeed, to the negation, but at least to the absence of all formal relation. That this absolute difference—or disparateness, as we may call it—affords no ground for relations becomes evident when we consider (1) that, if we had only a plurality of absolutely different presentations, we should have no consciousness at all (cf. § 11); and (2) that we never compare—although we distinguish—presentations which seem absolutely or totally disparate, as e.g. a thunderclap and the taste of sugar, or the notion of free trade and that of the Greek accusative. All actual comparison of what is qualitatively different rests upon at least partial likeness. This being understood, it is noteworthy that the recognition of unlikeness is, if anything, more “real or positive” than that of likeness, and is certainly the simpler of the two. In the comparison of sensible impressions—as of two colours, two sounds, the lengths or the directions of two lines, &c.—we find it easier in some cases to have the two impressions that are compared presented together, in others to have first one presented and then the other. But, either way, the essential matter is to secure the most effective presentation of their difference, which in every case is something positive and, like any other impression, may vary in amount from bare perceptibility to the extremest distance that the continuum to which it belongs will admit. Where no difference or distance at all is perceptible there we say there is likeness or equality. Is the only outcome, then, that when we pass from ab to ac there is a change in consciousness, and that when ab persists there is none? To say this is to take no account of the operations (we may symbolize them as acab:c, abab:o) by which the difference or the equality results. The change of presentation (c) and absence of change (o) are not here what they are as merely passive occurrences, so to put it. This is evident from the fact that in the former there is positive presentation and in the latter no presentation at all. The relation of unlikeness, then, is distinguished from the mere “position” or fact of change by (1) the voluntary concentration of attention upon ab and ac with a view to the detection of this change as their difference, and by (2) the act, relating them through it, in that they are judged unlike to that extent. The type of comparison is such superposition of geometrical lines or figures (as, e.g. in Euclid I. iv.): if they coincide we have concrete equality; if they do not their difference is a line or figure. All sensible comparisons conform essentially to this type. In comparing two shades we place them side by side, and passing from one to the other seek to determine not the absolute shade of the second but its shade relative to the first—in other words, we look out for contrast. We do not say of one “It is dark,” for in the scale or shades it may be light, but “It is darker”; or vice versa. Where there is no distance or contrast we simply have not two impressions, and, as said—if we consider the difference by itself—no impression at all. Two coincident triangles must be perceived as one. The distinction between the one triangle thus formed by two coinciding and the single triangle rests upon something extraneous to this bare presentation of a triangle that is one and the same in both cases. The marks of this numerical distinctness may be various: they may be different temporal signs, as in reduplications of the memory-continuum; or they may be constituents peculiar to each, from which attention is for the moment abstracted, any one of which suffices to give the common or identical constituent a new setting. In general, it may be said (1) that the numerical distinctness of the related terms is secured in the absence of all qualitative difference solely by the intellectual act which has so unified each as to retain what may serve as an individual mark; and (2) that they become related as “like” either in virtue of the active adjustment to a change of impression which their partial assimilation defeats, or in virtue of an anticipated continuance of the impression which this assimilation confirms.

It is in keeping with this analysis that we say in common speech that two things in any respect similar are so far the same. Identity. This ambiguity in the word “same,” whereby it means either individual identity or indistinguishable resemblance has been often noticed, and from a logical or objective point of view justly complained of as “engendering fallacies in otherwise enlightened understandings.” But apparently no one has inquired into its psychological basis, although more than one writer has admitted that the ambiguity is one “in itself not always to be avoided.”[113] It is not enough to trace the confusion to the existence of common names and to cite the forgotten controversies of scholastic realism. We are not now concerned with the conformity of thought to things or with logical analysis, but with the analysis of a psychological process. The tendency to treat presentations as if they were copies of things—the objective bias, as we may call it—is the one grand obstacle to psychological observation. Some only realize with an effort that the idea of extension is not extended; no wonder, then, if it should seem “unnatural” to maintain that the idea of two like things does not consist of two like ideas. But, assuming that both meanings of identity have a psychological justification, it will be well to distinguish them and to examine their connexion. Perhaps we might term the one “material identity” and the other “individual identity”—following the analogy of expressions such as “different things but all made of the same stuff,” “the same person but entirely changed.” Thus there is unity and plurality concerned in both, and herein identity or sameness differs from singularity or mere oneness, which so far entails no relation. But the unity and the plurality are different in each, and each is in some sort the converse of the other. In the one, two different individuals partially coincide; in the other, one individual is partially different; the unity in the one case is an individual presentation, in the other is the presentation of an individual.

In material identity the unity is that of a single presentation, whether simple or complex, which enters as a common Material Identity. constituent into two or more others. It may be possible, of course, to individuahze it, but as it emerges in a comparison it is a single presentation and nothing more. On account of this absence of individual marks this single presentation is what logicians call “abstract”; but this is not psychologically essential. It may be a generic image which has resulted from the neutralization of individual marks, but it may equally well be a simple presentation, like red, to which such marks never belonged. We come here from a new side upon a truth which has been already expounded at length, viz. that presentations are not given to us as individuals but as changes in a continuum. Time and space—the instruments, as it were, of individualization, which are presupposed in the objective sciences—are psychologically later than this mere differentiation.

The many vexed questions that arise concerning individual identity are metaphysical rather than psychological. But it Individual Identity. will serve to bring out the difference between the two forms of identity to note that an identification cannot be established solely by qualitative comparison; an alibi or a breach of temporal continuity will turn the flank of the strongest argument from resemblance. Moreover, resemblance itself may be fatal to identification when the law of being is change.

41. As regards the real categories, it may be said generally that these owe their origin in large measure to the anthropomorphic Real Categories. or mythical tendency of human thought—τὸ ὅμοιον τῷ ὁμοίῳ γινώσκεσθαι. Into the formation of these conceptions two very distinct factors enter—(1) the facts of what in the stricter sense we call “self-consciousness,” and (2) certain spatial and temporal relations among our presentations themselves. On the one hand, it has to be noted that these spatial and temporal relations are but the occasion or motive—and ultimately perhaps, we may say, the warrant—for the analogical attribution to things of selfness, efficiency and design, but are not directly the source of the forms of thought that thus arise. On the other hand, it is to be noted also that such forms, although they have an independent source, would never apart from suitable material come into actual existence. If the followers of Hume err in their exclusive reliance upon “associations naturally and even necessarily generated by the order of our sensations” (J. S. Mill), the disciple of Kant errs also who relies exclusively on “the synthetic unity of apperception.” The truth is that we are on the verge of error in thus sharply distinguishing the two at all; if we do so momentarily for the purpose of exposition it behoves us here again to remember that mind grows and is not made. The use of terms like “innate,” “a priori,” “necessary,” “formal,” &c., without further qualification leads only too easily to the mistaken notion that all the mental facts so named are alike underived and original, independent not only of experience but of each other; whereas but for the forms of intuition the forms of thought would be impossible—that is to say, we should never have a self-consciousness at all if we had not previously learnt to distinguish occupied and unoccupied space, past and present in time, and the like. But, again, it is equally true that, if we could not feel and move as well as receive impressions, and if experience did not repeat itself, we should never attain even to this level of spatial and temporal intuition. Kant shows a very lame and halting recognition of this dependence of the higher forms on the lower both in his schematism of the categories, and again in correcting in his Analytic the opposition of sense and understanding as respectively receptive and active with which he set out in his Aesthetic. Still, although what are called the subjective and objective factors of real knowledge advance together, the former is in a sense always a step ahead. We find again without us the permanence, individuality, efficiency, and adaptation we have found first of all within (cf. § 20, b and d). But such primitive imputation of personality, 'though it facilitates a first understanding, soon proves itself faulty and begets the contradictions which have been one chief motive to philosophy. We smile at the savage who thinks a magnet must need food or the child who is puzzled that the horses in a picture remain for ever still; but few consider that underlying all common-sense thinking there lurks the same natural precipitancy. We attribute to extended things a unity which we know only as the unity of an unextended subject; we attribute to changes among these extended things what we know only when we act and suffer ourselves; and we attribute further to them in their changes a striving for ends which we know only because we feel. In asking what they are, how they act, and why they are thus and thus, we assimilate them to ourselves, in spite of the differences which lead us by-and-by to see a gulf between mind and matter. Such instinctive analogies have, like other analogies, to be confirmed, refuted, or modified by further knowledge, i.e. by the very insight into things which these analogies have themselves made possible. That in their first form they were mythical, and that they could never have been at all unless originated in this way, are considerations that make no difference to their validity—assuming, that is, that they admit, now or hereafter, of a logical transformation which renders them objectively valid. This legitimation is, of course, the business of philosophy; we are concerned only with the psychological analysis and origin of the conceptions themselves.

42. As it must here suffice to examine one of these categories; let us take that which is the most important and central of the Causality. three, viz. causality or the relation of cause and effect, as that will necessarily throw some light upon the constitution of the others. To begin, we must distinguish three things, which, though very different, are very liable to be confused. (1) Perceiving in a definite case, e.g. that on the sun shining a stone becomes warm, we may say the sun makes the stone warm. This is a concrete instance of predicating the causal relation. In this there is, explicitly at all events, no statement of a general law or axiom, such as we have when we say (2) “Every event must have a cause”—a statement commonly known as the principle of causality. This again is distinct from what is on all hands allowed to be an empirical generalization, viz. (3) that such and such particular causes have invariably such and such particular effects. With these last psychology is not directly concerned at all: it has only to analyse and trace to its origin the bare conception of causation as expressed in (1) and involved in both these generalizations. Whether only some events have causes, as the notion of chance implies, whether all causes are uniform in their action or some capricious and arbitrary, as the unreflecting suppose—all this is beside the question for us.

One point in the analysis of the causal relation Hume may be said to have settled once for all: it does not rest upon or contain any immediate intuition of a causal nexus. The two relations that Hume allowed to be perceived (or “presumed to exist”), viz. contiguity in space of the objects causally related and priority in time of the cause before the effect, are the only relations directly discernible. We say indeed “The sun warms the stone” as readily as we say “The sun rises and sets,” as if both were matters of direct observation then and there. But that this is not so is evident from the fact that only in some cases when one change follows upon another do we regard it as following from the other: casual coincidence is at least as common as causal connexion. Whence the difference, then, if not from perception? Hume's answer,[114] repeated in the main by English psychologists since, is, as all the world knows, that the difference is the result of association, that when a change α in an object A has been frequently observed to precede a change β in another object B, this repetition determines the mind to a transition from the one to the other. It is this determination, which could not be present at first, that constitutes “the third relation betwixt these objects.” This “internal impression” generated by association is then projected; “for 'tis a common observation that the mind has a great propensity to spread itself on external objects.”

The subjective origin and the after-projection we must admit, but all else in Hume's famous doctrine seems glaringly at variance with facts. In one respect it proves too much, for not all constant sequences are regarded as causal, as according to his analysis they ought to be; again, in another respect it proves too little, for causal connexion is continually predicated on a first occurrence. The natural man has always distinguished between causes and signs or portents; but there is nothing to show that he produced an effect many times before regarding himself as the cause of it. J. S. Mill has indeed obviated the first objection epistemologically by adding to constant conjunction the further characteristic of “unconditionality.” But this is a conception that cannot be psychologically explained from Hume's premises, unless perhaps by resolving it into the qualification that the invariability must be complete and not partial, whereupon the second objection applies. “Unconditional” is a word for which we can find no meaning as long as we confine our attention to temporal succession. It will not do to say both that an invariable succession generates the idea, and that such invariable succession must be not only invariable but also unconditional in order to generate it. We may here turn the master against the disciple: “the same principle,” says Hume, “cannot be both the cause and the effect of another, and this is perhaps the only proposition concerning that relation which is either intuitively or demonstratively certain” (op. cit. p. 391). Unconditionality is then part of the causal relation and yet not the product of invariable repetition.

Perhaps the source of this element in the relation will become clear if we examine more closely the so-called “internal impression” of the mind, which according to Hume constitutes the whole of our idea of power or efficacy. To illustrate the nature of this impression Hume cites the instant passage of the imagination to a particular idea on hearing the word commonly annexed to it, when “'twill scarce be possible for the mind by its utmost efforts to prevent that transition” (op. cit. p. 393). It is this determination, then, which is felt internally, not perceived externally, that we mistakenly transfer to objects and regard as an intelligible connexion between them. But, if Hume admits this, must he not admit more? Can it be pretended that it is through the workings of association among our ideas that we first feel a determination which our utmost efforts can scarce resist, or that we feel such determination under no other circumstances? If it be allowed that the natural man is irresistibly determined to imagine an apple when he hears its name or, to expect thunder when he sees lightning, must it not also be allowed that he is irresistibly determined much earlier and in a much more impressive way when overmastered by the elements or by his enemies? But, further, such instances bring to light what Hume's “determination” also implies, viz. its necessary correlative, effort or action. Even irresistible association can only be known as such by efforts to resist it. Hume allows this when he says that his principles of association “are not infallible causes; for one may fix his attention during some time on any one object without looking farther” (op. cit. p. 393). But the fact is, we know both what it is to act and what it is to suffer, to go where we would and to be carried where we would not, quite apart from the workings of association. And, had Hume not confused the two different inquiries, that concerning the origin of the idea of causation and that concerning the ground of causal inference or law of causation, it could never have occurred to him to offer such an analysis of the former as he does.

Keeping to the former and simpler question, it would seem that when in ordinary thinking we say A causes this or that in B we project or analogically attribute to A what we experience in acting, and to B what we experience in being acted on; and the structure of language shows that such projection was made long before it was suspected that what A once did and B once suffered will be done and suffered in the same circumstances again. The occasions suitable for this projection are determined by the temporal and spatial relations of the objects concerned, which relations are matter of intuition. These are of no very special interest from a psychological point of view, but the subjective elements we shall do well to consider further. First of all, we must note the distinction of immanent action and transient action; the former is what we call action simply, and implies only a single thing, the agent; the latter, which we might with advantage call effectuation, implies two things, a patient as well as an agent. In scientific language the agent in an transitive act is called a causa immanens and so distinguished from the agent in effectuation or causa transiens. Common thought, however, does not regard mere action as caused at all; and we shall find it, in fact, impossible to resolve action into effectuation. But, since the things with which we ordinarily deal are complex, have many parts, properties, members, phases, and in consequence of the analytic procedure of thought, there ensues, indeed, a continual shifting of the point of view from which we regard any given thing, so that what is in one aspect one thing is in another many (cf. § 20). So it comes about that, when regarding himself as one, the natural man speaks of himself as walking, shouting, &c.; but, when distinguishing between himself and his members, he speaks of raising his voice, moving his legs, and so forth. Thus no sooner do we resolve any given action into an effectuation, by analytically distinguishing within the original agent an agent and a patient, than a new action appears. Action is thus a simpler notion than causation and inexplicable by means of it. It is certainly no easy problem in philosophy to determine where the resolution of the complex is to cease, at what point we must stop, because in the presence of an individual thing and a simple activity. At any rate, we reach such a point psychologically in the conscious subject, and that energy in consciousness we call attention. If this be allowed, Hume's critique of the notion of efficacy is really wide of the mark. “Some,”[115] he says, “have asserted that we feel an energy or power in our own mind; and that, having in this manner acquir'd the idea of power, we transfer that quality to matter, where we are not able immediately to discover it. . . . But to convince us how fallacious this reasoning is, we need only consider that the will, being here consider'd as a cause, has no more a discoverable connexion with its effects than any material cause has with its proper effect. . . . The effect is there [too] distinguishable and separable from the cause, and could not be foreseen without the experience of their constant conjunction” (op. cit. p. 455). This is logical analysis, not psychological; the point is that the will is not considered as a cause and distinguished from its effects, nor in fact considered at all. It is not a case of sequence between two separable impressions; for we cannot really make the indefinite regress that such logical distinctions as that between the conscious subject and its acts implies. Moreover, our activity as such is not directly presented at all; we are, being active; and further than this psychological analysis will not go.[116] There are, as we have seen, two ways in which this activity is manifested, the receptive or passive and the motor or active in the stricter sense—(cf. § 8) and our experience of these we project in predicating the causal relation. But two halves do not make a whole; so we have no complete experience of effectuation, for the simple reason that we cannot be two things at once. We are guided in piecing it together by the temporal and spatial relations of the things concerned. Hence, perhaps, some of the antinomies that beset this concept. In its earliest form, then, the so-called necessary connexion of cause and effect is perhaps nothing more than that of physical constraint. To this, no doubt, is added the strength of expectation—as Hume supposed—when the same effect has been found invariably to follow the same cause. Finally, when upon the basis of such associated uniformities of sequence a definite intellectual elaboration of such material ensues, the logical necessity of reason and consequent finds a place, and so far as deduction is applicable cause and reason become interchangeable ideas.

Belief.

43. The mention of logical necessity brings us to a new topic, viz. the “objectivity” of thought and cognition generally. The psychological treatment of this topic is tantamount to an inquiry into the characteristics of the states of mind we call certainty, doubt, belief—all of which centre round the one fact of evidence. Between the certainty that a proposition is true and the certainty that it is not there may intervene many grades of uncertainty. We may know that A is sometimes B, or sometimes not; or that some at least of the conditions of B are present or absent; or the, presentation of A may be too confused for distinct analysis. This is the region of probability, possibility, more or less obscurity. Leaving this aside, it will be enough to notice those cases in which certainty may be complete. With that certainty which is absolutely objective, i.e. with knowledge, psychology has no direct concern; it is for logic to furnish the criteria by which knowledge is ascertained.

Emotion and desire are frequent indirect causes of subjective certainty, in so far as they determine the constituents and the grouping of the field of consciousness at the moment—“pack the jury” or “suborn the witnesses,” as it were. But the ground of certainty is in all cases some quality or some relation of these presentations inter se. In a sense, therefore, the ground of all certainty is objective—in the sense, that is, of being something at least directly and immediately determined for the subject and not by it. Where certainty is mediate, one judgment is often spoken of as the ground of another; but a syllogism is still psychologically a single, though not a simple, judgment, and the certainty of it as a whole is immediate. Between the judgment A is B and the question Is A B? the difference is not one of content nor scarcely one of form: it is a difference which depends upon the effect of the proposition on the subject judging. (i.) We have this effect before us most clearly if we consider what is by common consent regarded as the type of certainty and evidence, the certainty of present sense-impressions whence it is said, “Seeing is believing.” The evident is here the actual, and the “feeling or consciousness” of certainty is in this case nothing but the sense of being taken fast hold of and forced to apprehend what is there. (ii.) The like is true of memory and expectation: in these also there is a sense of being tied down to what is given, whereas in mere imagination, however lively, this non-voluntary determination is absent (cf. § 26). Hume saw this at times clearly enough, as, e.g. when he says, “An idea assented to feels different from a fictitious idea that the fancy alone presents to us.” But unfortunately he not only made this difference a mere difference of intensity, but spoke of belief itself as “an operation of the mind” or “manner of conception that bestowed on our ideas this additional force or vivacity.”[117] In short, Hume confounded one of the indirect causes of belief with the ground of it, and again, in describing this ground committed the ὕστερον πρότερον of making the mind determine the ideas instead of the ideas determine the mind. (iii.) In speaking of intellection he is clearer: “The answer is easy with regard to propositions that are prov'd by intuition or demonstration. In that case the person who assents not only conceives the ideas according to the proposition, but is necessarily determin'd to conceive them in that particular manner” (op. cit. p. 395). It has been often urged—as by J. S. Mill, for example—that belief is something “ultimate and primordial.” No doubt it is; but so is the distinction between activity and passivity, and it is not here maintained that certainty can be analysed into something simpler, but only that it is identical with what is of the nature of passivity—objective determination. As Bain put it, “The leading fact in belief . . . is our primitive credulity. We begin by believing everything; whatever is is true” (Emotions and Will, 3d ed., p. 511). But the point is that in this primitive state there is no act answering to “believe” distinct from the non-voluntary attention answering to “perceive,” and no reflection such as a modal term like “true” implies. With eyes open in the broad day no man says, “I am certain there is light”; he simply sees. He may by-and-by come absolutely to disbelieve much that he sees—e.g. that things are nearer when viewed through a telescope—just as he will come to disbelieve his dreams, though while they last he is certain in these too. The consistency we find it possible to establish among certain of our ideas becomes an ideal, to which we expect to find all our experience conform. Still the intuitive evidence of logical and mathematical axioms is psychologically but a new form of the actual; we are only certain that two and two make four and we are not less certain that we see things nearer through a telescope.

Presentalion of Self, Self-Consciousness and Conduct.

44. The concept of self we have just seen underlying and to a great extent shaping the rest of our intellectual furniture; on this account it is at once desirable and difficult to analyse it and ascertain the conditions of its development. In attempting this we must carefully distinguish between the bare presentation of self and that reference of other presentations to it which is often called specially self-consciousness, “inner sense,” or internal perception. Concerning all presentations whatever—that of self no less than the rest—it is possible to reflect, “This presentation is mine; it is my object; I am the subject attending to it.” The presentation of self, then, is one presentation among others, the result, like them, of the differentiation of the original continuum. But it is obvious that this presentation must be in existence first before other presentations can be related to it. On the other hand, it is only in and by means of such relations that the concept of self is completed. We begin, therefore, with self simply as an object, and end with the concept of that object as the subject or “myself” that knows itself. The self has, first of all (a) a unique interest and (b) a certain inwardness, (c) it is an individual that (d) persists, (e) is active, and finally (f) knows itself. These several characteristics of self are intimately involved; so far as they appear at all they advance in definiteness from the lowest level of mere sentience to those moments of highest self-consciousness in which conscience approves or condemns volition.

The earliest and to the last the most important element in self—what we might perhaps term its root or material element—is that Self and the Body. variously styled the organic sensations—vital sense, coenaesthesis, or somatic consciousness. This largely determines the tone of the special sensations and enters, though little suspected, into all our higher feelings. If, as sometimes happens in serious nervous affections, the whole body or any part of it should lose common sensibility, the whole body or that part is at once regarded as strange and even as hostile. In some forms of hypochondria, in which this extreme somatic insensibility and absence of zest leave the intellect and memory unaffected, the individual doubts his own existence or denies it altogether. Ribot cites the case of such a patient, who, declaring that he had been dead for two years, thus expressed his perplexity:—“J'existe, mais en dehors de la vie réelle, matérielle, et, malgré moi, rien ne m'ayant donné la mort. Tout est mécanique chez moi et se fait inconsciemment.”[118] It is not because they accompany physiological functions essential to the efficiency of the organism as an organism, but simply because they are the most immediate and most constant sources of feeling, that these massive but ill-defined organic sensations are from the first the objects of the directest and most unreflecting interest. Other objects have at the outset but a mediate interest through subjective selection in relation to these, and never become so instinctively and inseparably identified with self, never have the same inwardness. This brings us to a new point. As soon as definite perception begins, the body as an extended thing is distinguished from other bodies, and such organic sensations as can be localized at all are localized within it. At the same time the actions of other bodies upon it are accompanied by pleasures and pains, while their action upon each other is not. The body also is the only thing directly set in motion by the reactions of these feelings, the purpose of such movements being to bring near to it the things for which there is appetite and to remove it from those towards which there is aversion. It is thus not merely the type of occupied space and the centre from which all positions are reckoned, but it affords us an unfailing and ever-present intuition of the actually felt and living self, to which all other things are external, more or less distant, and at times absent altogether. The body then first of all gives to self a certain measure of individuality, permanence and inwardness.

But with the development of ideation there arises within this what we may call an inner zone of self, having still more Inner Self. unity and permanence. We have at this stage not only an intuition of the bodily self doing or suffering here and now, but also memories of what it has been and done under varied circumstances in the past. External impressions have by this time lost in novelty and become less absorbing, while the train of ideas, largely increased in number, distinctness and mobility, diverts attention and often shuts out the things of sense altogether. In all such reminiscence or reverie a generic image of self is the centre, and every new image as it arises derives all its interest from relation to this; and so apart from bodily appetites new desires may be quickened and old emotions stirred again when all that is actually present is dull and unexciting. But desires and emotions, it must be remembered, though awakened by what is only imaginary, invariably entail actual organic perturbations, and with these the generic image of self comes to be intimately united. Hence arises a contrast between the inner self, which the natural man locates in his breast or φρήν, the chief seat of these emotional disturbances, and the whole visible and tangible body besides. Although from their nature they do not admit of much ideal representation, yet, when actually present, these organic sensations exert a powerful and often irresistible influence over other ideas; they have each their appropriate train, and so heighten in the very complex and loosely compacted idea of self those traits they originally wrought into it, suppressing to an equal extent all the rest. Normally there is a certain equilibrium to which they return, and which, we may suppose, determines the so-called temperament, naturel or disposition, thus securing some tolerable uniformity and continuity in the presentation of self. But even within the limits of sanity great and sudden changes of mood are possible, as, e.g. in hysterical persons or those of a “mercurial temperament,” or among the lower animals at the onset of parental or migratory instincts. Beyond those limits—as the concomitant apparently of serious visceral derangements or the altered nutrition of parts of the nervous system itself—complete “alienation” may ensue. A new self may arise, not only distinct from the old and devoid of all save the most elementary knowledge and skill that the old possessed, but diametrically opposed to it in tastes and disposition—obscenity, it may be, taking the place of modesty and cupidity or cowardice succeeding to generosity or courage. The most convincing illustrations of the psychological growth and structure of the presentation of self on the lower levels of sensation and ideation are furnished by these melancholy spectacles of minds diseased; but it is impossible to refer to them in detail here.[119]

Passing to the higher level of intellection, we come at length upon the concept which every intelligent being more or less Self as a Person. distinctly forms of himself as a person, M. or N., having such and such a character, tastes and convictions, such and such a history, and such and such an aim in life. The main instrument in the formation of this concept, as of others, is language, and especially the social intercourse that language makes possible. Up to this point the presentation of self has shaped that of not-self,—that is to say, external things have been comprehended by the projection of its characteristics. But now the order is in a sense reversed: the individual advances to a fuller self-knowledge by comparing the self within with what is first discernible in other persons without. So far avant l'homme est la société; it is through the “us” that we learn of the “me” (cf. § 36, note 1). Collective action for common ends is of the essence of society, and in taking counsel together for the good of his tribe each one learns also to take counsel with himself for his own good on the whole; with the idea of the common weal arises the idea of happiness as distinct from momentary gratification. The extra-regarding impulses are now confronted by a reasonable self-love, and in the deliberations that thus ensue activity attains to its highest forms—those of thought and volition. In the first we have a distinctly active manipulation of ideas as compared with the more passive spectacle of memory and imagination. Thereby emerges a contrast between the thinker and these objects of his thought, including among them the mere generic image of self, from which is now formed this concept of self as a person. A similar, even sharper, contrast also accompanies the exercise of what is very misleadingly termed “self-control,” i.e. control by this personal self of “the various natural affections”—to use Butler's phrase—which often hinder it as external objects hindered them. It is doubtful whether the reasoning, regulating self is commonly regarded as definitely localized. The effort of thinking and concentrating attention upon ideas is no doubt referred to the brain, but this is only comparable with the localization of other efforts in the limbs; when we think we commonly feel also, and the emotional basis is of all the most subjective and inalienable. If we speak of this latest phase of self as par excellence “the inner self,” such language is then mainly figurative, inasmuch as the contrasts just described are contrasts into which spatial relations do not enter.

45. The term “reflection,” or internal perception is applied to that state of mind in which some particular presentation or group Self-consciousness. of presentations (x or y) is not simply in the field of consciousness but there as consciously related to self, which is also presented at the same time. Self here may be symbolized by M, to emphasize the fact that it is in like manner an object in the field of consciousness. The relation of the two is commonly expressed by saying, “This (x or y) is my (M's) percept, idea or volition; I (M) it is that perceive, think, will it.” Self-consciousness, in the narrowest sense, as when we say “I know myself, I am conscious that I am,” &c., is but a special, though the most important, instance of this internal perception: here self (M) is presented in relation to self (with a difference, M′); the subject itself—at least, so we say—is or appears as its own object.

It has been often maintained that the difference between consciousness and reflection is not a real difference, that to know and to know that you know are “the same thing considered in different aspects.”[120] But different aspects of the same thing are not the same thing, for psychology at least. Not only is it not the same thing to feel and to know that you feel; but it might even be held to be a different thing still to know that you feel and to know that you know that you feel—such being the difference perhaps between ordinary reflection and psychological introspection.[121] The difficulty of apprehending these facts and keeping them distinct seems obviously due to the necessary presence of the earlier along with the later; that is to say, we can never know that we feel without feeling. But the converse need not be true. How distinct the two states are is shown in one way by their notorious incompatibility, the direct consequence of the limitation of attention: whatever we have to do that is not altogether mechanical is ill done unless we lose ourselves in the doing of it. This mutual exclusiveness receives a further explanation from the fact so often used to discredit psychology, viz. that the so-called introspection, and indeed all reflection, are really retrospective. It is not while we are angry or lost in reverie that we take note of such states, but afterwards, or by momentary side glances intercepting the main interest, if this be not too absorbing.

But we require an exacter analysis of the essential fact in this retrospect—the relation of the presentation x or y to that of self or M. What we have to deal with, it will be observed, is, implicitly at least, a judgment. First of all, then, it is noteworthy that we are never prompted to such judgments by everyday occurrences or acts of routine, but only by matters of interest, and, as said, generally when these are over or have ceased to be all-engrossing. Now in such cases it will be found that some effect of the preceding state of objective absorption persists, like wounds received in battle, unnoticed till the fight is over—such e.g. as the weariness of muscular exertion or of long concentration of attention; some pleasurable or painful after-sensation passively experienced, or an emotional wave subsiding but not yet spent; “the jar of interrupted expectation,” or the relief of sudden attainment after arduous striving, making prominent the contrast of contentment and want in that particular; or, finally, the quiet retrospect and mental rumination in which we note what time has wrought upon us and either regret or approve what we were and did. All such presentations are of the class out of which, as we have seen, the presentation of self is built up, and so form in each case the concrete bond connecting the generic image of self with its object. In this way and in this respect each is a concrete instance of what we call a state, act, affection, &c., and the judgments in which such relations to the standing presentation of self are recognized are the original and the type of all real predications. The opportunities for reflection are at first few, the materials being as it were thrust upon attention, and the resulting “percepts” are but vague. By the time, however, that a clear concept of self has been attained the exigencies of life make it a frequent object of contemplation, and as the abstract of a series of instances of such definite self-consciousness we reach the purely formal notion of a subject or pure ego. For empirical psychology this notion is ultimate; its speculative treatment falls altogether—usually under the heading “rational psychology”—to metaphysics.

46. The growth of intellection and self-consciousness reacts powerfully upon the emotional and active side of mind. To Conduct. describe the various sources of feeling and of desire that thus arise—aesthetic, social and religious sentiments, pride, ambition, selfishness, sympathy, &c.—is beyond the scope of systematic psychology, and certainly quite beyond the limits of an article like the present. But at least a general résumé of the characteristics of activity on this highest or rational level is indispensable. If we are to gain any oversight in a matter of such complexity it is of the first importance to keep steadily in view, as a fundamental principle, that as the causes of feeling become more complex, internal, and representative the consequent actions change in like manner. We have noted this connexion already in the case of the emergence of desires, and seen that desire in prompting to the search for means to its end is the primum movens of intellection (cf. § 35). But intellect does much more than devise and contrive in unquestioning subservience to the impulse of the moment, like some demon of Eastern fable; even the brutes, whose cunning is on the whole of this sort, are not without traces of self-control. As motives conflict and the evils of hasty action recur to mind, deliberation succeeds to mere invention and design. In moments of leisure, the more imperious cravings being stilled, besides the rehearsal of failures or successes in the past, come longer and longer flights of imagination into the future. Both furnish material for intellectual rumination, and so we have at length (1) concepts of general and distant ends, as wealth, power, knowledge, and—self-consciousness having arisen—that concept also of the happiness or perfection of self, and (2) maxims or practical generalizations as to the best means to these ends. Instead of actions determined by the vis a tergo of blind passion we have conduct shaped by what is literally prudence or foresight, the pursuit of ends that are not esteemed desirable till they are judged to be good. The good, it is truly urged, is not to be identified with the pleasant, for the one implies a standard and a judgment, and the other nothing but a bare fact of feeling; thus the good is often not pleasant and the pleasant not good; in talking of the good, in short, we are passing out of the region of nature into that of character. It is so, and yet this progress is itself so far natural as to admit of psychological explication. As already urged (§ 34), the causes of feeling change as the constituents of consciousness change; also they depend more upon the form of that consciousness as this increases in complexity. When we can deliberately range to and fro in time and circumstances, the good that is not directly pleasant may indeed be preferred to what is only pleasant while attention is confined to the seen and sensible; but then the choice of such good is itself pleasant—pleasanter than its rejection would have been.

The mention of deliberation brings us to the perennial problem of “the freedom of the will.” But to talk of will is to lapse into Freedom. the confusions of the old faculty—psychology. As Locke long ago urged: “The question is not proper, whether the will be free, but whether a man be free.”[122] In the absence of external constraint, when a man does what he likes, we say he is “externally free”; but he may still be the slave of every momentary impulse, and then it is said that he is not “internally” free. The existence and nature of this internal freedom is the problem. But if such freedom is held to imply a certain sovereignty or autonomy of self over against momentary propensions and blind desires, there can obviously be no question of its existence till the level of self-consciousness is reached and maxims or principles of action are possible. The young child, the brute and the imbecile, even when they do as they like, have not this freedom, though they may be said to act spontaneously. A resolutely virtuous man will have more of this freedom than the man of good moral disposition who often succumbs to temptation; but it is equally true that the hardened sinner has more of it than one still deterred in his evil ways by scruples of conscience. A man is internally free, then, whenever the ends he pursues have his whole-hearted approval, whether he say with Milton's Satan, “Evil be thou my good,” or with Jesus, “Thy will be done.” But this freedom is always within our experience a relative freedom; hence at a later time we often declare that in some past act of choice we were not our true selves, not really free. But what is this true self more than our ideal? Or perhaps we prefer to say that we were free and could have acted otherwise; and no doubt we might, if the place of the purely formal and abstract concept of self had been occupied by some other phase of that empirical self which is continuously but at no one moment completely, presented. It must then be admitted that psychological analysis in this case is not only actually imperfect, but must always remain so—so long, at any rate, as all that we discern by reflection is less than all we are. But this admission does not commit us to allowing the possible existence of a liberum arbitrium indiferentiae, sometimes called “absolute indeterminism”; for that would seem to differ in no respect from absolute chance or caprice. On the other hand, the rigidly determinist position can only be psychologically justified by ignoring the activity of the experiencing subject altogether. At bottom it treats the analysis of conduct as if it were a dynamical problem pure and simple. But motives are never merely so many quantitative forces playing upon something inert, or interacting entirely by themselves. At the level of self-consciousness especially motives are reasons and reason is itself a motive. In the blind struggle of so-called “self-regarding” impulses might is the only right; but in the light of principles or practical maxims right is the only might.[123] This superiority in position of principles is only explicable by reference to the inhibitory power of attention, which alone makes deliberation possible and is essentially voluntary; that is, subjectively determined. But no, it may be objected, deliberation in such cases is just the result of painful experiences of the evil of hasty action, and only ensues when this motive is strong enough to 'restrain the impulse that would otherwise prevail. Even if this be granted, it does not prove that the subject's action is determined for and not by him; it merely states the obvious fact that prudence and self-control are gradually acquired. Authoritative principles of action, such as self-love and conscience, are no more psychologically on a par with appetites and desires than thought and reason are on a par with the association of ideas.

Relation of Body and Mind.

47. The question of subjective initiative leads us naturally to that concerning the connexion of mind and organism, to which Parallelism. we now proceed. In development and efficiency, in the intensity and complexity of their processes, mind and brain keep invariably and exactly in line together. Striking and impressive instances of this correspondence are to be found in comparative psychology, and especially in mental pathology; but it is needless here to enlarge on a point which in the main is beyond dispute. In this correspondence lay the plausibility of the old materialism. But a closer scrutiny discloses an equally impressive disparity: we reject materialism, accordingly, while still maintaining this psychoneural parallelism to be a well-established fact. From this we must distinguish a second sense of parallelism founded on the disparity just mentioned as pertaining to the psychical and neural correlates. We may call this physiologico-psychological, or, more briefly, methodological, parallelism. It disclaims as illogical the attempt to penetrate to psychical facts from the standpoint of physiology, so persistently and confidently pursued by the old materialists. It also forbids the psychologist to piece out his own shortcomings with tags borrowed from the physiologist. The concepts of the two sciences are to be kept distinct, as the facts themselves to which they relate are distinct. Confusion is inevitable if the psychologist, for example, talks of his volition as the cause of his arm moving, when by arm movement he means the process described by the physiologist in terms of efferent excitations, muscular flexions, and so forth; or if the physiologist speaks of a sensation of red as produced by retinal stimulation due to light waves of a certain length, when by sensation he means what he immediately experiences on looking at a field poppy. This methodological convention, as we may call it, implies a more stringent interpretation of causation than that expounded by J. S. Mill and his contemporaries. It does not, however, forbid psychological inferences on the basis of physiological facts, nor vice versa. But in spite of this distinctness of the facts, and of the standpoints from which they are respectively studied, their causal relation cannot be simply ignored: it is, however, a problem that pertains strictly to the higher standpoint of philosophy. There have been in all four different theories of this relation within modern times: (1) that of mutual interaction—the common-sense view—very inconsistently maintained by Descartes; (2) the “occasionalism” substituted for this by Geulincx and the later Cartesians; (3) the pre-established harmony of Leibnitz; and (4) the monism of Spinoza, which reduced matter and mind to parallel attributes of the One Substance. The last of these—severed, however, from Spinoza's metaphysics—is still perhaps the prevailing theory, and to it the term psychophysical parallelism most properly applies. For whereas the parallelism first mentioned states a real correspondence between psychical processes and neural processes, but leaves open the question of a possible interaction between matter and mind, modern psychophysical parallelism is a pure hypothesis concerning the relation of psychical facts to physical theories, on the ground of which—as we shall presently see—any interaction between matter and mind is expressly denied.

But in the exposition of this hypothesis these two meanings of parallelism are frequently confused or interchanged. The same term “body” is applied both to an aggregate of matter and to the living organism. Now life must be regarded as either inherent in matter, or as the result simply of a particular material configuration, or as physically inexplicable. But, for the present at all events, it cannot be explained physically; nor are we even within measurable distance of such an explanation: so much is beyond cavil. Yet the hypothesis of psychophysical parallelism confines us to one or other of the former alternatives: at the same time its unwarrantable identification with psychoneural parallelism—where we find a real correspondence between mind and organism—tends to conceal the gravity of such assumptions. The standpoint of physiology, therefore, must be described not as identical with that of physics, but as intermediate between it and the standpoint of psychology. If the fact of life could be reduced to physical terms, physiology then, no doubt, would have to fall into line with physics, much as chemistry, for example, may have had to do. On the other hand, till a physical explanation of life is forthcoming, physiology belongs, with psychology, to the biological group of sciences, and cannot divest itself completely of the teleological concepts essential to them, not a vestige of which belongs to bare physics. It is just because of this community in their concepts that there actually is a certain “point to point” correspondence or parallelism between the psychical and the neural: as an organ a neuron is a unit; physically regarded, it ceases to be one. Yet this illicit identification of organism and material body is thought to be legitimate, inasmuch as physiological processes are found to rest invariably on a physical basis: and inasmuch as, though methodological parallelism forbids the physiologist to identify psychosis with neurosis, no limits can be imposed on his efforts to ascertain the mechanism of the neurosis itself. But if this be granted, is not psychophysical parallelism justified, in principle at all events? By no means: as little, for example, as an explanation of the mechanism of a locomotive would justify us in ascribing its origin, its maintenance or its guidance to the machine itself. When life and mind are explained by their mechanism the physicist may summon the biologist, as Mephistopheles did Faust, “Her zu mir”: then, but not before.

A favourite mode of stating psychophysical parallelism is that known as the Double Aspect Theory. In this, besides “Double Aspect” Theory. the unjustified identification of the first and third meanings, we find also an equally unjustified interpretation of parallelism in the second sense. All that methodology prescribes is that psychologists and neurologists—and, we may add, that physicists too—shall severally, as “specialists,” mind their own business. Again, all that the first two jointly ascertain is simply the fact of correspondence: the explanation of it is still to seek. Two propositions are now advanced which are held to meet this need. First—and negatively—the connexion, it is said, is not causal: mind does not act on body, nor body on mind: the changes on each side form two independent series, each “going along by itself.” In other words, the series themselves are said to exemplify what methodology enjoins on the sciences that investigate them—they mind their own business and never intrude into each other's domains. Nevertheless their interaction is not prima facie contradictory or absurd, and ordinary thought, as we have seen, assumes that it exists. What evidence, then, is there for denying it absolutely? Empirical evidence for such a universal negative there can hardly be; it must be established therefore—if established at all—on a priori grounds. Meanwhile two facts, already noticed, make seriously against it. On the psychical side sensations point to an intrusion of some sort, and are not psychically explicable (cf. § 16), and the like—for the present at all events—must be said of the fact of life on the physical side. Apart from all this, it seems plain that methodological parallelism, so far from justifying the denial of interaction, simply precludes its discussion on the dualistic level to which that parallelism is confined. The gulf implied is indeed not absolute—of so much, parallelism in the first sense assures us—but those who are forced to keep to their own side of it obviously are not the people to settle how it is crossed. We are aware that the dualism is not absolute, it is replied: it is only phenomenal, and the two series of phenomena are conditioned by an underlying unity of substance. Such is the second, and positive, proposition of the theory. Again asking for evidence, we are told that this underlying unity is unknown—in fact, unknowable. This unknowable substance is assumed, then, simply because—the impossibility of causal connexion being taken as established—no other alternative remains. The negative proposition is thus the foundation of the theory, and without it this agnostic monism becomes entirely arbitrary. We have, therefore, to continue our search for the grounds on which the possibility of interaction is denied. But it will be worth while first to examine certain ambiguities besetting the positive statement.

Difference of aspect may result solely from difference of standpoint, or it may be due to difference in the reality itself. The circle, seen as concave from within and as convex from without is an ancient instance of the first still in great favour; the pillar that was cloud and darkness to the Egyptians, but light to the children of Israel, may serve to exemplify the second. The former we may call the phenomenal, and the latter the ontal, meaning of “aspect.” With these two very different meanings our theory plays fast and loose, as suits its own convenience. To do this is easy—in so far as the reality is unknown and unknowable; and necessary—since in the end, the reality, however unknowable, must somehow include both the phenomenal aspects and all that pertains to them, and so far therefore be known. In dealing with “aspect” in the first sense, the one question to be raised concerns the nature and relation of the respective standpoints. To one belongs what we know as individual experience, and this is essentially concrete, immediate, and qualitatively diverse; to the other belongs an abstract, conceptual scheme, wholly quantitative, familiarly known as the mechanical theory. Between these there is plainly no such co-ordination as the inept comparison with the inside and the outside of a circle implies.[124] Neither is there, on the other hand, the same complete opposition; for the entire mechanical theory is based upon individual experience as enlarged and developed by inter-subjective intercourse. Both the sense-knowledge of the one and the thought-knowledge of the other relate to the one objective factor involved in both. So far, then, there is fundamentally only one standpoint—that of the subjective factor to the objective factor, which is immediately perceived in the one and mediately conceived in the other. The question here raised is thus primarily epistemological, but it is a question, as we have seen, in which psychology is intimately concerned. “Aspect” in the second sense is independent of standpoints. We have here to deal with attributes of the one reality, more or less in Spinoza's sense: this reality itself, as possessed of disparate attributes, is so far dual, and the question of causal connexion between these attributes is not escaped. For to know that a thing has invariably two distinct attributes does not enable us to determine straightway how the changes or “modes” of the one are connected with those of the other. (1) The same attribute might be always the initiating or independent variant, and then would come the question of finding out which of the two it was; or (2) it might be that now one, now the other, took the lead, the grounds of this alternation being then the topic for inquiry; or, finally, (3) it might be, as our theory assumes, that there was but a single series of double changes. The questions here raised are philosophical questions, but again they are questions in which psychology is intimately concerned. Our examination thus yields two results: first, there is fundamentally only a single standpoint—that of experience, now at the perceptual, now at the conceptual, level; and secondly, the distinction of aspects is not merely phenomenal, but pertains “somehow” to reality. The question is how; and this leads us to resume our inquiry into the grounds on which interaction is denied.

These grounds neither pertain to psychology nor to physiology. In spite of the outstanding difficulties connected with sensation and life, which these sciences severally raise, such denial is upheld mainly on the strength of an interpretation of the principle known as the conservation of energy—an interpretation of it, however, which many of the ablest physicists disallow. The energy of the physical world, it is maintained, is a strictly invariable amount; matter, therefore, cannot act on mind, for such action would entail a decrease, nor can mind act on matter, since that would entail an increase, of this energy. In other words, the material world is held to be a “closed system”; and as all the changes within it are mass-motions, there can be none which are not the effect and equivalent of antecedent mass-motions. But now this statement must be established on physical grounds: to assume it otherwise would be openly to beg the very question at issue. For if mind does act on matter, the physical mechanism is subject to changes from without, and so often its motions are not due to antecedent motions; and this—the common-sense view—cannot, of course, be summarily dismissed as impossible or absurd. Now, energy is essentially a metrical notion, and its conservation in finite and isolated material systems has been ascertained by careful quantitative experiments. To say that the energy of the material universe is constant is only a way of expressing the generalization of this result—is tantamount, in other words, to saying that it holds of all finite isolated systems. The whole universe may perhaps be called isolated, but we do not know that it is finite. We cannot, therefore, apply metrical concepts to it; and consequently we cannot interpret the conservation of energy as meaning that the physical part of it is a closed system. But if not a closed system, then the energy of a given group of bodies may be increased or decreased without interaction between that group and other bodies—may be increased or decreased by psychophysical interaction, that is to say. And, moreover, such psychophysical interaction would not invalidate the conservation of energy, rightly understood; for that merely means that the energy of a group of bodies can be altered only from without, and this might happen whenever such interaction occurred.[125] We seem, therefore, justified for the present in rejecting psychophysical parallelism as one of the three possible modes of relating mind and matter regarded as attributes of the real. Not only are there psychological as well as biological objections which it has not yet overcome, but there are so far no physical grounds in its favour.

At this point we may again for a moment turn aside to consider a modified form of the doctrine—the so-called Conscious “Conscious Automaton” Theory. Automaton Theory, an attempt to blend the old Cartesian views concerning the minds of man and brute. According to Huxley,[126] the best known modern exponent of this theory, “our mental conditions are simply the symbols in consciousness of the changes that take place automatically in the organism.” This consciousness is supposed “to be related to the mechanism of the body simply as a collateral product of its working, and to be as completely without any power of modifying that working as the steam-whistle . . . is without influence upon the locomotive's machinery”: thus “the feeling we call volition is not the cause of a voluntary act, but the symbol of that state of the brain which is the immediate cause of that act.” In other words, physical changes are held to be independent of psychical, whereas psychical changes are declared to be their “collateral products.” They are called collateral products, or “epiphenomena,” to obviate the charge of materialism, and to conform to the interpretation of the conservation of energy that we have just discussed. Such a theory is, strictly speaking, one of parallelism no longer: rather it adopts, instead, the first of the two possibilities we have noted above as opposed to parallelism. According to it, matter is the initiating or independent variant, on whose changes mind simply follows suit. It is open to two fatal objections. First, it is methodologically unsound: its psychology is physiological in the bad sense. It regards all states of consciousness as passive, i.e. as ultimately either “feelings” or “reflexes.” Volitional activity is declared illusory; and if this be true, intellectual activity must be illusory too. But to detect illusion requires experience of reality—we only know the sham by knowing the genuine first; and even passive states could not be experienced as such save by contrast with states that are active. To the physical side, then, we naturally turn for this knowledge which we are told is not to be found on the psychical; and we do so the more readily as, according to the present theory, the physical holds the primary place. But we turn in vain; for matter is inert, and its energy only “works” by taking the line of least resistance, like water running down hill. Moreover, such activity as we are in search of could only be found here in case the physical mechanism showed signs of being intelligently directed, and that would also be evidence that psychical activity is not illusory. Is, then, the physical side after all primary? No, we reply: the assumption is epistemologically unsound. This is our second objection. The order implied in the distinction of physical phenomena and psychical epiphenomena is contrary to all experience and indefensible. A physical phenomenon is either actually perceived or possibly perceptible; otherwise it is devoid of empirical reality altogether. But objects of perception are so far psychical; that is, they belong to immediate or individual experience. Therefore we cannot regard them as independent of this experience, nor this as their collateral product, i.e. as epiphenomena. Again, the phenomenality supposed to be common to both involves, as we have already seen, a fundamental identity in the standpoint of each: they belong to the same continuous experience at different levels. And lastly, their abstract, merely quantitative, character shows that it is the concepts of physics, and not the facts of immediate experience, that are symbolic, and so to say epithetic. The attempt—either empirically or speculatively—to outflank mind by way of matter is an absurdity on a par with getting into a basket in the hope of being able to carry oneself.

These epistemological considerations may help us to deal with the prime and ultimate argument for strict parallelism. When all is said and done, it is urged, still the interaction of mind and matter remains inconceivable. But this is hardly a sufficient reason for denying what is prima facie a fact. Occasionalists, from Geulincx to Lotze, have acknowledged the same obscurity in all cases of transeunt action. Yet they did not venture to deny that sensations were interruptions in the psychical series, the “occasions” for which were only to be found in the physical; nor that purposive movements were interruptions in the physical series, the “occasions” for which were only to be found in the psychical. And surely such a position is more in harmony with experience than that of the parallelisms, who maintain that each series “goes along of itself”—a statement which, as we have repeatedly urged, contradicts psychology and assumes the physical “explanation” of life. Whereas occasionalism leaves the question of ultimate means to be dealt with by a metaphysics which will respect the facts,[127] parallelism forecloses it on the basis of a ready-made metaphysics—modern naturalism, that is to say—in which psychology as an independent science is entirely ignored. Starting with a dualism as absolute as that of Descartes—but replacing his two substances by one, enjoying the otium cum dignitate of the Unknowable—starting, too, from the physical side, no wonder such a philosophy finds that what is for us the most familiar and of the supremest interest, the concrete world of sense and striving, is for it the altogether inconceivable, the supreme “world riddle.” And yet if the naturalist could deign to listen to the plainest teachings of psychology and of epistemology, the riddle would seem no longer insoluble, for his phenomenal dualism and his agnostic monism would alike disappear. The material mechanism which he calls Nature would rank not as the profoundest reality there is to know: it would rather become—what indeed “machine” primarily, connotes—an instrumentality subservient to the “occasions” of the living world of ends; and so regarded, it would cease to be merely calculable, and would be found intelligible as well. Psychophysical parallelism, then, we conclude, is not a philosophically tenable position; and—pending the metaphysical discussion as to the ultimate nature of interaction generally—we have to rest content with the second of the three possible modes of connexion above defined, as occasional ism formulates it. According to this, the two series, the psychical and the physical, are not independent and “closed” against each other; but in certain circumstances—e.g. in perception—physical changes are the occasion of psychical, and in certain circumstances—e.g. in purposive movements—psychical changes are the occasion of physical: the one change not being explicable from its psychical antecedents, nor the other from its physical.

Into the metaphysical discussion we cannot, of course, enter here. It must suffice to say that it will not be conducted on the lines of our present inquiry: it will not start from a dualism of matter and mind, either regarded as substances or as phenomena. Its problem will rather be the interaction of subject and object—a duality in the unity of experience, which by no means coincides with the dualism of matter and mind, neurosis and psychosis, and the like.

Comparative Psychology

48. Psychoneural parallelism is no doubt a well established generalization; nevertheless, concerning its exact range and its precise meaning there are differences of opinion. It is applicable, every one will allow, so soon as there is evidence of experiences individually acquired (cf. § 3); and from such point onwards, in ascending any biological phylum, we find that the psychical and neural aspects differentiate and develop together. But how when we descend? Interpreting the neural correlate physiologically, and not morphologically, as referring primarily to function and not to structure, we find that even in unicellular organisms it is still present as irritability and conductivity (leading to contraction, secretion, &c.). But as at higher levels psychosis is correlative to neurosis, the principle of continuity would seem to justify us in assuming a like correspondence here. Moreover, “learning by experience,” the comparative psychologists criterion, obviously presupposes some antecedent and underlying process, of which it is the differentiation and development. And our general analysis of mind, if correct, enables us to describe this process—“the irreducible psychical minimum,” of which we are here in search. We have such complete psychosis—and it is the simplest we know—in the emotional or diffused movements that follow immediately upon sensation; and these are so far purposive—though not intentional—that they tend to heighten or retain what is pleasurable, and to alleviate or remove what is painful. Given that plasticity, which is the psychological presupposition of all acquisition, then learning by experience is a possible development from such a primitive stage.

But though every psychosis have its concomitant neurosis, it is uncertain how far the converse holds good. The action of the heart, for example, depends upon neuroses of which we have now no direct consciousness. Facts of this kind have led to three hypotheses concerning the lowest forms of life, differing more or less from that just proposed. (i.) Perfectibility and instinct are found, it is said, to be in inverse ratio. Hence in the lowest forms of life there is no “learning by experience,” because a stationary state of complete adjustment to environment has been already attained, and all reactions have therefore become “secondarily automatic”: consciousness, having served its purpose, has disappeared. To such a very Buddhistic psychology it may be objected: (1) that even organic refiexes tell upon the so-called vital sense or coenaesthesis, and so far—the irreducible minimum being still intact—do not preclude all possibility of learning, should occasion arise; and (2) that the psychical life, even of a Protozoan, does not, according to the best evidence, show any such mechanical finality as is here supposed.[128] (ii.) According to the second view, which is advocated by Herbert Spencer, the behaviour of the lower organisms is wholly made up of such reflexes, supposed to be devoid of all psychical concomitants; but consciousness—so far from having disappeared—first comes upon the scene at the opportune moment when the increasing complexity of the mechanism calls for its guidance. Psychologically this hypothesis is less defensible than the last, and it has already been dealt with at some length (cf. § 7). It not only assumes, as that does, far more uniformity in the interaction of organism and environment than the facts warrant, but in regarding life as prior to mind, and as the means of its evolution, it burdens science with two insoluble problems instead of one. For even if it were possible chemically to build up protoplasm, we should still be as far from organisms as a heap of bricks are from putting themselves together as a house. (iii.) The last view we have to notice is essentially an extension of the preceding, and is chiefly interesting as a reductio ad absurdum of that. The physics of colloidal substances—at present wanting, but confidently expected “in the near future” by certain biologists—is the key which is to unlock the mysteries of protoplasm. Certain organisms, regarded as varieties of such a substance, react positively to a given physical property of the environment, and others negatively: thus a moth flies towards the light, and a centipede runs from it—the one is positively, the other negatively, “heliotropic”; the radicle of a seed, growing downwards, is, positively, the plumule, growing upwards, is, negatively, “geotropic.” Instincts are but complexes of such tropisms, and owe their character entirely to the symmetrical form and definite structure of the colloidal substance. Now if it facilitate the work of the biologist to say that when what we ordinarily regard as a hungry caterpillar climbs to the tip of a branch it is forced so to do by positive heliotropism; that then positive chemiotropism sets up mastication of the young buds; and that, lastly, “we can imagine this process leading to the destruction of the substances in the skin of the animal that are sensitive to light, and upon which the heliotropism depended,” so leaving it free to crawl downwards and come in contact with the new buds which have in the meantime unfolded[129]—if such language serve any useful purpose, all well and good; only it must be applied to the hungry man too: in short, all behaviour must be described in the same terms. For the champion of colloids to betake himself to consciousness as he approaches the higher forms of life is as much a breach of methodological parallelism as it is for the psychologist to fall back upon protoplasm as he approaches the lower. But to suppose that psychical processes first appear in the complicated form of association of ideas—which learning by experience is taken to imply—and at the same time to assume that such experience, even when it appears, is “ultimately due to the motions of colloidal substances,” these are incongruous absurdities which only the grossest ignorance would be bold enough to maintain.

Concluding, as we have done, that mind and matter—as we may provisionally call them—do really interact, we naturally infer that organic structures are not the result solely of material processes, but involve the co-operation of mental direction and selection: in other words, we are led to regard structure as partly shaped and perfected by function, rather than function as solely determined by structure, itself mechanically evolved. And such a view is justified by the fact that mechanical evolution is primarily a process of “degradation” rather than development, a case of facilis descensus contrasting with the upward struggle of life per aspera ad astra. Still, the notion of life or mind as formative and directive has its difficulties. In the first place, we have no experience of mind organizing matter-no experience of the actual process, that is to say-however sure we may feel of the fact.[130] Hence the occasionalism to which here, at any rate, science is confined. But even so, the difficulty is not wholly removed. In the handicrafts whence we derive the conception of organs the artificer handles, but does not literally order, his tools—as if they too were intelligent. The conscious direction of such movements is doubtless facilitated by the fact that many of the complex co-ordinations actually involved in them are carried out automatically, thanks to structural modifications, either inherited or acquired. And, regarding life phylogenetically, we can imagine this process carried back indefinitely. Indeed, if it be illogical to talk of mechanisms evolving themselves and giving rise to the beings whose ends they serve, we have no choice but to accept this dualism of mind-shaping and matter inert. No choice, that is, unless we can establish the primacy of the psychological standpoint. Here we have duality but not dualism, and the object is not inert, i.e. is not matter. But still there remain two difficulties—possibly resolvable into one—the plasticity already referred to as involved in all biological development and hereditary transmission; as to these, psychology is almost wholly in the dark.[131]

Authorities.—Historical: There are few good works on the history of psychology; the only one in English, R. Blakey, History of the Philosophy of Mind from the Earliest Period to the Present Time (London, 1848), is poor. F. A. Carus's Geschichte der Psychologie (Leipzig, 1808) is at least useful for reference. A work bearing the same title by H. Siebeck (the first art consisting of two divisions—(i.) Die Psychologie von Aristoteles, (ii.) Die Psychologie von Aristoteles bis zu Thomas von Aquino (Gotha, 1880 and 1884) is thoroughly and carefully done. Siebeck has also contributed a series of articles, “Zur Psychologie der Scholastik,” to the Archiv f. d. Gesch. d. Philos. (vols. i.-iii.). Die Philosophie in ihrer Geschichte (I. Psychologie), by Professor Harms (Berlin, 1878), is also good. T. A. Ribot's La Psychologie anglaise contemporaine (3rd ed., 1892) and La Psychologie allemande contemporaine (2nd ed., 1885) are lucid and concise in style, though the latter work in places is superficial and inaccurate. Of Max Dessoir's Geschichte der neueren deutschen Psychologie the section dealing with the 17th-century writers prior to Kant went into a second edition in 1897; it contains a useful collection of material. From Les Origines de la psychologie contemporaine (2nd ed., 1908), by the neo-Thomist scholar Mgr. D. Mercier, much may be learnt, though its purpose is not primarily historical.

Positive: The recent output of systematic works on psychology has been voluminous. Among the most important of these may be mentioned J. Sully's The Human Mind (2 vols., 1892); W. James, Principles of Psychology (2 vols., 1890); G. F. Stout, Analytic Psychology (2 vols., 1896); A Manual of Psychology (2nd. ed., 1901); H. Höffding, Outlines of Psychology (1891; translated from the Danish); G. T. Ladd, Psychology, Descriptive and Explanatory (1894); W. Wundt, Grundriss der Psychologie (4th ed., 1901, translated); F. Jodl, Lehrbuch der Psychologie (2 vols., 2nd ed., 1902). Dealing mainly with experimental psychology are: Külpe, Grundriss der Psychologie auf experimenteller Grundlage dargestellt (1893; translated); Ebbinghaus, Grundzüge der Psychologie (3rd ed., 1908), Bd. I.; and E. B. Titchener, Experimental Psychology: a Manual of Laboratory Practice (2 vols., 1901); C. S. Myers, Experimental Psychology (1908).

Of the older more advanced textbooks Professor Volkmann's Lehrbuch der Psychologie (2 vols., 3rd ed., 1885; edited by Cornelius) is written in the main from a Herbartian standpoint. To the honoured name of Lotze belongs a distinguished place in any enumeration of modern productions in philosophy; his Medicinische Psychologie (Göttingen, 1852) is still valuable. A large part of his Mikrokosmos (3 vols., 3rd ed., 1876-1880; trans. into English, 2 vols., 1885) and one book of his Metaphysik (2nd ed., 1884 also trans. into English) are, however, devoted to psychology. The doctrine of evolution has been as fruitful in this study as in other sciences that deal with life. In this respect Herbert Spencer's Principles of Psychology (2 vols., 3rd ed., 1881) and Data of Ethics (1879) occupy a foremost place. Dr Alexander Bain's standard volumes, The Senses and the Intellect (4th ed., 1894) and The Emotions and the Will (3rd ed., 1875), contain a good deal of “physiological psychology,” but no adequate recognition of the importance of the modern theory of development. Wundt's Physiologische Psychologie (3 vols., 6th ed., 1908 seq.) is indispensable to the student of this subject.

Specially interesting as treating psychological problems on new lines are La Psychologie des idées-forces, by A. Fouillée (2 vols., 1893)—perhaps the best French contribution to recent psychology; its cardinal point is the fundamentally dynamical character of the psychical. R. Avenarius, Kritik der reinen Erfahrung (2 vols., 1888-1890; 2nd ed., 1908), is an attempt, on the model of Kirchhoff and Mach's treatment of physics, to describe experience, taking the relation of the central nervous system to the environment as starting point. Its strange and forbidding terminology prevented the timely recognition of its merits; but since the author's death in 1896—from overwork and disappointment—quite a literature has grown up, partly expository, partly controversial; devoted to this latest critique. H. Cornelius, Psychologie als Erfahrungswissenschaft (1897), rather epistemological than psychological, claims affinity with the critiques of Kant and Avenarius. In J. Rehmke's Lehrbuch der allgemeinen Psychologie (2nd ed., 1905)—a psychology with a soul, and claiming to be philosophy as well—the problems of perception and of psychoneural interaction are discussed at length. F. Brentano, Psychologie vom empirischen Standpunkte (1874), vol. i., treats presentations and judgments as fundamentally distinct, feeling and willing, on the other hand, as fundamentally one. His influence on Austrian psychologists has been considerable, and is more or less apparent in the following: K. Twardowski, Zur Lehre vom Inhalt und Gegenstand der Vorstellungen (1894); A. Meinong, Psychologisch-ethische Untersuchungen zur Werththeorie (1894), and also numerous important papers; v. Ehrenfels, System der Werththeorie (2 vols., 1897-1898); A. Höfler, Psychologie (1897).

Important as treating of particular topics are C. Stumpf, Tonpsychologie (2 vols., 1883-1890); A. Lehmann, Die Hauptgesetze des menschlichen Gefühlsleben (trans. from the Danish; 1892); various monographs by T. A. Ribot on diseases of memory, will, personality, on the psychology of attention, of the emotions, of general ideas, &c., all translated into English; J. M. Baldwin, Social and Ethical Interpretations in Mental Development (1897); W. Wundt, Völkerpsychologie (3 vols., 1900); W. McDougall, An Introduction to Social Psychology (1908).

There are several periodicals devoted exclusively to psychology, the chief being the American Journal of Psychology; the Psychological Review; Zeitschrift für Psychologie und Physiologie der Sinnesorgane; L'Année psychologique; the British Journal of Psychology; and Archiv für die gesammte Psychologie.

(J. W.*)

  1. Principles of Nature and Grace, § 3.
  2. Lectures on Metaphysics, ii. 153.
  3. And it is precisely for want of this mediation that Kant's “two stems of human knowledge, which perhaps may spring from a common but to us unknown root,” leave epistemology still more or less hampered with the old dualism of sense and understanding.
  4. Evolution of Sex, by Geddes and Thomson, 1st ed. p. 265.
  5. Cf. Darwin, Descent of Man, i. 56.
  6. T. H. Huxley, Hume, “English Men of Letters Series,” (1879), p. 171.
  7. Huxley, op. cit. p. 172.
  8. Examination of Sir W. Hamilton's Philosophy, ch. xii. fin.
  9. A meaning better expressed, as said above, by experience.
  10. Cf. Kant's Principle of the Anticipations of Perception: “In all phenomena the real, which is the object of sensation, has intensive magnitude.”
  11. The biological principle referred to is that known as von Baer's law, viz. “that the progress of development is from the general to the special.”
  12. Compare Spencer's Principles of Psychology, i. §§ 217, 8.
  13. D. Hartley, Observations on Man (6th ed., 1834), pp. 66 sqq.
  14. It may be well to call to mind here that Alexander Bain also regarded emotional expression as a possible commencement of action, but only to reject it in favour of his own peculiar doctrine of “spontaneity,” which, however, is open to the objection that it makes movement precede feeling instead of following it—an objection that would be serious even if the arguments advanced to support his hypothesis were as cogent as only Bain supposed them to be. Against the position maintained above he objects that “the emotional wave almost invariably affects a whole group of movements,” and therefore does not furnish the “isolated promptings that are desiderated in the case of the will” (Mental and Moral Science, p. 323). But to make this objection is to let heredity count for nothing. In fact, wherever a variety of isolated movements is physically possible there also we always find corresponding instincts, “that untaught ability to perform actions,” to use Bain's own language, which a minimum of practice suffices to perfect. But then these suggest gradual ancestral acquisition.
  15. Foster, Text-Book of Physiology, § 597.
  16. 16.0 16.1 To cover more complex cases we might here add the words “or trains of ideas.”
  17. “That there are ideas, some or other, always present in the mind of a waking man, every one's experience convinces him; though the mind employs itself about them with several degrees of attention. Sometimes the mind fixes itself with such intention . . . that it shuts out all other thoughts and takes no notice of the ordinary impressions made on the senses; . . . at other times it barely observes the train of ideas . . . without directing and pursuing any of them; and at other times it lets them pass almost quite unregarded as faint shadows that make no impression” (Essay, ii. 19, §§ 3, 4).
  18. As, e.g. in interpreting the conduct of children as if they were already “grown-up” persons; cf. J. Ward, Il. of Spec. Phily. (1882), pp. 369 fin. 374; James, Prin. of Psychy. (1890), i. 196.
  19. The Works of Thos. Reid, supplementary note, p. 932.
  20. Other languages give more prominence to this distinction; compare γνῶναι and εἲδέναι, noscere and scire, kennen and wissen, connaître and savoir. On this subject there are some acute remarks in a little-known book, the Exploratio philosophica, of Professor J. Grote. Hobbes, too, was well awake to this difference, as e.g. when he says, “There are two kinds of knowledge; the one, sense or knowledge original and remembrance of the same; the other, science or knowledge of the truth of propositions, derived from understanding.”
  21. Bain, Logic, i. 3.
  22. Common language seems to recognize some connexion even here or we should not speak of harsh tastes and harsh sounds, or of dull sounds and dull colours and so forth. All this is, however, superadded to the sensation, probably on the ground of similarities in the accompanying organic sensations.
  23. Physiologische Psychologie, 1st ed., p. 421; the doctrine reappears in later editions, but no equally general statement of it is given.
  24. The following brief passage from his Principes de la nature et de la grace (§ 4) shows his meaning: “Il est bon de faire distinction entre la Perception, qui est l'état intérieur de la Monade représentant les choses externes, et l'Apperception, qui est la Conscience, ou la connaissance réflexive de cet état intérieur, laquelle n'est point donnée à toutes les âmes, ni toujours à la même âme. Et c'est faute de cette distinction que les Cartésiens ont manqué, en comptant pour rien les perceptions dont on ne s'apperçoit pas, comme le peuple compte pour rien les corps insensibles” (Op. Phil. Erdmann's ed., p. 715).
  25. Herbart and Fechner describe subconscious presentations generally as existing below the threshold. On the other hand, we have spoken of subconscious sensations as existing beyond it. In view of the important differences between the two forms of presentations primary and secondary, this distinction of ultra-liminal and subliminal seems convenient and justifiable.
  26. This doctrine of the involution and evolution of ideas we owe to Leibnitz. Herbart attempted in a very arbitrary and a priori fashion to develop it into a physical statics and dynamics with the result—usual to extreme views—that later psychologists neglected it altogether. There are now signs of a fresh reaction, and we shall continually come across evidence of the wide range and great importance of the doctrine as we proceed.
  27. Examination of Sir W. Hamilton's Philosophy, 3rd ed., p. 329.
  28. For a detailed account of the various sensations and perceptions pertaining to the several senses the reader is referred to the articles Vision; Hearing; Touch; Taste; Smell, &c.
  29. Senses and Intellect, 4th ed. (1894), p. 101.
  30. Nothing shows this more plainly than the newly-coined term epiphenomenon now applied in this connexion.
  31. Reception does not in English suggest the taking back of the Latin recipere; it expresses only the comparative passivity of sense. In contrast to percipere (to take entire possession of) it implies the absence of that assimilation which is essential to perception; and finally it contrasts appropriately with retention.
  32. This distinction, though continually overlooked, is vitally important. By psychological analysis we mean such analysis as the psychological observer can reflectively make, by psychical analysis only such analysis as is possible in the immediate experience of the subject observed.
  33. Cf. G. H. Lewes, Problems of Life and Mind (1879), vol. iii. pp. 250 sqq.; H. Spencer, Principles of Psychology, vol. i. § 60.
  34. Cf. W. A. Nagel, “Die Phylogenese specifischer Sinnesorgane,” Bibliotheca zoologica (1894), pp. 1-42.
  35. Cf. Stumpf, Tonpsychologie, ii. 58 seq.
  36. As a matter of fact there are no objects absolutely black, none that are devoid of all lustre and completely absorbent of light. But this does not affect the argument.
  37. It is assumed that the physiological differentiation of the retina has advanced from the centre, where vision is most distinct, towards the margin where it is least so; and it is found that stimulation of the margin yields none but achromatic sensations, stimulation of a certain intermediate zone only sensations of yellow or blue, and central stimulation alone sensations of every hue. Further, total colour-blindness is extremely rare, whereas red-green colour-blindness is comparatively common.
  38. Cf. Bastian, The Brain as an Organ of Mind (1880), pp. 691 sqq.; Ferrier, The Functions of the Brain (1886), 2nd ed. pp. 382 sqq.; James, Principles of Psychology (1890), ch. xxvi.
  39. We are ever in danger of exaggerating the competence of a new discovery; and the associationists seem to have fallen into this mistake, not only in the use they have made of the concept of association in psychology in general, but in the stress they have laid upon the fact of movement when explaining our space-perceptions in particular. Indeed, both ideas have here conspired against them—association in keeping up the notion that we have only to deal with a plurality of discrete impressions, and movement in keeping to the front the idea of sequence. Mill's Examination of Hamilton (3rd ed., p. 266 seq.) surely ought to convince us that, unless we are prepared to say, as Mill seems to do, “that the idea of space is at bottom one of time” (p. 276), we must admit the inadequacy of our experience of movement to explain the origin of it.
  40. To illustrate what is meant by different complexes it will be enough to refer to the psychological implications of the fact that scarcely two portions of the sensitive surface of the human body are anatomically alike. Not only in the distribution and character of the nerve-endings but in the variety of the underlying parts—in one place bone, in another fatty tissue, in others tendons or muscles variously arranged—we find ample ground for diversity in “the local colouring” of sensations. And comparative zoology helps us to see how such diversity has been developed as external impressions and the answering movements have gradually differentiated an organism originally almost homogeneous and symmetrical. Between one point and another on the surface of a sphere there is no ground of difference; but this is no longer true if the sphere revolves round a fixed axis, still less if it also runs in one direction along its axis.
  41. The improvements in the sensibility of our “spatial sense” consequent on practice, its variations under the action of drugs, &c., are obviously no real contradiction to this; on the contrary, such facts are all in favour of making extensity a distinct factor in our space experience and one more fundamental than that of movement.
  42. Thus a place may be known topographically without its position being known geographically, and vice versa.
  43. Cf. on this point Poincaré, La Science et l'hypothèse, pp. 74 sqq.
  44. Thus Locke says, “Our simple ideas [i.e. presentations or impressions, as we should now say] are all real . . . and not fictions at pleasure; for the mind can make to itself no simple idea more than what it has received” (Essay, ii. 30, 2). And Berkeley says, “The ideas imprinted on the senses by the Author of Nature are called real things; and those excited in the imagination, being less regular, vivid and constant, are more properly termed ideas or images of things, which they copy or represent” (Prin. of Hum. Know., pt. i. § 33).
  45. The distinction between the thing and its properties is one that might be more fully treated under the head of “Thought and Conception.” Still, inasmuch as the material warrant for these concepts is contained more or less implicit in our percepts, some consideration of it is in place here.
  46. Ideation—“a word of my own coining,” says James Mill.
  47. Treatise of Human Nature, bk. i. pt. i. § 1.
  48. Moreover, as we shall see, the distinction between present and past or future psychologically presupposes the contrast of impression and image.
  49. Organic sensations, though distinguishable from images by their definite though often anatomically inaccurate localization, furnish no clear evidence of such adaptations. But in another respect they are still more clearly marked off from images, viz. by the pleasure or pain they directly occasion.
  50. The following scant quotation from Fechner, one of the best observers in this department, must suffice in illustration. “Lying awake in the early morning after daybreak, with my eyes motionless though open, there usually appears, when I chance to close them for a moment, the black after-image of the white bed immediately before me and the white after-image of the black stove-pipe some distance away against the opposite wall. . . . Both [after images] appear as if they were in juxtaposition in the same plane; and, though—when my eyes are open—I seem to see the white bed in its entire length, the after-image—when my eyes are shut—presents instead only a narrow black stripe owing to the fact that the bed is seen considerably foreshortened. But the memory-image on the other hand completely reproduces the pictorial illusion as it appears when the eyes are open” (Elemente der Psychophysik, ii. 473).
  51. Die Lehre vom Tastsinne, &c., pp. 86 seq.
  52. As we have seen that there is a steady transition from percept to image, so, if space allowed, the study of hallucinations might make clear an opposite and abnormal process—the passage, that is to say, of images into percepts, for such, to all intents and purposes, are hallucinations of perception, psychologically regarded.
  53. On this term cf. below, §§ 24, 28.
  54. Cf. Drobisch, Empirische Psychologie (1842), § 31; Höffding, “Ueber Wiederkennen, Association und psychische Activität,” in Vierteljahrsschr. f. wissenschaftl. Philosophie, Bd. xiii. and xiv. To Höffding we are also indebted for the term Bekanntheitsqualität, which has suggested the γ character used above. Cf. also Ward, “Assimilation and Association,” Mind (1894-1895).
  55. Recent experiments, however, seem to prove that the after-percept is not the sole factor, and often is not a factor at all in such successive comparison (so-called); but that what is now termed “the absolute impression” may supplement it or even replace it Eltogether. As to what is meant by absolute impression, cf. 14, c.
  56. Hence the earlier process has been named “impressional association” (Stout, Analytic Psychology, 1896, ii. pp. 27-29), and again “animal association” (Thorndike, Animal Intelligence, an Experimental Study of the Associative Processes in Animals, 1898, pp. 71, 87, and passim). But it seems preferable to confine the term “association” to the later process, in which alone the component presentations have that amount of distinctness and individuality which the term properly connotes.
  57. Some light is perhaps here thrown on the reciprocal relation of “association by contrast” and “association by similarity” as severally the differentiation of partial similars and the integration of partial dissimilars.
  58. J. Steiner, Die Functionen des Centralnervensystems u.s.w., 2te Abth. Die Fische (1888), pp. 50, 126, 19 seq., 101.
  59. W. Bateson, “The Sense-Organs and Perceptions of Fishes,” Journ. Marine Biol. Assoc. (1890), p. 239.-5
  60. Cf. Stout, Manual of Psychology (1899), vol. ii. ch. i.; also F. H. Bradley, “Memory and Inference,” Mind (1899), pp. 145 sqq.; and especially Thorndike, Animal Intelligence, cited above.
  61. So Hume, Treatise of Human Nature, pt. i. § 4 (Green and Grose's ed., p. 321); also Lotze, Metaphysik, 1st ed., p. 526.
  62. Experience-continuum would perhaps be a better name, since it is only a preliminary to a true memory record, as we shall presently see.
  63. This connexion of association with continuous movements of attention makes it easier to understand the difficulty above referred to, viz. that in a series A B C D . . . B revives C but not A, and so on—a difficulty that the analogy of adhesiveness or links leaves unaccountable. To ignore the part played by attention in association, to represent the memory-continuum as due solely to the concurrence of presentations, is perhaps the chief defectmof the associationist psychology, both English and German. Spencer's endeavour to show “that psychical life is distinguished from physical life by consisting of successive changes only instead of successive and simultaneous changes” (Principles of Psychology, pt. iv. ch. ii., in particular pp. 403, 406) is really nothing but so much testimony to the work of attention in forming the memory-continuum, especially when, as there is good reason to do, we reject his assumption that this growing seriality is physically determined.
  64. A term borrowed from Lotze (Metaphysik, 1st ed., p. 295), but the present writer is alone responsible for the sense here given to it and the hypothesis in which it is used.
  65. Apart, that is to say, of course, from the reduplications of the memory-train spoken of below.
  66. This contrast of thread and tissue is suggested, of course, by Herbart's terms Reihe and Gewebe. It is justified by the fact that memory proper follows the single line of temporal continuity, while ideation furnishes the basis for manifold logical connexions.
  67. It is a mark of the looseness of much of our psychological terminology that facts of this kind are commonly described as cases of association. Dr Bain calls them “obstructive association,” which is about on a par with “progress backwards”; Mr Sully's “divergent association” is better. But it is plain that what we really have is an arrest or inhibition consequent on association, and nothing that is either itself association or that leads to association.
  68. Any full discussion of paramnesia, as these very interesting states of mind are called, belongs to mental pathology.
  69. As, e.g. James Mill (Analysis of the Human Mind, ch. x.), who treats this difficult subject with great acuteness and thoroughness.
  70. Cf. W. James, Principles of Psychology, i. 629 sqq.; L. W. Stern, “Psychische Präsenzzeit,” Z. f. Psych., (1897), xiii. 325 sqq.
  71. To this rate the “indifference point” mentioned above is obviously related. It has also been called “adequate time” or “optional time.” It is, however, a tempo that varies with the subject-matter attended to; when effective attention is more difficult the tempo is slower than it is when to attend is easy.
  72. Cf. Wundt, Logik, i. 432.
  73. H. Ebbinghaus, “Ueber das Gedächtniss: Untersuchungen zur experimentellen Psychologie” (1885).
  74. Cf. J. Jacobs and F. Galton on the “Span of Prehension,” Mind, (1887), pp. 75 sqq.; Bourdon, “Influence de l'âge sur la mémoire immédiate,” Rev. phil. (1894) xxxviii., 148 sqq.
  75. Cf. Dietze, “Untersuchungen über den Umfang des Bewusstseins u.s.w.,” Phil. Studien (1885), pp. 362 sqq.; L. W. Stern, “Psychische Präsenzzeit,” Ztschr. f. Psychologie (1897), xiii. 325 sqq.; Daniels, “Memory After-image and Attention,” Am. Jour. of Psychology (1893), vi. 558 sqq.
  76. W. G. Smith, “The Place of Repetition in Memory,” Psychological Rev. (1896), pp. 20 sqq. The figures given are unquestionably low, partly, as the writer points out, in consequence of the method employed, but partly, as his detailed tables show, in consequence of the lax attention of three out of his eight subjects. Objections have been taken to the plan of this investigation, but it is doubtful if they invalidate the result here mentioned. Cf. Jost, “Die Associationsfestigkeit in ihrer Abhängigkeit von der Vertheilung der Wiederholungen,” Ztschr. f. Psychologie, xiv. 455 sqq.
  77. The following are among the more important papers on rhythm: T. L. Bolton, “Rhythm,” Am. Journ. of Psychology (1894), pp. 145 sqq.; E. E. Meumann, “Untersuchungen z. Psychologie u. Aesthetik des Rhythmus,” Phil. Studien (1894), x. 249 sqq., 393 sqq.; M. K. Smith, “Rhythmus und Arbeit,” Phil. Studien (1900), xvi. 71 sqq. 197 sqq.; Arbeit und Rhythmus (1899), by K. Bucher, a well-known economist, bringing out the teleological aspects of rhythm.
  78. There are still other forms of what seems at first sight to be regressive association, but none that do not admit of explanation without this assumption.
  79. See Spencer, Data of Ethics, chs. i.-iv.; G. H. Schneider, Freud und Leid des Menschengeschlechts, ch. i.
  80. In the lowly organisms that absorb food directly through the skin such bitter juices as exist naturally might at once produce very violent effects—comparable, say, to scalding; and the reflexes then established may have been continued by natural selection so as to save from poisoning the higher organisms, whose absorbent surfaces are internal and only guarded in this way by the organ of taste. Some light is thrown on questions of this kind by the very interesting experiments of Dr Romanes; for a general account of these see his Jelly-fish, Star-fish, and Sea-urchins, ch. ix.
  81. This is one among many cases in which the study of a vocabulary is full of instruction to the psychologist. The reader who will be at the trouble to compare the parallel columns under the heading “Passive Affections,” in Roget's Thesaurus of English Words and Phrases, will find ample proof both of this general statement and of what is said above in the text.
  82. Observation and experiment show that the physical signs of pain in the higher animals consist in such changes as a lowered and weaker pulse, reduction of the surface temperature, quickened respiration, dilatation of the iris, and the like. And so far as can be ascertained these effects are not altogether the emotional reaction to pain but in large measure its actual accompaniments, the physical side of what we have called its tone. The following is a good description of these general characteristics of feeling: “En même temps, il se fait une série de movements généraux de flexion, comme si l'animal voulait se rendre plus petit, et offrir moins de surface à la douleur. Il est intéressant de remarquer que, pour l'homme comme pour tous les animaux, on retrouve ces mêmes mouvements généraux de flexion et d'extension répondant aux sentiments différents de plaisir et de la douleur. Le plaisir répond à un movement d'épanouissement, de dilatation, d'extension. Au contraire, dans la douleur, on se rapetisse, on se referme sur soi; c'est un movement général de flexion” (C. Richet, L'Homme et l'Intelligence: la douleur, p. 9).
  83. It has been definitely formulated, but in physiological language, by Bain as the Law of Novelty: “No second occurrence of any great shock or stimulus, whether pleasure, pain, or mere excitement, is ever fully equal to the first, notwithstanding that full time has been given for the nerves to recover from their exhaustion” (Mind and Body, p. 51). Cf. also his Emotions and Will, 3rd ed., p. 83.
  84. Preyer, Akustische Untersuchungen, p. 59.
  85. Cf. Fechner, Vorschule der Aesthetik, ii. 263. Fechner's full style for it is “Princip der ökonomischen Verwendung der Mittel oder des kleinsten Kraftmasses.”
  86. Essays, Scientific, Political and Speculative, vol. ii., Ess. I. and VIII.
  87. As it is impossible to say that any distinguishable presentation is absolutely simple, the hypothesis of subconsciousness would leave us free to assume that any pleasantness or unpleasantness that cannot be explained on the score of intensity is due to some obscure harmony or discord, compatibility or incompatibility, of elements not separately discernible. But this, though tempting, is not really a very scientific procedure. If a particular presentation is pleasurable or painful in such wise as to lead to a redistribution of attention, it is reasonable to look for an explanation primarily in its connexion with the rest of the field of consciousness. Moreover, it is obvious—since what takes place in subconsciousness can only be explained in analogy with what takes place in consciousness—that, if we have an inexplicable in the one, we must have a corresponding inexplicable in the other. If the feeling produced by what comports itself as a simple presentation cannot be explained by what is in consciousness, we should be forced to admit that some presentations are unpleasant simply because they are unpleasant—an inexplicability which the hypothesis of subconsciousness might push farther back but would not remove.
  88. “To look at anything in its elements makes it appear inferior to what it seems as a whole. Resolve the statue or the building into stone and the laws of proportion, and no worthy causes of the former beautiful result seem now left behind. So, also, resolve a virtuous act into the passions and some quantitative law, and it seems to be rather destroyed than analysed, though after all what was there else it could be resolved into?” Sir A. Grant, Aristotle's Ethics, Essay IV., “The Doctrine of the Mean,” i. 210 (2nd ed.).
  89. Mind (1884), ix. 188 sqq.; and, again, Principles of Psychology, ch. xxv. Very similar views were advanced independently and almost at the same time by the Danish physiologist C. Lange; hence the name James-Lange theory, by which their views are commonly known. Of Lange's work a German translation was published in 1887.
  90. “Physical Basis of Emotion,” Psychological Review (1894), p. 518. In this reply to criticisms Professor James is supposed to have modified his views: it would be nearer the truth to say that he has made admissions incompatible with them.
  91. Text-Book of Psychology (1890), p. 383.
  92. G. H. J. Berkley, “Two Cases of General Cutaneous and Sensory Anaesthesia without marked Psychical Implications,” Brain (1891), xxv. 441 sqq.
  93. “Experiments on the Value of Vascular and Visceral Factors for the Genesis of Emotion,” Proc. Roy. Soc. (1900), lxvi. 390 sqq.; and Nature, lxii. 328 sqq.
  94. Of the three principles Darwin advances in explanation of emotional expression that which he places last—perhaps because it admits of less definite illustration—seems both psychologically and physiologically more fundamental than the more striking principle of serviceable associated habits which he places first; indeed the following, which is his statement of it, implies as much: “Certain actions which we recognize as expressive of certain states of mind are the direct result of the constitution of the nervous system, and have been from the first independent of the will, and to a large extent of habit” (Expression of the Emotions, p. 66). It is in illustration of this principle too that Darwin describes the movements expressive of joy and grief, emotions which in some form or other are surely the most primitive of any.
  95. As such an instance may be cited Plato's story of Leontius, the son of Aglaeon, in Rep. iv. 439 fin.
  96. Bain, Emotions and Will, 3rd ed., p. 438.
  97. It must be noted that, though we still retain our psychological standpoint, the higher development of the individual is only possible through intercourse with other individuals, that is to say, through society. Without language we should be mutually exclusive and impenetrable, like so many physical atoms; with it each several mind may transcend its own limits and share the minds of others. As a herd of individuals mankind would have a natural history as other animals have; but personality can only emerge out of intercourse with persons, and of such intercourse language is the means. But important as is this addition of a transparent and responsive world of minds to the dead opaqueness of external things, the development of our psychological individual still remains a purely individual development. The only new point is—and it is of the highest importance to keep it in sight—that the materials of this development no longer consist exclusively of presentations elaborated by a single mind in accordance with psychical laws. Nevertheless that combination of individual experiences which converts subjective idiosyncrasy and isolation into the objectivity and solidarity of Universal Mind only affects the individual in accordance with psychical laws, and we have no need therefore to overstep our proper domain in studying the advance from the non-rational phase to the phase of reason.
  98. Locke, so often misrepresented, expressed this truth according to his lights in the following: “The earth will not appear painted with flowers nor the fields covered with verdure whenever we have a mind to it. . . . Just thus is it with our understanding: all that is voluntary in our knowledge is the employing or withholding any of our faculties from this or that sort of objects and a more or less accurate survey of them” (Essay, iv. 13, 2).
  99. Ruskin, in his Fors clavigera, relates that the sight of the word “crocodile” used to frighten him when a child so much that he could not feel at ease again till he had turned over the page on which it occurred.
  100. Treatise of Human Nature (Green and Grose's ed.), pt. i. § vii. p. 331.
  101. Cf. Berkeley, Principles of Human Knowledge, Introd. § 16, Hume, op. cit. § 7.
  102. See Wundt, Logik, i. 107 seq., where this process is happily styled “die kategoriale Verschiebung der Begriffe.”
  103. Cf. Hamilton: “To judge (κρίνειν, judicare) is to recognize the relation of congruence or of confliction in which two concepts, two individual things, or a concept and an individual, compared together, stand to each other” (Lectures on Logic, i. 225).
  104. As to these it must suffice to refer to what has been already said; cf. § 11 and § 28.
  105. Cf. e.g. F. Schumann, “Zur Psychologie der Zeitanschauung,” Ztschr. f. Psychologie, xvii. 130, 136.
  106. G. F. Stout, Analytic Psychology, i. 50 seq.
  107. A. Meinong, “Ueber Gegenstände höherer Ordnung u.s.w.,” Ztschr. f. Psychologie (1899), xxi. 182 sqq. Special mention must be made of an earlier paper by C. v. Ehrenfels (“Ueber Gestaltqualitätes,” Vierteljahrsschr. f. wissensch. Philosophie, 1890, pp. 249 sqq.), round which the whole subsequent discussion of this topic centres. Cf., too, Stout, op. cit. bk. i. ch. iii.
  108. Assuming, of course, that the change is the simplest or directest possible, i.e. a change of “colour proper” without change of saturation.
  109. Essay concerning Human Understanding, II. xvi. § 1.
  110. “We may regard one of the words here printed as one, in that by a definite act we unite a plurality of letters in our image and separate it from its neighbours: we may also regard the one word as many when we attend to the transition from one letter to another and mark each step” (Sigwart, Logic, ii. § 66).
  111. Wundt, Logik; eine Untersuchung der Principien der Erkenntniss (2nd ed., 1893), i. 59 sqq.
  112. Hume's numerical difference, that is to say, is really distinctness, not quantitative difference.
  113. Cf. J. S. Mill, Logic, bk. i. ch. iii. § 11, and Examination of Hamilton, 3rd ed., ch. xiv. p. 306, note; also Meinong, “Hume-Studien” II., Wiener Sitzungsberichte (Phil. Hist. Cl.). ci. 709.
  114. Treatise of Human Nature, pt. iii. § xiv., “of the idea of necessary connexion.”
  115. Hume here has Locke and Berkeley specially in view. Locke as a patient and acute inquirer was incomparably better as a psychologist than a man addicted to literary foppery like Hume, for all his genius, could possibly be. On the particular question, see Locke, Essay, bk. ii. c. 21, §§ 3-5.
  116. In an article (Mind, 1886, p. 317) Mr F. H. Bradley created some stir by declaring that “the present use of these phrases [active energy] is little better than a scandal and a main obstacle in the path of English psychology.” In Mind for 1902 and 1903 he has made important contributions towards clearing up the supposed confusion, and the subject is still being debated. But the main contention of the text, that activity is for psychology at all events ultimate and unanalysable, seems still to await refutation. A brief notice of some of the diverse views obtaining will be found in an address, “The Problems of General Psychology,” by J. Ward, Philosophical Review (1904), pp. 608 sqq.
  117. Treatise of Human Nature, Green and Grose's ed., i. 396.
  118. “Bases affectives de la personnalité” in Revue philosophique, xviii. 149.
  119. This subject has a very wide literature. The following are specially interesting: Ribot, Les Maladies de la personnalité (3rd ed., 1889); Boris Sidis and S. P. Goodhart, Multiple Personality (1905); Morton Prince, The Dissociation of a Personality (1906).
  120. So—misled possibly by the confusions incident to a special faculty of reflection, which they controvert—James Mill, Analysis, i. 224 seq. (corrected, however, by both his editors, pp. 227 and 230), and also Hamilton, Lect. i. 192.
  121. It has been thought a fatal objection to this view that it implies the possibility of an indefinite regress; but why should it not? We reach the limit of our experience in reflection, or at most in deliberate introspection, just as in space of three dimensions we reach the limit of our experience in another respect. But there is no absurdity in supposing a consciousness more evolved and explicit than our self-consciousness, and advancing on it as it advances on that of the unreflecting brutes.
  122. Essay concerning Human Understanding, II. xxi. §§ 16 sqq.
  123. The right is only relative, of course, when the maxims are “hypothetical”—to use Kant's phrase,—but it is absolute when the maxim is “categorical.”
  124. In fact, if there were, since it is only as we contemplate finite portions of the circle that the distinction of concave and convex is present, the nearer we approximated to its elements the more this difference of aspect would disappear. If on the physical side we called these elements atoms, there would be an answering element of “mind-stuff” on the psychical; and there would be no more unity and no other diversity in a given man's mind than in his brain regarded as a complex of primordial atoms. Wild as all this seems, yet views of the kind have been seriously put forward more than once as the logical outcome of psychophysical parallelism.
  125. The possibility is enough: we cannot tell what actually happens, and do not, therefore, know how far the direction of matter by mind calls for a modification or limitation of physical hypotheses. Cf. Ward, Naturalism and Agnosticism (3rd ed., 1906), ii. 73-86.
  126. Essay on “Animal Automatism,” Collected Essays, vol. i.
  127. Cf. Lotze, Metaphysik, § 61 fin.
  128. Cf. H. S. Jennings, Behaviour of the Lower Organisms (1906).
  129. Cf. J. Loeb, Comparative Psychology (1901), pp. 188 sqq.—an interesting book, full of psychological crudities.
  130. But of course a thoroughgoing spiritualism ought to explain the very existence 'of matter as really the appearance or manifestation of mind.
  131. On the subject of comparative psychology generally, see Animal Behaviour (1900), by Professor C. Lloyd Morgan; L. T. Hobhouse, Mind in Evolution (1901).