You are on page 1of 291

J. M.

Williams

Luminance and the Hess Effect

Luminance-Dependent Latency as Measured by the Hess Effect


by John Michael Williams Copyright (c) 1980, 2011, John Michael Williams (jmmwill@comcast.net, as of 2013-02-07) Permission is granted to reproduce one PDF copy and one hard copy of this document, including the Preface, for personal and private, noncommercial use. All other rights are reserved.

Preface to the PDF Online Version


The original 1980 doctoral dissertation manuscript was typed double-spaced in hardcopy on a typewriter. The result was copied by Southern Illinois University and bound in letter-size, single-sided paper. Parts of it were rewritten and published in the journal, Vision Research, in 1983, in an article coauthored with Professor Alfred Lit, the author's doctoral dissertation advisor: Williams, J. M., & Lit, A. (1983) "Luminance-dependent visual latency for the Hess effect, the Pulfrich effect, and simple reaction time". Vision Research, 23, 171-179. The current, PDF, version was composed by scanning the original manuscript on a UMAX S-12 scanner at 600 dpi to TIFF bitmaps. The result then was converted to Open Office Writer v. 3.2 format. Pages which were predominantly text were converted using the Tesseract OCR program. Nontext (usually graphical) pages were imported as bitmap figures. The Writer document then was rectified by hand; this entailed considerable manual editting, especially for formulas, tables, equations, and references. Page layout was modified slightly in this process, although pagination generally was preserved. Figures were combined with figure captions, although these were on separate pages in the dissertation hardcopy manuscript. Formulas sometimes were reformatted to fit on one line. A few minor errors, mostly of the typo variety, have been corrected, and a couple of minor additions were made in the introductory matter. Words originally typed underlined for emphasis generally were put in italics, often bolded. The 2011-01-01 version corrects a few newly-found typoes and improves the formatting of some equations. The original dissertation ms is available from University Microfilms, Ann Arbor, Michigan (Abstract #8213538, 1980). A web site dedicated to the Pulfrich, Hess, and related effects, with explanations, literature, and animations, is maintained by the author at www.siu.edu/~pulfrich.

J. M. Williams

Luminance and the Hess Effect

LUMINANCE-DEPENDENT LATENCY AS MEASURED BY THE HESS EFFECT

by John Michael Williams B. A., Columbia University, 1969 A. M., University of Chicago, 1976

A Dissertation Submitted in Partial Fulfilment of the Requirements for the Degree of Doctor of Philosophy

Department of Psychology in the Graduate School Southern Illinois University August, 1980

J. M. Williams

Luminance and the Hess Effect

ii

J. M. Williams

Luminance and the Hess Effect

ACKNOWLEDGEMENTS The author particularly thanks Professor Alfred Lit for many discussions on the present subject, and for comments on a preliminary draft of this Dissertation. Mr. Charles J. Popp and Mr. John R. Holstrom designed and built the machinery of the target-movement mechanism in about 1973, using Eye Institute funds from a USPHS Research Grant awarded to Dr. Lit. A FORTRAN program used for daily Special

datum-analysis was written by Mr. Arthur Menendez.

thanks are due Mr. Popp for much valuable advice and help during the two years it took to install and calibrate the apparatus as described herein. The present study was supported in part by the National Eye Institute of USPHS through a Research Grant (EY00383) awarded to Professor Alfred Lit. The author held a Graduate

Fellowship and subsequently a Dissertation Research Award during the period of the study.

iii

J. M. Williams

Luminance and the Hess Effect

TABLE OF CONTENTS Page INTRODUCTION ..................................................... 1 1. Spatiotemporal Illusions ...................................... 8 2. Simple Reaction Time ......................................... 13 3. The Hess Effect .............................................. 16 4. Validity of the Geometric t Formula ......................... 19 VISUAL IATENCIES 1. The Visual System ............................................ 26 2. Mechanisms of Visual Latency ................................. 32 3. Intensity-Dependent Visual Latency: Literature ............... 47 4. Selected Theories of Intensity-Dependent Latency ............. 48 THE HESS EFFECT 1. Literature ................................................... 54 2. Proposed Theory .............................................. 66 EXPERIMENTAL METHOD 1. Subjects ..................................................... 87 2. Apparatus a. Reaction Time and Pulfrich Apparatus ...................... 87 b. Hess Target-Movement Mechanism ............................ 89 c. Hess Lightbox ............................................. 97 d. Calibrations (1) Photometry ........................................... 103 (2) Optical Filters ...................................... 106

iv

J. M. Williams

Luminance and the Hess Effect

(3) Hess Target Relative Displacement .................... 106 (4) Hess Target Speed .................................... 108 3. Procedure ................................................... 109 a. Reaction Time ............................................ 109 b. Pulfrich Effect .......................................... 109 c. Hess Effect .............................................. 110 d. Stimulus Intensity Range ................................. 113 RESULTS 1. Reaction Time ............................................... 115 2. Pulfrich Effect ............................................. 116 3. Hess Effect ................................................. 118 4. Least-Squares Approximate Fits .............................. 123 5. Latency Comparisons for Individual Subjects ................. 128 6. Miscellaneous Observations .................................. 129 DISCUSSION ..................................................... 132 TABLES ......................................................... 137 FIGURES ........................................................ 192 BIBLIOGRAPHY ................................................... 224 APPENDICES 1. Literature on Intensity-Dependence of Latency ............... 258 2. Normal Equations for Selected Latency Theories .............. 269 3. Solution of Linear First-Order Ordinary Differential Equation ................... 280

J. M. Williams

Luminance and the Hess Effect

VITA SHEET ..................................................... 283

vi

J. M. Williams

Luminance and the Hess Effect

GUIDE TO TABLES Subject Matter Lightbox Design ......................... Reaction Time ........................... Reaction Time vs. Hess t ............... Pulfrich Effect ......................... Table(s) I II-VI VII VIII-XII Page(s) 137 138-142 143 144-148 149 -156, 185 157-173, 185
174-185

LD Hess Effect .......................... XIII-XX, XLVI AB Hess Effect .......................... JMW Hess Effect ......................... Selected Hess-Effect Standard Deviations ................ Combined Hess-Effect Data for Theoretical Fit ................ Fit of Various Theories to Hess-Effect Data ................... XXI-XXXVI, XLVI XXXVII-XLVI XLVII-XLIX L-LI LII

186-188 189-190 191

vii

J. M. Williams

Luminance and the Hess Effect

INTRODUCTION In the traditional associationistic view, "sensations" somehow were organized by the mind into "perceptions," the latter being mental responses to objects or other stimuli (e.g., Boring, 1950, ch. 25; Allport, 1955, pp. 46-50: Dodwell, 1978). The

sensations were considered the afferent mental events mediated by the various senses; these senses usually were described as the the ancient five Aristotelian senses, which correspond to readily recognizable anatomical units. Before the present century, in a typical elaboration of the "five senses," the Aristotelian sense of touch, for example, would be replaced by distinct senses of taction, kinaesthesis, warmth, cold, and pain. A given sense would be considered as

such to be a legitimate objective source of sensations, such sensations having a sort of independent existence somewhere within the mind, body, or both. A major advance at the dawn of

modern psychology was Wundt's realization that introspection was an appropriate tool of analysis in dealing with sensations in the laboratory. But laboratory methods made the shortcomings of

introspection obvious, and mind yielded its place to the more demonstrably valid matter. During the nineteenth and early twentieth centuries, precise use of the terms sensation and perception filled a gap in the theory and in the rhetoric of psychology while concrete progress was being made in understanding the stimulus-response mechanisms.

J. M. Williams

Luminance and the Hess Effect

For example, hand-in-hand with the new touch "sensations" named in the preceding paragraph, new anatomical entities were being recognized such as Meissner and Pacinian corpuscles, end-bulbs of Krause, and the free nerve-endings. The structure of the

"mind" grew to be resolvable in terms of an experimental datum-base of neural pathways, reflex reaction times, complication factors in personal equation, and, eventually, conditional reflexes. As agreement was achieved in the laboratory in terms

more operational than those of the armchair, the old sensation-perception dialectic gradually fell into disuse. Careful measurement and the use of operational definition have replaced the mind and its "common sense" (of Aristotle and others) with an incredibly complex but somewhat more comprehensible central nervous system (CNS), the "sensations" of mind being replaced by membrane permeabilities, neural impulses, neurotransmitters, and similar quantifiable concepts. The old

five senses survive only as a nosological convenience (e.g., Hyvarinen, 1973). The resultant replacement of ambiguous terminology has had a salubrious influence in the design of psychological inquiry: In the past, vitalistic "sensations" often were invoked to excuse the capricious results of experiments in which "independent" variables were allowed, inadvertently, to depend on responses of the subject. Even the "soul" was used occasionally as an un-

known factor to excuse faulty control of experimental variables which now can be measured accurately but in previous years could

J. M. Williams

Luminance and the Hess Effect

not be described or standardized for replication. For these reasons, in most frequent usage, "sensations" as analytic tools have yielded their place to the more hardy "perceptions" -- the latter still being definable usably as hypothetical or often overt responses to experimentally-manipulable stimuli (cf. Heinemann, 1978). The terms sensation and perception still

are convenient and are used (in fact interchangeably) when referring to a response in a afferent system (Graham, 1951, p. 870 note) In the present work, the "perception lag" of Roufs (1963), for example, will not be considered as referring to anything distinguishable from the "sensation" lags of Bills (1920), Grossman (1967, p. 201), or others. However, responses occurring after the "per-

cept" has been formed probably ought not be referred to as sensations; such later responses probably best would not be considered perceptions, either, unless they derived meaningfully from stimulus inputs occurring later than the inputs upon which the percept in question be purported to depend. An example might be worthwhile, Suppose that tachis-

although the terminology be of moot value:

toscopic localization of an object depended on a percept forming in the cerebral cortex, whereas recognition of that object depended on a percept forming in the brainstem. If localization

always occurred first, then recognition never could be considered a sensation or perception relevant to localization; conversely, localization could be considered a sensation or perception with respect to some response measuring recognition. If experimental

manipulations permitted localization to be retarded (as a particular,

J. M. Williams

Luminance and the Hess Effect

identifiable perceptual response) so that it followed recognition as defined above, only then could it be that recognition might function as a perceptual component of localization. The concept of latency as used in the present work derives from the physiological concept of the latent period between a stimulus and a response (Michaelis, 1963, p. 764). Although a

distinction can be drawn between a latent period (before a response occurs) and an implicit time before a response peak or other feature can be recognized by the experimenter, latency in the present work will refer to a time lapse between stimulus and some specific event (response) within the organism, regardless of whether or not some response occurring earlier than that event also might be identified. Examples in the literature of usage of

"latency" and "implicit time" may be found in Bartley (1934), Johnson and Bartlett (1956), Rietveld and Tordoir (1965), and van der Tweel, Estevez, and Cavonius (1979). By the definition adopted in the present work, during a latent period nothing seems to be happening, at least not as measured by the response criterion or recording instrument being used. A physiological example is given in Zierler (1974, p. 86):

If a brief shock be delivered to an excised frog muscle, no measurable change in tension of the muscle is observed for about 1.4 ms; a decrease in tension then occurs and typically persists 1.8 ms or more; and finally the sought response, a twitch contraction, actually begins. would be 1.4 + l.8 = 3.2 ms. The total latent period in this example Similarly, responses in the afferent

J. M. Williams

Luminance and the Hess Effect

systems do not occur instantaneously but must follow the stimulus onset after various delays, the delay depending on the response being measured. The delay, or latency, for any particular res-

ponse feature is assumed to occur consistently and repeatably under adequately controlled conditions; this assumption is what allows experimental investigation of latency. An analysis of

latencies at various stages of central nervous system processing has been assumed possible from the earliest days of modern psychology (see Poffenberger, 1912). Sufficiently different task re-

quirements in an experiment may be assumed to engage significantly different cell populations in the central nervous system, thereby making a psychophysical analysis of central latencies possible and circumventing the need to insert electrodes or to perform partial dissections of living organisms; this indirect analysis is especially important in research with humans. The visual system is convenient for latency analysis because not only can the two retinas be stimulated entirely independently, but also their lateral halves can be stimulated separately for very different central effects: The decussation

of the optic chiasm sends projections of the left hemiretina of each eye exclusively to the left dorsal lateral geniculate nucleus (LGNd) and thence to the primary visual projection area of the left lobe of the occipital cortex. Thus, retinal stim-

ulation to the left of the foveal fixation point may be depended upon to stimulate directly the left occipital lobe only. The same

situation holds for the right hemiretinas and for the right occip-

J. M. Williams

Luminance and the Hess Effect

ital lobe, of course.

Except possibly for a shared, bilateral-

ly-projecting, ambiguous vertical strip of retina about one degree of visual angle wide and passing through the central fovea, the lateral separation of each hemiretina allows functional analysis of visual response of the two sides of the brain entirely separately--in terms of gross behavioral response, recorded scalp potentials, or any other available method. It might be added here that the bi-

latera1ly projecting one-degree strip just mentioned is a recent finding for the monkey retina (Stone, Leicester, & Sherman, 1973; Bunt, Minckler, & Johanson, 1977) not recognized in Poggio (l974). Additional complications occur in stereopsis, which involves differences in response of the two eyes as such. In terms of the

stimulus alone, the two eyes must image two slightly different views (usually of the same scene) for stereopsis to occur. In the con-

text of the above, central processing for stereopsis must combine differences (e.g., by retinal disparity) in the two images, so therefore the combination(s) should entail some sort of unique perceptual response(s) and associated latencies. Such latencies,

which may be studied by means of the optical illusions discussed below, as well as these discussed by Julesz (1978), should be distinguished from the latencies necessarily involved at individual retinal points when the eyes or the targets viewed are placed in motion. Merely because stereoacuity remains reasonably good

when dichoptic targets are set in lateral motion (Bekesy, 1969; Lit, 1964) does not prove that the retinal latencies associated with that motion are involved in stereopsis. In any event, best

J. M. Williams

Luminance and the Hess Effect

stereoscopic acuity is attained in the foveal one-degree strip (above) which seems to project to both primary visual cortical receiving areas. On a purely hemiretinal basis of analysis, the

stereoscopic combination could be by binocular inhibition at sites as early as the LGNd (see Singer, 1977); probably, as reviewed in Berlucchi (1972), however, stereopsis depends on interhemispheric projections beyond the primary visual projection areas (e.g., Blakemore, l97O; Mitchell & Blakemore, 1970; see also Payne, Elberger, Berman, & Murphy, 1980). Implications of these infer-

ences will be discussed in greater detail below. Before leaving the topic of latency as such, it might be mentioned that the value of accurate latency analysis is recognized in medical diagnosis: Examples would be the hammer-thump and consequent reflexive knee-jerk, the waveform decomposition of an electrocardiogram as correlated with the heart-sounds, and the various metabolic clearance measures of kidney function. The

Pulfrich effect, an illusion discussed below and evidently depending upon visual latency, has been used successfully to diagnose central demyelinating diseases such as multiple sclerosis (e.g., Frisen, Hoyt, Bird, & Weale, 1973; Rushton, 1975). Finally, it might be added here that the visual persistence of a response already evident can be studied in the same circumstances as can its latency (see Bowen, Pola, & Matin, 1974; Sakitt & Long, 1979). Decay-time and persistence would have almost the same

meaning in this context, although ordinarily one would refer to the persistence of some special feature, but to the decay-time of

J. M. Williams

Luminance and the Hess Effect

some given (scalar) measure of a response.

For example,

Bidwell's ghost, an afterimage described in Brown (1965), might be said to persist .2 second because of some particular combination of decay-times of various retinal responses to light. Sakitt and Long (1979) have studied spectral characteristics of persistence of features they believed mediated by different populations of retinal receptors, and the concept of persistence as related to vision and to memory has been reviewed by Coltheart (1980) 1. SPATIOTEMPORAL ILLUSIONS.

Several illusions in vision depend on dichoptic presentation of stimuli such that the time-course of neural activity in one retina (and associated central projections) is made different from the corresponding time-course in the other. The response

to stimulation thus is made sooner or more rapidly for one eye than for the other; this difference in response may be considered to involve both a binocular retinal latency difference and a difference in rise- and decay-time of events at some central perceptual site(s) at which the different responses of the two eyes initiate whatever behavioral or (electro)physiological change is being used to study the illusion in question. Two illusions of the type just described are the Pulfrich effect and the Mach-Dvorak effect. In the context of object-per-

caption (for example, as studied in the vision-laboratory research program at the author's present location), both of these effects involve a single target-object which moves back-and-forth, usually along a straight path, in a horizontal direction parallel to the

J. M. Williams

Luminance and the Hess Effect

observer's frontal plane.

With no illusion present, the target

is seen, as might be expected, to be moving back-and-forth at a single, well-defined distance; this distance may be estimated as seen by asking the observer to set a comparison object to the same distance in depth as the nearest point on the apparent path of the moving target. With no illusion present, such distance-set-

tings measure the localization acuity of the observer under the given experimental conditions (Lit, l960, l964; Lit & Hamm, 1966; Lit & Vicars, 1970; Lit & Finn, 1976). In the illusion of the Pulfrich effect (Pulfrich, 1922; Lit, 1949) as measured by distance settings, a filter placed before one eye makes the target appear to be displaced in depth either toward or away from the observer: While the two images of the

target move across their respective retinas, the latency at any given point in the unfiltered eye evidently remains shorter than that of the corresponding retinal point in the filtered eye; this latency difference advances the neural effect in the unfiltered eye relative to that in the filtered eye. At any given instant,

the same, unique target thus corresponds to two (retinotopically) spatially displaced neural responses (e.g., at the ganglion cell layer of the retina), which responses determine all subsequent responses of the observer including distance-settings. The retino-

topically advanced displacement of the neural effect in the unfiltered eye evidently is processed by the visual system as though it originated from a retinal image disparity. The illusion thus

occurs because the stereoscopically-fused images of the moving

J. M. Williams

Luminance and the Hess Effect

10

target cannot be discriminated from the images of an object actually displaced in depth from the real path of target motion. The ob-

tained direction of apparent displacement of the target from the real path at any given moment is found to depend on which eye is filtered and on the direction in which the target is moving. Julesz

and White (1969) have shown that a filter over one eye can facilitate stereopsis under certain circumstances just as would a delay in stimulus presentation to that eye. In explaining precisely the obtained magnitude and direction of the Pulfrich effect (e.g., Lit, 1949), it can be assumed that, at any given instant, the two eyes are converged to some arbitrary fixation point and that the two images of that fixated point fall directly on the two foveas. A visible object farther away than

the fixated point will be imaged on the two retinas also--but the images of the more distant object will not fall directly upon their respective foveas: Instead, there will be a retinal disparity in

the locations of the images: if the disparity be great enough, the more distant object will be seen "double." If the eyes then be

converged to the distant object just mentioned, the previous fixation point in turn itself may be seen "double," but the newly fiixated object will be seen "single," there being no retinal disparity. Reasoning by this purely geometrical approach, it can

be understood that the retinal disparities of nearby objects must differ (for any given fixation distance) systematically from those of distant objects. It may be assumed that persons with normal

binocular vision learn early in life to interpret these different

J. M. Williams

Luminance and the Hess Effect

11

disparities as discriminative cues to differences in distance of the objects to which the images correspond. Normally, for a given

convergence of the eyes, a moving object at a well-defined, constant distance will be associated with a well-defined, constant retinal disparity of the object's dual images on the two retinas. This

disparity will remain about constant although both images of the object be in motion across the two retinas. But if a filter be

placed before one eye, then, among other things, the response of that eye will become slower than that of the unfiltered eye (evidence for increased visual latency as a function of decreased retretinal illumination is outlined below). Thus, an additional dis-

parity in the two retinal images will be created, as it were, in tenms of the neural responses to the two retinal images. For

constant effect of the filter, and for constant target speed, the effectively added disparity caused by the filter will be constant, too: So, we have s = v t, in which the added effective

disparity s must be constant if target speed v and added filter delay t are constant. For example, in the Pulfrich effect, suppose an object happened to be moving from left to right at a constant distance from the observer. Then no matter what the (fixed) distance to

which the eyes happened to be converged at any given instant, the dual images of the object, reversed by the optics of the eyes, would be moving right-to-left with some particular retinal disparity proper to the real distance between object and observer. Now, if a filter happened to be before the left eye, the response

J. M. Williams

Luminance and the Hess Effect

12

in the right eye effectively would be advanced somewhat; the advance would change the percept so that for judgement of depth, the image in the right eye were advanced farther to the left than otherwise it would have been. This change in disparity would be

the same as the change caused by removing the object to a distance farther away from the observer. Thus, by the geometry, placing

a filter before the left eye would be expected to increase the binocularly perceived distance of an object moving left-to-right--and this is just what is observed. The magnitude of the change

in disparity caused by a filter is found (Lit, l96Oa, b) approximately to be proportional to the speed of the moving object. This finding suggests as above by purely geometrical argument that the given filter must in fact induce a given constant retardation (latency difference); alternative explanations such as constant spatial displacement of retinal response thus are ruled out, although refinements based on target spatial-frequency spectrum (Wist, Brandt, Diener, & Dichgans, 1977) are not. The validity of the geomet-

rical explanation is discussed below, but as an approximation under very general conditions it is reasonably accurate. Applying the

geometrical argument to the classical Pulfrich simple pendulum, it is shown that the pendulum bob will appear to describe an almost-elliptical trajectory in an approximately horizontal plane, the perceived apparent path enclosing the real path and intersecting the latter near the extremes of the target's horizontal traverse; at these extremes, the target speed goes to zero and the direction of target motion is reversed (Pulfrich, l922). Distance-settings,

J. M. Williams

Luminance and the Hess Effect

13

have been used to map the apparent path (Weale, l954; cf. Rogers, Steinbach, & Ono, 1974). In the Mach-Dvorak effect (Lit, 1978; Burr & Ross, 1979), no filter is used. The binocular difference in time-course of

central processing is induced directly by presenting the target only intermittently, but with stimulus onset delayed slightly for one eye relative to the other: A temporal phase-difference in

stimulation thereby has the same effect as an interocular latency difference, either of which may alter perception just as would a binocular spatial disparity. No latency of response is needed

to explain the Mach-Dvorak effect if exposure duration for each eye be brief (Lit, 1978). In the Mach-Dvorak effect, as in the

Pulfrich effect above, if the target should happen to be in simple harmonic motion, it will be seen to move approximately in a horizontal, apparently el1iptica1 path, the departure in depth from the real path at any instant being governed by the direction of target motion and by the eye chosen to receive each stimulus onset first (Dvorak, 1872). Other latency-dependent phenomena have been reviewed in Bills (1920), Vogelsang (1928), and Roufs (1974). 2. SIMPLE REACTION TIME. Simple reaction time depends in part on an inferred perceptual latency but does not require binocular viewing. Other

things being equal, higher stimulus or background luminances are found to produce shorter reaction times (e.g., Teichner & Krebs,

J. M. Williams

Luminance and the Hess Effect

14

1972; see also Menendez, 1979).

Lit (1968) has formulated the

hypothesis that, in general, visual latency is governed by the luminance of the target or the background, whichever has the higher value at stimulus onset. In the simple reaction time paradigm, the observer is asked to make some response as soon as possible after stimulus onset. The stimulus then is presented and the latency to respond measured. The reaction-timed response usually is recorded in apparatus for which the latency-implicit time distinction is not meaningful. The change in the visual field corresponding to stimulus onset may involve changes in illuminance, wavelength, or spectral purity at the retinal location of the target image. Individual differences

in reaction time are related closely to the various "personal equations" (e.g., Wolf, 1865; Sanford, 1888; Stroobant, 1891; Pulfrich, 1922; Boring, 1950, ch. 8; Woodworth & Schlosberg, 1954, ch. 2) which in the past were used in attempts to equate otherwise aberrent observations in astronomy. Simple visual reaction time measurements involve a temporal change in the visual field. For sharp spatial gradients in a

time-invariant visual field, the data are reasonably well explained by Ing1ing and Drum's (1973) hypothesis that chromatic borders are not enhanced by the visual system unless they correspond to luminance borders. Such border enhancement can be seen as the "Mach

band" at each end of a region of gradual luminance change (e.g., Ratliff, 1965; Fiorentini, 1972) or as the bright line visible at an edge (sudden change) in luminance; this latter kind of enhance-

J. M. Williams

Luminance and the Hess Effect

15

ment is found itself to be enhanced by eye movements (Lukas, Tulunay-Keesey, & Limb, 1980). A border between two chromatically

distinct areas in a region in which luminance is constant is not enhanced (van der Horst & Bouman, 1967). ilar in the temporal domain: The situation is sim-

The (heterochromatic) flicker pho-

tometer demonstrates that chromatic fusion occurs at lower temporal frequencies than luminance-fusion provided a luminance-difference be present between the fields being alternated. Thus, it

is possible greatly to enhance flicker of two chromatically distinct alternating fields of equal luminance by changing the luminance of one field slightly. However, flicker is affected slightly

or not at all by changing the dominant wavelength of one field--whether or not the two luminances of the two fields be equal at the start. So, it is not unexpected that the temporal change in the

visual field to which simple reaction time is measured should obey the same laws as does flicker: Simple visual reaction time is

found to depend upon a change in luminance-contrast at target presentation (Pollack, 1968; Lit, Young, & Shaffer, 1971) and on a wavelength-dependent (or possibly a purity-dependent) change in chromatic contrast provided the target contains no luminance tranSient (Weingarten, 1972; Breton. 1977; Nissen & Pokorny, 1977). These last studies show that reaction time to a target matching its background in luminance tends to be long but depends systematically on the target-background color difference.

J. M. Williams

Luminance and the Hess Effect

16

3. THE HESS EFFECT. The Hess effect (Hess, 1904), subject of the present Dissertation, depends on stimulation of one or both retinas by the images of two moving targets (see Figure 2). The two tar-

gets are visibly distinct and the task is (vernier) localization. There is no retinal disparity involved, although geometric explanations of the Hess and Pulfrich effects otherwise are quite similar. Only one retina being necessary, the Hess effect may be

referred to as a monocular illusion; however, the effect allows independent comparison of any two eyes (presumably mainly of any two retinas), including the two eyes of a single observer. comparison might be made to subtract away response biases and reveal retinal disease-states or injuries, if any be present. In any case, the targets used to obtain the Hess effect typically are glass or metal rods, or rectangular pieces of paper, which can be aligned or displaced slightly as in a vernier acuity task. Cathode-ray tubes may be used to present the targets, but speed and phosphor persistence (the latter caused by photoluminescence) can be a problem even at moderate image intensities. In the Hess effect, the targets move together in a direction parallel to their short edges, usually, and with equal velocity in a plane parallel to the observer's frontal plane. The appearance of two such targets on the right side of Figure 2A is indicated by the left side of Figure 2A; stimuli are on the right, and retinal perceptual responses are on the left. The Such

same relations of target, target image, and retinal response are

J. M. Williams

Luminance and the Hess Effect

17

shown also in Figure 1, for a single target.

As shown in Figure

2B, if the luminance of one target be increased above that of the other, the target of higher luminance will appear to be displaced relative to the other in the direction of target motion. This

illusory lateral displacement is the Hess effect in its phenomenal form. Thus, as in simple reaction time (above), higher

target luminances again appear to be correlated with shorter visual latencies. At least three distinct approaches may be used

to obtain measures of the magnitude of the Hess effect: (a) The observer may observe the effect as just described and estimate the apparent displacement phenomenologically (Figure 2B). (b)

The targets may be presented as above but actually physically displaced by some constant distance in the direction of target motion. The observer then may adjust target relative luminance

until the Hess-effect latency difference (as inferred) compensates the physical displacement so that the targets appear aligned. The amount of additional stimulus luminance needed to make the lagging target seem to be aligned with the leading target then measures the effect as a function of the given physical displacement. (c) The targets may be presented with a constant lum-

inance difference and the observer asked to adjust their relative physical displacement until they appear to be aligned. The amount

of physical displacement needed to compensate the given luminance difference then measures the effect as a function of the difference. This last approach is shown in Figure 2C.

J. M. Williams

Luminance and the Hess Effect

18

The apparatus of the present study allows the observer to use the third method (c) described in the preceding paragraph to null a given luminance-difference by adjusting target relative displacement in the direction of target motion. The targets are In the

kept at a constant speed while in view of the observer.

present approach, an average target displacement s obtained under each given set of luminance conditions is used to compute a sample value of the corresponding inferred latency-difference t for target speed v using the formula

(target relative displacement at apparent alignment) (target speed)

s/v

(1)

For the Pulfrich effect (e.g., Lit, 1949), for any given, fixed target speed, this formula yields a latency difference t

which is found to increase (a) with decreases in prevailing level of retinal illumination (for fixed difference in target log troland value) or (b) with increases in difference of target log troland value (for fixed illuminance of the target image of higher illuminance). A latency explanation of the Hess effect, as pointed out by Roufs (l974, eq. 10) in a theoretical discussion, implies that there exists a continuous function f such that, for an obtained Hess-effect target relative displacement s at prevailing illuminance level L and target luminance difference L, s/v = t = f(L,L). (2)

J. M. Williams

Luminance and the Hess Effect

19

The Hess and Pulfrich effects are equivalent formally in that formula (2) above applies to either; the two effects differ in that the Hess requires only monocular stimulation, while the Pulfrich requires binocular stimulation. This difference allows an

analysis of central visual latencies based on comparison of obtained latency functions f under identical stimulus conditions for the two effects. 4. VALIDITY OF THE GEOMETRIC t FORMULA. The validity of formula (l) above, which is to say the constancy of proportionality of numerator s to denominator v for constant illumination, has been tested by Lit (l96Oa, b) for the Pulfrich effect. Lit's results at low to moderate photopic s/v

levels of retinal illuminance show a consistent change in

amounting to an almost linear decrease of about 0.5 ms/degree/s for speeds between 4.5 degree/s and about ll degree/s. Further

decreases at higher speeds occurred (especially at the higher target troland values) but were much smaller per unit change in target speed. At target speeds below about 4 degree/s, s/v

was found to increase rapidly as an inverse function of target speed. Cursory examination of data graphed in Lit (l96Oa, b)

suggests an inverse exponential or hyperbolic dependence of s/v on v. Closer inspection of Figure 3 of Lit (l96Ob) suggests that each of the four graphs given of s/v versus target speed (relative target illumination as parameter) has the same shape. This can be confirmed by reading t = s/v off each graph

J. M. Williams

Luminance and the Hess Effect

20

at values of, say, v = 2.5, 5, 10, and 20 degree/s and regressing t on log E (level of more intensely illuminated target); this was done by the present author, and the resulting coefficients of determination (about .99), in spite of the small sample, seem to confirm the apparent linear relation between s/v and log E at each target speed. The y-intercepts (14.4 0.4) of the four

regressions are about constant; the regression slopes m of about 8.9 (1.2) themselves may be regressed upon v to yield a cursory m' = -.14v + 10.3, coefficient of determination about .8. Thus, the approximate linear relation between t and log E seems to vary rather mildly with v, for the data used in these estimates. Lit's (l96Ob, Table III) data also seem to show a small tendency of s/v to increase with increasing target thickness, especially for targets at lower speeds and for smaller differences in log retinal illuminance (these smaller differences happen to correspond to lower troland values for the more intensely illuminated targets). This possible increase of t = s/v with

target thickness and the decrease with target speed just discussed together suggest that prolongation of stimulation is a common factor in both effects. One explanation of the seemingly anomalous results would be in terms of the traditional concept of a critical duration of summation of effect which depends inversely upon image illumination in small retinal regions stimulated by the moving target images. This dependence has been discussed by Keller (1941),

J. M. Williams

Luminance and the Hess Effect

21

Matin (1968), Roufs (1972), and Montellese, Sharpe, and Brown (1979). Cone's (1964b) approach is similar. The additional

assumption that t varies inversely with total absorbed luminous energy during exposure duration t0 leads to the following qualitative expression (target troland value E held constant):

dE

[ ]

dE

(3)

for = k (dE)-1 and

dE

E t0 dt,

with K and k proportionality constants.

Expression (3) is referred to small retinal regions and time intervals much greater than t0.

Now, for Lit's (l96Oa, b) black-on-white targets, it can be seen from equation (3) that increasing v decreases t0, thus. increasing dE linearly but decreasing hyperbolically, consistent with the data. For white-on-black targets (such as in the

present study), increasing v decreases t0 again but this time decreasing dE and increasing . when v is decreased. Similar situations obtain

J. M. Williams

Luminance and the Hess Effect

22

Another, perhaps equivalent, explanation of the departures of s/v from constancy in Lit's (l96Oa, b) data might be that the more slowly moving targets increase perceptual contributions of the more slowly responding retinal elements normally mediating visual latency only at mesopic or scotopic levels. For example, persistence of some rod response might affect localization at the lower target speeds (see Breitmeyer, 1980). Still a third explanation might be that there exist various retinal elements which are stimulated by the moving target-images but which do not respond independently of spatially-adjoinirg retinal regions. These elements might mediate a fast-spreading

local response of the retina which quickens the time-course of local retinal response. Thus, should such elements exist, the

moving target images might, through them, reduce latencies in any adjoining regions of the retina which should happen to be stimulated by light soon enough. Thus, rapidly moving target images

always would be arriving in regions with quickened response and so would have smaller latencies than more slowly moving images. This explanation does not account for any effect of target thickness on obtained values of s/v; however, support for existence of local changes in retinal response can be found in the ordinary center-surround organization of ganglion-cell receptive fields, in the Westheimer effect (Westheimer, 1965; 1967), and in distal retinal anatomy (see Kaneko, 1979; also Rodieck & Stone, 1965). A fourth explanation of the above changes in s/v with target speed might be found in terms of a hypothetical quickening

J. M. Williams

Luminance and the Hess Effect

23

of the activity of the entire retina because of rapid target image motion in some relatively small retinal region. Thus, the obser-

ved decreases in s/v for the Pulfrich effect at the higher target speeds would result from the target speed as such (e.g., McIlwain, 1964; Barlow, Derrington, Harris, & Lennie, 1977; Wist, Brandt, Diener, & Dichgans, 1977). Speculations aside, the validity of formula (1) above for the Pulfrich-effect data has not been established by Lit's (l96Oa, b) studies except only approximately for target speeds above about 5 degrees/s. The present study of the Hess effect

involved targets 1/3 degree thick and moving at a single speed of 4 degree/s (thus, incidentally, stimulating individual retinal points for durations of some 87 ms). mechanical convenience. This speed was chosen for

However, the Lit (l96Oa, b) data allow

a correction to be made in the t's obtained in the present study if generalization of the present data to some other target speed were desired. The various linearities discussed above The correction based on the

make ad hoc correction convenient.

Pulfrich data would be a few ms for generalization to any higher target speed and 15 ms or more for generalization to some lower speed, depending on retinal-illuminance conditions. An hyper-

bola fits the Lit (l96Ob, Figure 3) data quite well (see Lit, 1949). It should be mentioned here that preliminary data gathered in the course of the initial testing of the Hess apparatus (under stimulus conditions the same as those of the present study itself) suggest that, as might be expected from the Pulfrich data just

J. M. Williams

Luminance and the Hess Effect

24

cited, t calculated by formula (1) above does increase greatly for the Hess effect at target speeds lower than those used in the present study. But those preliminary data also suggest that t Therefore,

changes negligibly for target speeds above 4 degree/s.

in the Hess effect, formula (l) above will be assumed to have about the same validity as for the Pulfrich effect. Empirical

correction of formula (1) therefore probably would require addition to s/v of a factor likely to be linearly dependent upon the difference in log troland value of the targets. probably would be an hyperbolic function of v. It must be emphasized that all the above questions of the "validity" of formula (1) refer to the use of this formula directly to obtain a measure of visual perceptual latency difference under some given set of experimental circumstances. The theory This factor

proposed below is not based upon this formula but rather yields theoretical values of a hypothetical visual latency difference depending upon various stimulus conditions including target retinal illuminance. The theoretical latency-difference is applied to

the geometric conditions of the Hess experiment, and, granting the validity of formula (l), thus generates predictions as to what the available data should be in terms of the empirical values of measured during the experiment. The accuracy of the proposed s/v

theory is tested validly by the present study only to the extent that formula (1) validly measures the visual latency being predicted by the theory.

J. M. Williams

Luminance and the Hess Effect

25

The importance of the data actually obtained by Lit (l96Oa,b) is that they suggest that the Hess and the Pulfrich effects do not depend upon a visual latency difference itself governed solely by prevailing conditions of illumination. Wist, Brandt, Diener,

and Dichgans (1977) have pointed this out in a study of the Pulfrich effect using gratings which made it possible to filter the spatial-frequency spectral content of their targets. values of t = s/v thereby were obtained for targets containing just those higher spatial frequencies which are reduced when width of the rod (bar) shaped targets used in Lit (l96Oa, b) and in the present study is increased. Thus, another facet of the Larger

Pulfrich effect (and presumably the Hess effect) was revealed which, like the results of Lit (l96Oa, b) suggest that the geometric latency formula s/v measures something more than a simple luminance-dependent latency difference.

J. M. Williams

Luminance and the Hess Effect

26

VISUAL LATENCIES 1. THE VISUAL SYSTEM. The visual system in mammals, including humans, would seem to be entirely visual from the eye up to the thalamic and midbrain centers; it is at these centers that the first interactions occur with afferents from nonvisual modalities or with central nervous system neural circuits. In the present work, the term central

nervous system (CNS) is being used so as to exclude the neural retina; reasons for this will be discussed below. Suffice it to

mention here that there is little evidence at present for a CNS efferent pathway to the mammalian retina (e.g., Rodieck, l973, ch. 21). Such efferents might place mammals at a selective disCurrently, it would seem that

advantage (e.g., Williams, 1979).

the only likely efferents to the neural retina would be the autonomic (sympathetic) fibers controlling tonus of the retinal blood vessels; these fibers presumably follow the central retinal artery into the globe, branching and subdividing to regulate the arteriolar flow which nourishes the inner layers of the neural retina. The evident lack of retinal central efferents justifies the treat~ ing of all latencies of afferent response up to the dorsal lateral geniculate nucleus (LGNd) and the superior colliculus (SC) purely as visual latencies. In humans, the LGNd contains the single subcortical afferent synapse known to exist between retina and the primary visual receiving area of the striate cerebral cortex (area 17 of Brodmann).

J. M. Williams

Luminance and the Hess Effect

27

The situation is known to be more complex in cats (Singer, Tretter, & Cynader, 1975; Tretter, Cynader, & Singer, 1975) in which Area 18 also receives direct primary visual afferent input via the LGNd. The fast-conducting "Y-system" of retinal ganglion cells seemingly preserves its identity through the LGNd so that it is primarily retinal "Y-cells" which provide visual afferent inputs for the cat Area 18. The cat retinal "X-cells" seem to relay at the LGNd The axons of cat retinal "Y-cells"

to project to Area 17 only.

(see Tretter, Cynader, & Singer, 1975) are those that bifurcate just before reaching the LGNd to project also to the SC. The

cat "X-cell" pathway does not appear to branch before the LGN. The Area 18 projections notwithstanding, there is no particular reason to suspect that the cat visual system is atypical among those of mammals. Perhaps one important point has been made by

Singer, Cynader, and Tretter (1975) which would bear emphasis: Cats tend to have a more highly interconnected CNS than other mammals such as primates. The greater part of neurophysiological

research in vision has been done with cats, to which creatures the "X-Y" classification most immediately applies (see Rodieck, 1979). As an exhaustive categorization, the "X-Y" dichotomy probably now has become outmoded, even the erstwhile third-alternative "W-cells" having been found to be heterogeneous and too vaguely defined for neurophysiological purposes (Rodieck, 1979). However, the structur-

al basis for the spatial nonlinearity of "Y-ce11" receptive fields versus the consistent linearity of "X-cell" receptive fields (Hochstein & Shapley, 1976) has not been elucidated yet and remains a

J. M. Williams

Luminance and the Hess Effect

28

significant unknown factor in retinal anatomy.

Functional class-

ifications based on spatial linearity of receptive fields remain useful (e.g., Kratz, Webb, & Sherman, 1978). Species differences aside, purely visual latencies in principle might be measured not only at LGNd and SC, but at afferent regions of the pulvinar nucleus, the pretectal area bordering the SC, the basal ganglia, and at other regions described in references cited below. Apart from the technical difficulties

of recording from any such sites, a basic problem in the analysis of the mechanisms of visual latency resides in the physiological variability among individual members of a species, a variability which makes absolute interindividual comparisons difficult a priori-especially for individuals as variety-blessed as ordinary human beings. Even the best stereotaxic atlas can be no better than Indi-

the uniformity of the individuals of the species studied.

vidual variability, minimal at the retina, becomes overridingly important in the complexities of the CNS, which is liable to link stimulus and response by any equipotential means available (e.g., Hebb, 1938, l942). The question of whether a given anatomical element of the visual system is influenced by nonvisual afferent input is not an idle one: Psychophysical techniques usually are the only avail-

able ones for studying visual latency in normal human beings. If neurophysiological inferences as to visual function are to be drawn validly from psychophysical data, psychophysical response measures must be chosen such as will be mediated solely by the

J. M. Williams

Luminance and the Hess Effect

29

visual system as such--or at least which are known to be influenced minimally by uncontrolled CNS sources of variability. Merely

assuming that some component of a response-measure is due to "nonvisual CNS" or "motor" effects begs the question that the response change (as a function of visual stimulus-change) is purely "visual." In any given case, it must be proven physiologically that

a given response component meaningfully "is" nonvisual and may be subtracted or factored away to leave a purely visual latency, amplitude, number of cases, or whatever, suitable for psychophysical study. Thus, there are only a few purely "visual"

responses such as the electroretinogram (ERG) and the retinal reflection photodensity which are available for use with normal, intact human beings. Other measures require neurophysiological

support if inferences are to be drawn on the visual system as a distinct entity. Single-unit or gross-electrode recording, of

course, presents no insurmountable difficulty when subhuman species are being studied; this approach has a psychophysics of its own. In most research done with humans, results of controlled experimental conditions are combined with knowledge of pathological human cases and of research with animals. Experimen-

tal control of extraneous stimulation serves to standardize or regularize all CNS activities of the individual being studied, thus leaving the obtained behavioral differences validly functions of the various visual stimulus conditions. Behavioral

measures of visual mechanisms typically have been scalp-electrode recordings of the electroencephalogram (EEG) and of visual-evoked

J. M. Williams

Luminance and the Hess Effect

30

potentials (VEPs), the latter (e.g., Regan, 1975) presumably being mostly cortical in origin. The apparently cortical

source of visual-evoked brain potentials has had not a little effect in biasing studies of the brain toward the cortex to the exclusion of the subcortical centers, at least until this decade past. Another biasing factor, one might suppose, would be that

brain injuries which can be survived tend to be cortical--whence a great deal of clinical interest. A recent technical advance

holds great promise for the functional analysis of the visual CNS: Superconducting magnetic-field detectors for magnetoencephalography (Brenner, Lipton, Kaufman, & Williamson, 1978) have been shown to localize brain responses more precisely than is possible with the EEG or VEP. It can be shown by Maxwell's equations that in any small region in which energy is conserved in electromagnetic form, the same changing electric fields which are measured by the EEG or VEP also must be accompanied effectively simultaneously by changing magnetic fields of well defined amplitude yielding the same instantaneous relative intensities. This has been known for a

century, but the magnetic fields have been too weak, compared with the electric fields, to be detected with apparatus available until very recently. The relative weakness, as may be seen by exam-

ination of the forces involved, derives from the shorter distances over which a magnetic force can be exerted and still have a measurable effect on a charge-carrier such as an electron in a piece of monitoring equipment: First of all, magnetic monopoles

J. M. Williams

Luminance and the Hess Effect

31

are not known to exist, but electric monopoles are ubiquitous. For an electron, the electronic spin-magnetic dipole has a field which, with its tiny dipole moment, falls off as the cube of the distance; the same electron has an electric field deriving from its monopolar negative charge which falls off as the square of the distance. The inverse-cubic falloff of the magnetic field

allows the electron to be localized more precisely by magnetometry but to be detected with greater difficulty. Secondly, even for

electric and magnetic dipoles of equal dipole moment, for the electric field strength to be equalled by that of the magnetic (inductive) field strength (e.g., Halliday & Resnick, 1962, Table 34-1), an incredibly large current would have to flow for dipoles or charge movements with dimensions on the order of those of a neuronal cell membrane (about lO-8 m). Thus, in the past, the

magnetic dipoles present in biomembranes could not be localized either while static or while being displaced discretely during typical changes of biological membrane structure during neuronal response. Thirdly, the electromagnetic field changes caused by electric charge displacement involve only that charge itself, which is to say, only a limited, known array of charge-carriers. So, the

magnetic field strength just mentioned in the second point above only could be achieved by incredibly high velocities of charge-displacement to attain the necessary huge current. Thus, the obvious

requirement that neural activity correlate with ionic displacements across biomembranes precludes accompaniment of such activity by magnetic fields of strength anywhere near the strength of the

J. M. Williams

Luminance and the Hess Effect

32

electric field of the ions.

Nevertheless, the magnetic fields These fields sim-

existed and were known to exist for a century.

ply could not be detected (not to mention be identified) in the presence of the random magnetic-dipole fields caused by thermal motion in the crystal lattices of the metal of the probes available to be applied at the scalp. Supercooled detectors have a lack of

thermal noise (as well as a zero loss of stored current), which explains their effectiveness. In any event, the more exclusively cortical signal of the magnetoencephalogram should allow signal subtraction from EEG or VEP to obtain the deeper, subcortical components, if any, of these latter electrical measures. 2. MECHANISMS OF VISUAL LATENCY. The components of the visual afferent system may be considered the retina, the retinal direct projections to the LGN and SC, the striate and parastriate cerebral cortices, and perhaps other retinal projections to the thalamus and mesencephalon. The direct pathway from retina to striate cortex (via optic nerve, chiasm, optic tract, LGNd, and the optic radiations of the corona radiata) is monosynaptic, with lateral (retinotopic), recurrent, and interocular inhibition at the LGNd. Some retinal ganglion

cell axons bifurcate before reaching the LGNd and follow the superior brachium below the pulvinar to the SC and to pretectal nuclei. Other retinal ganglion cells project only to the mes-

encephalon, while others project only to the striate cortex. One possible reflex path is shown in Figure 3. Descriptions of

J. M. Williams

Luminance and the Hess Effect

33

the afferent pathways may be found in Duke-Elder and Wybar (1961) Lockhart, Hamilton, and Fyfe (l969, pp. 345-362), Brindley (1970a, b), Warwick and Williams (1973, ch. 7), Poggio (1979), Poppel, Held and Dowling (1977), Blake (1979), Grusser (1979), Rodieck (1979). and van Essen (1979). Recent reviews of the CNS interactions

with retinal afferents, especially at the LGNd, may be found in Singer (1977) and in Burke and Cole (1978). A concise summary

of much of the known cerebral cortical organization has been published by Hubel and Wiesel (1979). The SC has not been so well

studied, but a recent review has been done by Gordon (1975), and other pertinent data on the SC may be found in Singer, Tretter, and Cynader (1975), Tretter, Cynader, and Singer (1975), Schiller (1978), and Rodieck (1979). The structure of some mesencephalic

visual connections are discussed by Hobson and Scheibel (1980), and a review of SC visual-motor integration appears in Wurtz and Albano (1980). To the present author's knowledge, no retinotopic

functional regions have been found in the human brain except at the LGN, the SC, and various cortical areas, the last of which all seem to map from fibers originating in Area 17, the LGNd, or the SC. In the retinocortical system, components of response latency would be localized mainly to the retina, the LGNd, or the occipital cortical visual areas (Brodmann's Areas 17, 18, and 19). Cal-

losal interhemispherical connections have been found for Areas 17, 18, and 19, but perhaps only for cortical regions retinotopically processing information from regions near the retinal vertical

J. M. Williams

Luminance and the Hess Effect

34

meridian through the fovea (Berlucchi, 1972).

The reviews

mentioned in the preceding paragraph document considerable modulation of LGNd activity by the cerebral cortex as well as by subcortical (including brainstem-reticular) brain regions. Much of these data are obtained from animal studies but may be expected to hold for humans to greater or lesser extent, depending on such species-specific characteristics as foveation, color vision, and binocular overlap of visual field. Modulatory activity

at the LGNd might be involved in learning (e.g., Bartlett & John, 1973; Kruger, 1979); but considering the one-to-one retinal ganglion cell-LGNd principal cell connectivity (e.g., Poggio, 1974), the single synapse in any single afferent path through the LGNd would seem to provide little flexibility in its contribution to response latency at the visual cortex. The fibers of the optic nerve are believed all to be myelinated (Ogden & Miller, 1966; Magoon & Robb, l980). Based

on Rodieck (1979, Table I), on Ogden and Miller (l966), and on scale drawings and photographs of the brain (e.g., Duke-Elder & Wybar, 1961), action-potential latency between a retinal ganglion cell discharge and LGNd response would amount to 5 or 10 me in humans (see also Sumitomo, Ide, & Iwama, 1969; Tretter, Cynader, & Singer, 1975). Depending on ganglion-cell location in the

retina and on fiber diameter, as well as on spacing of axonal nodes of Ranvier, conduction times would range from a minimum of about 2 ms to perhaps 20 ms, with an average of perhaps 4 ms. similar latency due to conduction time would be expected between A

J. M. Williams

Luminance and the Hess Effect

35

LGNd and Area 17.

The values for the cat given in Singer, Tret-

ter, and Cynader (1975) would total to about 2 to 10 ms between retina and Area 17. Thus, if action-potential delay across a given LGNd afferent-pathway synapse could be varied by a factor of two by CNS modulatory activity, the difference could not amount to more than a few ms, a reasonable fraction of the total latency to Area 17 due to conduction time, but a small fraction of most latencies tabulated in Appendix I below. The circuitous nature of visual

cortical pathways beyond Area 17 (e.g., Kaas, 1978) would allow indefinite delay of behavioral response and cannot be analyzed further by the present approach. It is interesting that the

relatively complex retina of the frog can accomplish delays of 20 seconds (Pickering & Varju, 1969) before generating an optic nerve response to visual stimulation. Because the data be1ow gathered for the present study will not directly represent activity of any single component of the CNS, the latencies derived will be examined only in terms of two CNS subdivisions: A retinal component localized by assumption to

regions distal to the first myelinated portions of the optic nerve; and a CNS component proper localized proximal to the optic disc. As mentioned above, the classical reaction time (RT)

literature shows that further deductions would be tenable; however, any deeper analysis would be less directly supportable by data and less clearly amenable to quantification than the simple retina-CNS dichotomy chosen.

J. M. Williams

Luminance and the Hess Effect

36

As an example of the use of the retina-CNS dichotomy, consider the measurement of minimum possible reaction time for a behavioral response. The reasoning then is simply that the min-

imum RT must depend on the shortest and "simplest" CNS pathway{s) available, and that the effects of persistence of the response in the same or parallel pathways must be negligible. Thus, by res-

tricting the analysis to (minimum) latency alone, an implicit assumption of simplicity allows for relatively straightforward, if not necessarily accurate, interpretations of the data. Even

the extraordinarily simple anatomy of the vestibuloocular reflex arc (Szentagothai, 1950) allows considerable central modulation of activity; functional pathways (e.g., Hobson & Scheibel, 1980, pp. 24 ff.) thus include much more than a reflex itself unless the response being studied be defined in terms of minimum possible latency. The example is illustrated in Figure 3. The eye-blink

reflex, a response to visual stimulation with sudden onset, can be conditioned and has a latency of 50 to 110 ms (Kimble, 1961, pp. 55-59; see also Bixler, Bartlett, & Lansing, 1967), possibly the shortest latency of any easily-observable behavioral response It is almost irresistable to partition the obtained latency for such a response into additive sensory and motor components. For example, Brenner, Lipton, Kaufman, and Williamson (1978) performed just such an analysis of somatosensory cortical evoked magnetic fields, obtaining a (somato)sensory component of about 70 ms and a motor component of about 100 ms. Clearly, along

J. M. Williams

Luminance and the Hess Effect

37

this line of reasoning, the eye-blink reflex of Figure 3 must have a negligible CNS time component: In fact, the eye-blink

latency is of the same order as latencies obtained for peaks of the ERG a- or b-wave (see Appendix I below). Data from a

study by Bixler, Bartlett, and Lansing (1967) allow estimation of eyelid-muscle contraction time at 30 ms or so, leaving, for a total RT of about 60 ms, a total central component of some 30 ms. Considering the intensity levels at which Bixler, Bartlett, and Lansing (1967) were stimulating, it is possible to combine their findings with those of Alpern (1968). In the latter study, com-

putations based on Pulfrich-effect data from Lit (1949) and confirmed under slightly different conditions by Prestrude (1971, Figure 7) allowed estimation of an absolute visual latency at something near 15 ms, all but a few ms, presumably, localized at the retina. Although the experimental situations involved

were very different (as were the observers), the studies just cited would allow, as an example, the following latency analysis of the eyeblink reflex: 15 ms (retinal) + 15 ms (CNS) + 30 ms (peripheral motor) = 60 ms RT. latencies were fairly high. The intensities for these In any case, this sort of approach,

which ignores any dependence of CNS or motor delay upon stimulus intensity, derives directly from the separation of reaction time into sensory and motor components, historically at the very roots of modern psychology. Historical details may be found in

Boring (1950, ch. 8) or Woodworth and Schlosberg (1954, ch. 2).

J. M. Williams

Luminance and the Hess Effect

38

But reaction time involves absolute latency and the Hess effect seems to involve a relative latency. It is not obvious

how Hess-effect data alone might be used to localize a latency difference even for a simple choice between retina and other-CNS. Other information on the time-course of visual response to light must be considered for the relative contribution of the retinal mechanisms to be estimated. The classical research on the physiological substrate of visual latency was done by Adrian and Matthews (l927a, b), who showed for the conger eel that the gross BRG and subsequent gross optic nerve action potential did not have the same neural origin. These authors found that (within limits) if the duration, intensity, or area of retinal stimulation should be increased, the latency to any criterion level of optic-nerve gross action-potential would be decreased. These findings are consistent with In addition,

the human data summarized in the Introduction above.

Adrian and Matthews (l927a) also found that there seemed to be a relatively fixed time delay of about .1 second between the first ERG a-wave and the initiation of the gross optic nerve action potential; thus, the obtained experimental variations in optic-nerve latency could be attributed to retinal processes more closely related to the photochemical transduction of light to retinal neural response than to the subsequent transduction of retinal neural response to the optic-nerve potential. This conclusion

was investigated further and supported in another classic study by Bernhard (l94O) on frog and human evdked responses.

J. M. Williams

Luminance and the Hess Effect

39

Bernhard (1940) based an electrophysiological study of the visual system on the human occipital "alpha-blocking" response to a light suddenly presented in an electrically-shielded, darkened room. The alpha-blocking response apparently occurs simultan-

eously over the entire scalp, although originating, evidently, in the occipital cortex. This simultaneity has been interpreted in

terms of the field of a 10-Hz electric dipole located at the calcarine cortex of the cerebrum and radiating alpha over the entire skull (Vaughan, 1974). Regardless of the geometry of its source,

there is little doubt that the alpha-blocking response originates mainly in the visual cortex, an inference accepted by Bernhard. In one study of five human observers, Bernhard (1940) obtained data which showed an exponential-like decline of blocking time from 600 to about 180 ms for a five log-unit increase of stimulus luminance ranging from lO-4 to 10 millilamberts. An

only slightly shallower curve of decline was obtained for optic-nerve latencies of dark-adapted frogs for the same range of luminances, the frog latencies ranging from about 400 down to 80 ms. Bernhard

(1940) evidently did not use artificial pupils and so, unfortunately, could not equate retinal illuminances corresponding to the stimulus luminances used. Bernhard (1940) proceeded to another study in which the human ERG was recorded while alpha-blocking latency again was measured as before. At three low-photopic levels of luminance,

occipital alpha-blocking latency was found to range from about 180

J. M. Williams

Luminance and the Hess Effect

40

to 240 ms, decreasing with increasing luminance as before; approximately the same change in ERG b-wave latency (viz., 60 ms) was obtained under the same conditions. Assuming the b-wave to rep-

resent bipolar-cell activity in the retina, Bernhard (1940, Figure 26) concluded that the main dependence of visual latency upon stimulus intensity is due to retinal processes localized distal to the source of the ERG b-wave; such processes would include activity of the photoreceptors, the horizontal cells, Muller cells, and their interconnections. It should be mentioned that recent evidence

has been advanced that the b-wave of the ERG is not neural in origin, leaving the a-wave, possibly, as the only neural component of the ERG with an important intensity-dependent latency (Baron, Boynton, & Hammon, 1979. Green, 1980). cf. Szamier, Ripps, & Chappell, 1980; Vogel & Nevertheless, as the a-wave seems to arise in the

photoreceptors (see below) and to precede the b-wave by a constant interval at least in the toad eyecup (Mansfield & Daugman, 1978), Bernhard's (1940) inferences about distal retina would seem to hold regardless of the anatomical source of the b-wave. Approaching the problem of latency-analysis from another aspect, Bernhard (1940) made a series of measurements of alpha blocking latency and motor reaction time. The obtained results,

with reaction timed both for overt response and for bicep electromyogram, showed that the "motor component" of reaction time remained at about 80 ms (for overt response and at about 42 ms for elec-

J. M. Williams

Luminance and the Hess Effect

41

tromyogram) regardless of stimulus intensity under the experimental conditions (Bernhard, 1940. pp. 77-79). Further experiments

by Bernhard (1940) on intensity-dependent visual masking will not be discussed in the present paper. Bernhard (1940) thus showed, probably in the first study of its kind in humans, that the moiety of the intensity-dependence of visual latency could be attributed to retinal transduction processes. Various aspects of Adrian and Matthews's (l927a) and Bernhard's (1940) findings have been corroborated by Monnier (1949; 1952), Lennox (1956), Cobb and Dawson (1960), Prestrude (l971), Heron, Regan, and Milner (1970), and Mansfield and Daugman (1978). As

discussed by Brindley (1970, pp. 63-64), many of the older studies (including Adrian and Matthews's) found a dependence of ERG latency upon retinal stimulus area. More recent designs (e.g., Brindley

& Westheimer, 1965) have eliminated effects of the increase in intraocular scattered light which accompanies an increase in stimulus area. When scattered light is rendered ineffective, ERG Waveforms of

latencies are not found to depend upon stimulus area:

local and global ERG then are found to have about the same shape and temporal characteristics (Brindley, 1970, p. 64), although, of course, the larger areas still yield greater ERG amplitudes in proportion to the area stimulated. Because the ERG seems to show no

change in latency with change in stimulus area, any area-dependence of more central visual latencies would be expected to originate at

J. M. Williams

Luminance and the Hess Effect

42

site(s) at which some CNS correlate of the greater ERG amplitude would be effective by convergence of retinotopic inputs. At such

a region of convergence, the greater area would have its earlier effect (hypothetically) (a) by spatial summation for earlier achievement of some threshold response amplitude and (b) by the statistically more likely earlier occurrence of the shortest-latency retinal single-unit response to the larger-area stimulus. In view

of the ERG data, the retinal inner nuclear layer would be the most distal possible locus of such convergence. If it be required to

select just one such convergence site, the striate cortex would suffice; but the greater convergence of the less precisely retinotopic projections to the optic midbrain would seem to offer a more likely locus for CNS determination of dependence of visual latency upon area of the stimulus retinal image. Of the more recent studies just cited, Monnier's (1952) data suggest a minimum retinocortical conduction time of about 5 ms (these data refer to high intensity flashes yielding ERG b-wave latencies of 25 to 30 ms and corresponding earliest VEP cortical-positive potentials occurring 33 to 42 ms after stimulus onset). The est-

imate above based on gross-anatomical estimated axonal conduction times yielded a comparable 4 ms, as the reader might recall. The

shortest-latency ERG components, the early and late receptor potentials, become evident in the ERG only at the highest stimulus intensities and occur 60 microseconds (early) and perhaps l millisecond (late) after stimulus onset (Brown & Murakami, 1964a, b; Cone, 1964c, 1965; Goldstein & Berson, 1970). The early receptor

J. M. Williams

Luminance and the Hess Effect

43

potential survives death and tissue fixation (Brown & Murakami, 1964b; Arden, Bridges, Ikeda, & Siegel, 1968) and, whether it be a "biological" response at all, must arise at or very near the site of the visual-pigment chromophore. Various photovoltaic effects

in inorganic semiconductors (Pankove, 1975, ch. 14) differ from the early receptor potential only in that the latter seems to have the action spectrum of rhodopsin and seems to be proportional in amplitude to the fraction of visual pigment present when the eliciting flash is delivered. With respect to rod-cone duplicity theory, the identity of the photoreceptor class(es) originating the ERG does not seem to be an important problem. Work by Arden and Weale (1954), Gold-

stein and Berson (1970), Conner and MacLeod (1977), and Gordon, Shapley, and Kaplan (1978) all suggest that rods and cones both would have been contributing to the measurements of Adrian and Matthews (1927a) and Bernhard (l940). Gouras (1966) and Gouras

and Link (1966) provide data on monkey ganglion cells which suggest that either rods or cones will control optic-nerve action potentials, depending upon whichever of these photoreceptors has the shorter latency at some distal neural stage of processing ("between b-wave origin and ganglion cell"; Gouras, 1966, paraphrased). A similar convergence of rods and cones in terms of adaptation level was shown by Rodieck and Rushton (1976) for the cat. The

specific functional dependence of latency upon stimulus intensity, however, may well differ for rods and cones at the photoreceptor level, a question which will not be addressed further at this point.

J. M. Williams

Luminance and the Hess Effect

44

As discussed above, the importance of the more distal retinal elements for intensity-dependence of visual latency was established reasonably well by Adrian and Matthews (l927a, b) and by Bernhard (l940). An obvious choice of mechanism would be the

phototransduction process, which occurs in the photoreceptor outer segments. Data on the vertebrate phototransduction process may

be found in Barlow and Fatt (1977), Hubbell and Bownds (1979), and in Hagins (1979). Vertebrate phototransduction is not entirely

understood at present but seems to depend on processes different from phototransduction in invertebrates (e.g., Rodieck, 1973, XI.A.3). The difference presumed most important for the present

purposes is that the invertebrate photoreceptor signal seems to involve a change in membrane potential proportional to the logarithm of stimulus intensity, whereas the vertebrate plasma membrane potential seems to change linearly with stimulus intensity (MacLeod, 1978). Information on invertebrate photoreceptors may

be found in Wolken (1971). deVoe (1967), and Ratliff (1974); Levinson (1972) has published a theory based mainly upon invertebrate generator potentials, and Hartline (1934) showed that Bloch's law held for single Limulus ommatidia. Even in the ver-

tebrate photoreceptors, though, a logarithmic dependence of normalized response amplitude upon stimulus intensity can be shown: This has been done by Baylor, Lamb, and Yau (1979), for single rod outer segments the response of which is neither zero nor sat~ urated.

J. M. Williams

Luminance and the Hess Effect

45

From Barlow and Fatt (1977), Hubbell and Bownds (1979), and Hagins (1979), the phototransduction process in vertebrates would seem to begin with photoisomerization of 11-cis retinal to all-trans retinal. The action spectrum of this change presum-

ably depends upon the protein to which the retinal is linked. The early receptor potential, as mentioned above, is simultaneous with this isomerization (or follows it with a latency of some tens of microseconds) and may or may not be related to the isomerization causally. In any case, the photoisomerization initiates an event

which is amplified rapidly in power and evident1y culminates in a blockage of Na+-channe1s by Ca2+ (cf. Liebman & Pugh, 1979) in the photoreceptor plasma membrane, a blockage which stops the "dark current" of sodium ions, hyperpolarizes the outer segment, and probably generates the ERG a-wave as well as a synaptic response relaying the effect of the photoisomerization to the horizontal and bipolar cell dendrites. In Necturus, at least, the photo-

receptor inner segments generate a "slow potential" with the same latency as the ERG a-wave (Bortoff, l964), and this "slow potential" thus would reflect the latency of the intracellular amplification process itself--or perhaps the latency of a horizontal cell feedback (v. Mansfield & Daugman, 1978) to the inner segment in response to a receptor potential received and amplified by the horizontal cell shortly before. The receptor potential

initiates events in more central pathways perhaps as analyzed and reviewed for the cat by Grusser (1979). Ke1ly and Wilson (1978)

suggest that flicker response might be based on two serial stages

J. M. Williams

Luminance and the Hess Effect

46

of diffusion in the distal retina. Until very recently, for lack of data the psychophysical literature has avoided specific postulation of cellular mechanisms which might underly the intensity-dependence of visual latency. Descriptive theories were suggested as early as Ives (1922) and Hecht (1928); and more recently by Veringa (1961), Vaughan, Costa, and Gilden (l966), Alpern (1968), and Sperling and Sondhi (1968). Similar efforts have been made by Kelly (1969b), Prestrude (1971), Mansfield (1973), and Kelly and Wilson (1978). Many such theories

were reviewed by van de Grind, Grusser, and Lunkenheimer (1973). Mansfield and Daugman's (1978) study included a least-squares comparison of various theories for fit of data gathered for the toad eyecup. Latency formulas derived from various of these and

other theories will be given below. It might be added here that in the present work, the word transduction is intended merely to refer to a process amenable to explanation in biophysical terms (e.g., Ashmore & Falk, 1979): Usually, there is nothing intended to imply that the "transduced" energy is conserved in the strict1y physical sense. In fact, the

relatively constant structure of most neural elements would suggest constant or negative entropy, and thus a constant expenditure of energy even at ultrastructural levels. With this caution, the

studies summarized in the present section above may be interpreted to support the conclusion that the phototransduction process itself is likely to be the major mechanism determining the dependence of visual latency upon stimulus intensity.

J. M. Williams

Luminance and the Hess Effect

47

3. INTENSITY-DEPENDENT VISUAL LATENCY:

LITERATURE.

The suggestions of Adrian and Matthews (l927a, b), Bernhard (1940), Monnier (1949, 1952), and others regarding the overriding importance of the retina in controlling visual-system latency were examined further in a search of the literature. results of this search are entered in Appendix I below. The reader will find that the studies in Appendix I are arranged in increasing order of range of stimulus intensity for which visual latency values were published in the respective studies. Many of the entries in Appendix IB were values measured The

by ruler directly from published graphs of ERG waveforms, reaction times, etc., and so incorporate a certain error due to the present author. The somewhat inconsistent but reasonably clear tendency

of latency-range to vary directly with intensity-range across studies can be seen in Appendix IB. This covariation is a

fundamental fact necessary for full understanding of the present Dissertation. Exceptions, even within studies confined to special

stimulus conditions, are rare (see Rogers & Anstis, 1972). For the few studies for which data were available, ratios of latency-range for ERG response versus for other-CNS response were computed and squared, the result being entered in the last column of Appendix IB: other-CNS. This variance ratio exceeds l/4 when the estimated

retinal dependence of latency upon intensity exceeds that of the Such statistics, as well as the other data in the ApReaction time, specifpendix, tend to show (across studies) that a large fraction of total visual latency can be varied at the retina. ically, as a function of intensity is reviewed in Menendez (1979).

J. M. Williams

Luminance and the Hess Effect

48

4. SELECTED THEORIES OF INTENSITY-DEPENDENT LATENCY. In addition to those based on the geometric t of the Introduction (pp. 8 ff. above), several descriptive theories have been mentioned in the section on Mechanisms above. Only theories

readily adapted to describe the Hess effect will be discussed in the present section. Theories applicable to the Hess and related spatiotemporal illusions all must generate absolute latency functions which approach an approximately-horizontal asymptote as stimulus intensity is increased. The generalized graph of such a function can

be found in Lit (1949, Figure 8); an analytic expression for such a graph typically would take the form of a negative exponential t = a exp(-b I) (4)

or of an hyperbola of some sort; for example, t = b a + , Ic or, (5a)

1/t

= b

a +

, Ic

(5b)

the latter of which is the same as t = a' Ic + b' . (5c)

In the above expressions, t represents latency and I stimulus intensity, with a, b, and c parameters.

J. M. Williams

Luminance and the Hess Effect

49

The exponential is a natural function in the context of differential equations describing the visual process (e.g., Fuortes & Hodgkin, 1964; Alpern, 1968; Levinson, 1972); the hyperbola is natural in the context of Laplace transforms of such differential equations (e.g., Kumar, 1975) or more directly in terms of chemical or ionic equilibria of photopigments or of closely-associated membranes (e.g., Baylor & Hodgkin, 1974; Naka & Rushton, 1966; Briggs & Haldane, 1925). The hyperbolic tangent curve

used by Naka and Rushton (1966), of course, itself is a function of exponentials: tanh(x) = = By definition, for any variable x, sinh(x) cosh(x) = exp(x) exp(x) + exp(-x) exp(-x) (6) 1 1 + exp(-2x) 1 1 + exp(-2x)

This kind of function, as shown by Mansfield and Daugman (1978), may be used to describe inverse relative latencies (1/t) asymptoting to a constant for large values of stimulus intensity x. In a very general sense, then, the differential equa-

tions cannot be avoided unless a pragmatic or a purely empirical approach be taken. An empirical route was adopted by Keller

(1941, Figure 2), who found an hyperbolic expression for the critical duration of summation of Bloch's law (equivalent to a latency of termination of complete summation: See Matin, 1968; cf. Kong

& Wasserman, 1978) in the context of Hecht's photochemical theory of intensity discrimination.

J. M. Williams

Luminance and the Hess Effect

50

Much of the theoretical literature on visual latency has been reviewed in Harmon and Lewis (1966, esp. pp. 552-553), Prestrude (1971), Rogers and Anstis (1972), Teichner and Krebs (1972), Mansfield (1973), van de Grind, Grusser, and Lunkenheimer (1973), Roufs (1974), and Mansfield and Daugman (1978). Of the many

theories cited in these works, several are compared below with the present proposed theory. Normal equations for least-squares

estimation of the respective theoretical parameters may be found in Appendix II of the present work, and the theoretical expressions themselves are as follows: a. Exner's: t = [a (log I) + K]-1 , (7)

the parameter a being described as maximal for rods and decreasing for cone vision (see Pieron, 1922; Kleitmann & Pieron, 1925). b. Charpentier's: t = a I-(1/n) + K , (8)

in which the parameter K is described as differing for rods and cones, with n equa1 to about 4 (see Kleitmann & Pieron, 1925; Pieron, 1922). c. Pieron's: t - t = k Ib , (9a)

J. M. Williams

Luminance and the Hess Effect

51

which is equivalent formally to Charpentier's and may be transformed to t t = I

I +

(9b)

for = -b, and = (k/t)1/ . In the original Pieron (1923a) form of (9a) above, t = asymptotic value of latency k = about 60, for flashes enduring beyond 30 ms and for t in ms and I in candles/m2). b = about -.33 for extended sources. (see also

Mansfield, 1973; Rogers and Anstis, 1972, Table I). d. Ives's (1917): t = t l - t2 = [a log I1 + b]-1 - [a log I2 + b]-1 , (10)

each term of which on the right is the same as Exner's expression for absolute latency above. e. Hecht's (1928): t - t = [k log(I/I0)]-1 , (ll)

which also has been used by Bartlett and MacLeod (1954).

J. M. Williams

Luminance and the Hess Effect

52

f. Lythgoe's (1938): t + c = 0.02 log(Eu/El) , (12)

in which the small constant c represents the (residual) difference in latency between the eyes with no filters in place. g. Roufs's (1963): t = t 1 - t0 = -T ln(E1/E0) , (13)

which is formally identical to Lythgoe's above and for which T = about 10 ms. h. Alpern's (1968): t0 - t = K I-1/(n
+ 1)

(14)

again formally equivalent to Charpentier's or Pieron's above and in which Lit's (1949) data for the Pulfrich effect have yielded, for I representing troland value , t0 K and n = = = 71.9 ms 78 ms about 4.

i. Enright's (1970): t = log(Eu/El) [Y + (S/2) log(EuEl)] , (15)

in which Eu = troland value of target of higher luminance, El = troland value of target of lower

J. M. Williams

Luminance and the Hess Effect

53

luminance, and, again using Lit's (1949) data on the Pulfrich effect, Y S = = 25.7 ms -5.75 ms. and

In the literature, under some circumstances an average or "ambient" illumination level has been invoked in order to justify certain theoretical manipulations (e.g., Ives & Kingsbury, 1916; Enright, 1970). A general approach to the quantification of

average effects of stimuli differing by finite amounts among themselves (or against their backgrounds: See Rogers, Steinbach,

and Ono, 1974) has been formulated by Helson (l948), and, from a different tack, by Land and associates (e.g., Wyszecki & Stiles, 1967, pp. 444-446; McCann, Savoy, & Hall, 1978). The theories

listed above absorb adaptation effects in their free parameters; it is worth remembering, though, that adaptation (as a response) has not an inconsiderable effect on visual latencies, as witness the results of Rogers and Anstis (1972) for the Pulfrich effect.

J. M. Williams

Luminance and the Hess Effect

54

THE HESS EFFECT 1. LITERATURE. In view of Roufs's (1974) brief but thorough review of the major latency effects in vision, the present discussion will be limited to the Hess effect per se. The Hess and other laten-

cy-dependent phenomena (at least by assumption) have been treated in Poffenberger (1912), Bills (1920), Vogelsang (1928), Bernhard (1940), and Teichner and Krebs (1972). As an historical tidbit,

it might be mentioned here that the recognition of a time-lag between visual stimulus and response, a lag suggesting theoretical use of a differential equation below, evidently had been achieved at least as early as 1613: Le Pere d'Agui11on reportedly noted that in general the optical image vanishes with removal of the object, but that a motion impressed [by the image] on the visual system persists in time (in Plateau, 1878, Sect. I, p. 6). Vogelsang (1928, pp. 141-l43) apparently was the first to attribute the Hess effect to its namesake, Carl von Hess of Wurzburg, Germany. As reported by Vogelsang (1928, pp. 142-143),

the Hess effect originally was used by Pulfrich (or perhaps Pulfrich's associate, Fertsch) to explain the fami1iar Pulfrich effect, an effect not widely known during the 1920s (see Lit, 1949). A purely phenomenal effect closely related to the Hess effect is the "fluttering heart" illusion in which saturated-red heart-shaped pieces of paper pasted onto a deep blue background appear to pulsate when the array is moved back-and-forth before

J. M. Williams

Luminance and the Hess Effect

55

the eyes.

The pulsation, upon close inspection, is found to be

an irregular lateral displacement of parts or all of the hearts against their background. A history of the fluttering-heart

illusion is given in Roufs (1974), and a possible confusion of nomenclature is pointed out by McDougall (1905): McDougall (1905)

suggests that the "true" fluttering-heart illusion is obtained at photopic levels of illuminance only, and that a similar scotopic illusion due to rod mediated afterimages should carry a different name. Cursory observations by the present author (unpublished)

suggest that the fluttering-heart illusion (or something fitting the description) can be obtained under mesopic conditions in which both the redness of the hearts and their fluttering would be reported. An exhaustive series of experiments by Szili (1892) of the

stimulus conditions in which a fluttering-heart illusion could be obtained included conditions under which light- and dark-grey pieces of paper were used. The papers, which differed only in

reflectance and therefore only in luminance, were found to appear displaced so that the darker piece consistently lagged the lighter piece in the direction of motion. Attempts to measure the mag-

nitude of this effect were made by moving the array above a set of rulings (Szili, 1892, pp. 381-383). A similar phenomena1 study was done by Charpentier (1893), who used a rotating opaque disc with a narrow, radial, open sector covered with glass filters of various colors. The sector was

visible to the observer constantly and was illuminated by sunlight from behind. The sector could be varied in angular width, but

J. M. Williams

Luminance and the Hess Effect

56

evidently the varicolored subregions of the sector could not be displaced relative to one another in the direction of (rotary) motion. With different filters covering different extents of

the sector, the sector usually was seen to separate, the region(s) of the sector of higher luminance (it must be inferred) appearing to advance in the direction of disc rotation. Although Char-

pentier (1893) was investigating possible effects of different rise-times of sensation as functions of color and claimed to have equalized the regions of the varicolored sector in luminance (Charpentier, 1893, p. 569, "L'intensit lumineuse tait rendue comparable . . ."), it seems likely in view of later work (Guth, 1964) that Charpentier's photometry was not accurate enough to eliminate luminance differences, the art of photometry not having been standardized until after Ives (below). The criticism of

faulty photometry also was raised by McDougall (1904, p. 105), who found no wavelength-dependent latencies in similar experiments. In any case, Charpentier (1893) used visual estimates of

the apparent relative displacements of the regions of the sector to calculate latency differences (or persistence-differences, possibly) among the colors. To obtain the data used in these

calculations, Charpentier (1893) simply recorded, under various conditions, what part of adjoining sector-widths seemed to be overlapping. The apparent overlap then was used as a distance

in the formula, latency = distance/rotary speed. A problem with Charpentier's phenomenological approach is

J. M. Williams

Luminance and the Hess Effect

57

that any real (or apparent) displacement greater than 0 must be considered subject to various illusions (e.g., the Brown effect, Ansbacher, 1944; see also Coren & Ward, 1979) in addition to any illusion(s) being measured. In the Brown effect, for example, a

circular arc-segment subtending 30 or more of a disc is found, upon rotation of the disc, to shrink, Doppler-like, to a mere apparent point of light. There is no reason to doubt that the

apparent width of an illuminated sector will change upon changing the speed of rotation of the disc: For example, if, as in Char-

pentier's (1893) work, latency be computed on the basis of a phenomenal lining-up of the leading edge of one sector with the trailing edge of the other, there can be no guarantee that the two leading edges of the sectors also should appear displaced by any wel1-defined apparent distance. When no physical measure of the

apparent displacement is available, only an apparent latency can be computed. An improvement may be made by requiring judgements

based on the apparent alignment of the two leading (or lagging) edges of the sectors: An apparent displacement of 0 at least

cannot be distorted; furthermore, the criterion of alignment is the more nearly defined for the observer, the more identical the aligned edges appear when viewed individually. Although the magnitude of his visual latencies eluded his attempts at measurement, Charpentier (1887) did attribute the major visual latencies and persistences to events in the retina; as discussed above, this belief was well-founded. Carl von Hess's name is associated with the phenomenon which

J. M. Williams

Luminance and the Hess Effect

58

is the subject of the present Dissertation by virtue of several successful experimental approaches yielding objective and reliable data. After phenomenological observations suggesting shorter

absolute latencies under conditions of higher target retinal illuminance (or light adaptation), Hess (1904, p. 231), as translated by the present author, writes: Against a black background and viewing with a dark-adapted eye, it was possible to use as stimulus [a flat cardboard strip] made of two closely-joined halves, one of light grey and the other of dark grey. Under favorable circumstances, the dark grey half of the strip was found very distinctly to lag behind the light half when the strip was moved [in the direction of the joining line] (v. Plate 2). . . . It is practical to perform these experiments with very narrow stimulus-strips (2 to 3 mm wide); the two halves of the strip easily may be made to move relative to one another by use of a vernier-controlled mechanism. Knowing the speed of motion makes it possible to derive the latency difference involved. In the present context, the data show . . . that an achromatic excitation ["Erregung"] of 1ow stimulus-intensity ["Lichtstrke des Reizlichtes"] can appear visibly more delayed than one of higher intensity. Hess (1904, p. 231) then goes on to describe a different mechanism which allowed better control of target luminance, but which sacrificed the vernier-controllable match above for a measure of apparent target separation no better than Charpentier's A slot in the rotating drum [illuminated from within] is set to 1/2 cm in width and about 6 cm in length ["length" being parallel to the axis of drum rotation]. The slot is backed with translucent white paper and half of its length is covered with smoke-darkened glass. . . . Then, when the drum was made to turn with an angular velocity which made the period of rotation about 1.3 sec, the darkened half of the slot appeared [to lag the brighter half] so that the leading edge of the

J. M. Williams

Luminance and the Hess Effect

59

darker half lined up with the trailing edge of the brighter half. As the circumference of the drum totalled 62 cm, this observation yielded the result that the darker half of the slot provided an excitation ["Erregung"] which appeared delayed by about .Ol sec relative to the brighter half. It cannot be much more obvious that this example illustrates a definite decline of excitation for which a valid measure exists. The delay depends on the difference in intensities ["Lichtstrken"] of illumination as well as on their absolute levels; furthermore, it depends on the state of adaptation of the eyes as well as on the use of a fixation point allowing extrafoveal observation . . .. It might be noted that the explanation given by Hess in the first quote above rests upon latency, while the second seems to involve persistence of the trailing edge of the brighter half of the slot. Hess's calculation of the apparent latency difference involved in the effect was no different than Charpentier's and others; but the vernier apparatus allowing match-quality data to be gathered, the recognition that stimulus luminance (and not wavelength) governed the effect, and the careful study of the conditions under which the effect could be measured justify attribution of the effect to Hess by Vogelsang (l928, p. 141). McDougall (l904, p. 105) also reported the Hess effect for phenomenal observation similar to that of Charpentier as discussed above. However, nothing so accurate as Hess's (l904)

vernier apparatus appeared again until Ives's (1917, 1918) papers on "visual diffusivity." Ives was concerned especially with flicker photometry and eventually developed a theory (Ives, 1922) of visual temporal response. This theory was based on three hypothetical serial

J. M. Williams

Luminance and the Hess Effect

60

stages of retinal processing:

(a) a (reversible) photochemical

stage, (b) a diffusion stage of heat-like conduction governed by the heat equation of differential calculus, and (c) a perceptual stage governed by a threshold rate of change of the diffusion response. In a set of experiments which essentially were improve-

ments on Charpentier's, Ives (1917, 1918) obtained the Hess effect by using a rotating disc. This stimulus-disc was constructed of The

sheet metal and could be rotated at a constant low speed.

stimulus-disc consisted of an outer annulus which could be moved against an inner disc and locked in place at any given angle of displacement. Both the inner disc and the outer annulus had

a slot which could be aligned with the slot in the other by proper positioning of the outer annulus. The slots were narrow and ran

along a radius of the stimulus-disc when aligned; either or both slots could be transilluminated by a source concealed from the observer, and the exact amount of displacement of the inner disc relative to the outer annulus could be read from a scale on the disc. The apparatus is described in full detail in Ives (1917).

In the Hess-effect experiments, Ives's stimulus-disc was set in motion, and the observer was instructed to fixate the axis of the disc. The slots were illuminated with light of different intensity

and/or spectral composition, and the annulus and inner disc then were varied in physical displacement until the observer reported that the slots appeared to be lined up along a single radius. magnitude of the (Hess) effect then was read off in terms of the physical misalignment (if any) when the slots apparently were The

J. M. Williams

Luminance and the Hess Effect

61

aligned.

The disc had to be stopped each time the physical The obtained data could

displacement of the slots was changed.

be used to compute the theoretical latency by the formula used by Hess (above; but for real latency), which has the same form as expression (l) above. A decade after the Ives studies, Edmund (1928) explored a binocular analogue of the Hess effect. The main apparatus was

a black velvet ribbon mounted as a vertical belt on a motorized pulley; the belt carried a narrow, sewn-on white strip of cloth. One eye was used to view the belt directly, and a base-in prism was used to produce an image of the belt in the other eye. strip and its image thereby were made to move parallel to one another and in the same direction, as seen by the viewer. With The

eyes carefully fixated, Edmund (1928) was able to study latencies induced by filtering the light to one eye. The assumption was

that any visible disparities must be caused by temporal effects of the filter upon visual response to the moving images. Edmund

(1928) reported that although a filter produced some apparent delay of the leading edge of the white strip, the major phenomenal effect was a spreading-out and prolonging of the trailing edge as seen by the filtered eye. Edmund (1928) suggested that

the Pulfrich effect occurred because the eyes respond mainly to the very different trailing edges of the target percepts. would be a persistence more than a latency explanation. This Edmund's

(1928) dichoptic presentation of stimuli raises questions not

J. M. Williams

Luminance and the Hess Effect

62

necessarily related to the phenomenal nature of his data:

The two

eyes may be assumed to differ in the transmittance of their ocular media including their macular pigment, if any (see Norren & Vos, 1974). Dependence upon dichoptic data for information on ret-

inal (as opposed to other CNS) latency entails the extra inference that, given these known interocular differences, the normal visual system develops so as to compensate for them (if large) or to suppress their effects below a jnd (if small). Theories, as

in Alpern (1968), have been constructed to assume that equal illuminances yield equal latencies at all binocularly-paired levels of processing. The first relatively recent study of the Hess effect was that of Guth (l964). Guth used a belt of opaque black movie

film which had paired, misaligned slots cut out of it at regular intervals. The belt was set horizontally in motion, and spectrally

narrow-band lights of same or different dominant wavelength were presented in the slots. The observer's task was to adjust the

intensity of light in one of the slots until the pair seemed to line up, as in a vernier acuity task. Different physical align-

ments of the slots could be obtained by presenting the same lights in a different set of trials using a different belt of film. Giv-

en the slot radiances and filter characteristics at match (apparent alignment), Guth (1964) was able to determine slot relative luminances required to produce equal (inferred) latencies in the two slots; these relative luminances were found to fall approximately along a photopic visibility curve as determined by

J. M. Williams

Luminance and the Hess Effect

63

brightness-matching and threshold in the same apparatus for the same observers. Guth (1964) therefore concluded that any pre-

vailing latency differences could be attributed to luminosity but not wavelength differences in the stimuli, consistent with earlier opinions of McDougall (1904). Of course, in retrospect,

a change in luminance-contrast occurred in Guth's study with every passage of the target image across any given retinal point--in Guth's (1964) as in every study cited here so far--so wavelength differences, in the light of current knowledge, would not be expected to have had any effect, as was explained in the section on reaction time above. In a binocular study of a Hess-effect analogue, Prestrude and Baker (1968) used two rotating images of white discs, each disc having a black line segment inscribed partway along a radius. When stereoscopically fused, the combined view was of a single disc with a rotating radial line. Inserting a filter over one eye

than caused an apparent split in the fused radial line, the two disparate halves corresponding to the two line segments of the two fused discs. By changing the relative spatial phase of the

discs, the observer could advance the apparently lagging line segment until the two segments again appeared as a single radial line; the amount of phase-advancement could be recorded by the experimenter. A latency-difference due to the filter thus could

be estimated (formula (1) above) with something on the order of the precision of vernier acuity for moving targets. Prestrude

J. M. Williams

Luminance and the Hess Effect

64

and Baker (1968) compared the precision of their method favorably with the precision attained by Lit (1949) for the Pulfrich effect and by Roufs (1963) for two-flash simultaneity. It should be

noted that the dichoptic presentation of Prestrude and Baker (1968), as that of Edmund (1928), strictly does not test vernier acuity because no vernier alignment of any two physical objects is involved. The Prestrude and Baker task would seem to fall into

a category of hyperacuities (Westheimer, 1979) for dichoptic coincidence (or disparity) without depth. A later study was performed (Prestrude & Baker, 1971) in which white 1ine-segments were inscribed on black discs and the resultant latency-differences, inferred as above, were compared experimentally with those of the original configuration using black line-segments on white discs. These latency differences were

found to depend strongly upon retinal adaptation level set by the larger area of the background rather than the negligible areas of the line segments (recall Helson, 1948), although the dependence upon area as such was not investigated. Furthermore, an intense

monocular bleach was found to produce a smaller latency-difference when the black-on-white stimuli were presented than when the white-on-black stimuli were. This last result suggested that the

bleach had provided an "equivalent background" which somehow combined with the physical background light so that the net total adaptation level was higher for a white than a black disc. This

"equivalent background" approach, in which the retinal effects of a bleach are transformed into the effects of an illuminated back-

J. M. Williams

Luminance and the Hess Effect

65

ground (the Crawford transform; Barlow, 1964), was confirmed by Prestrude and Baker (1971) by further studies in which the time-course of recovery from the initial bleach was examined: For induced

latency difference t, the differences in obtained t's for white-on-black versus black-on-white at the same (unique) difference in retinal illuminance were found to be equal to the differences in obtained t's at the same (unique) time after termination of the bleach. The Crawford transform of photopigment bleach into

adaptation level also has been used to explain effects of preadaptation on latency in the Pulfrich effect (Rogers & Anstis, 1972). A recent study of the Hess effect was done by Wilson and Anstis (1969), who used small moving shadows cast in two parallel horizontal slots in an opaque fieldstop immediately backed by a translucent, diffusing screen. Each slot and its immediately-adThe shadows were

jacent background could be filtered separately.

cast by vertical rods driven by synchronous motors and could be adjusted in phase by the observer until the shadows seemed to be aligned. This apparatus, which unfortunately does not allow

independent control of target and background luminance, has yielded some interesting generalizations: Wilson and Anstis

(1969) graphed their data to show agreement with data of Lit (1949) on the Pulfrich effect; this agreement allowed a fit of the obtained latency differences t by a power function of the form t = a I-b + t0 . (16)

J. M. Williams

Luminance and the Hess Effect

66

This expression is equivalent algebraically to Pieron's formula (9a) above. Written as in (16), the exponents b obtained by

Wilson and Anstis (1969) were near 1/3, possibly coincidentally agreeing with typical such exponents obtained in scaling brightness against intensity (e.g., Krantz, 1972, p. 675; Stevens & Marks, 1980) and agreeing with such previous data on visual latency as the electrophysiological results summarized by Riggs and Wooten (1972, p. 708) and given by Mansfield and Daugman (1978) for the toad eyecup. Particularly noteworthy is the negative exponen-

tial-like graph of data in Wilson and Anstis (1969, Figure 2b): This graph is remarkably similar to Bernhard's for alpha-blocking latency in humans (Bernhard, 1940, Figure 20) and to another graph of optic-nerve latency in frogs (Bernhard, l940, Figure 22). Similar graphs appear in Bartley (1934) and in Prestrude (1971). Although a great deal of the rhetoric of such appearances derives power from an aesthetic commonality among research-oriented figure-compositors, undoubtedly there is even a more important relationship in that all three such well-scaled graphs derive from responses depending upon the same vertebrate retina. 2. PROPOSED THEORY. The approach is based in part on conversations with Alfred Lit on the subject of visual persistence. The intent is to

specify a source for the latencies assumed to govern spatiotemporal illusions such as the Mach-Dvorak, Hess, and Pulfrich effects. Geometric considerations of stimulus and response are

J. M. Williams

Luminance and the Hess Effect

67

combined with postulation of a CNS process determining the magnitude of the effect measured. It might be mentioned again (see

above) that for sufficiently brief target exposure duration, the Mach-Dvorak effect seems to depend solely upon the stimulus geometry: The only relevant "latency" is the interocular delay

between shutter exposures; target lateral movement during the unexposed period of the shutter cycle determines a retinal disparity evidently generating the depth illusion as well as an apparent target movement. Only if exposure duration be pro-

longed enough to obtain an ED effect (Lit, 1978) would the following theory apply nontrivially in the Mach-Dvorak effect. The basic approach in the present theory is the specification of the form of the function f(L, L) in equations (2) above. The solution proposed fulfills the requirement of Roufs

(1974, eq. 10) that arguments in intensity I (proportional to luminance L or estimated retinal illuminance E) be logarithmic. Because of the present subject matter, the following will be cast entirely in terms of the Hess effect, in which the task is vernier alignment of two moving targets imaged on the same retina. With no illusion present, the stereoscopic depth settings of the Pulfrich effect depend on a matching of disparities which will be assumed geometrically the same as the vernier matching of the Hess effect. Both illusions thus are assumed to depend on the same

geometric latency difference t as discussed in the section on Spatiotemporal Illusions above.

J. M. Williams

Luminance and the Hess Effect

68

a. A DISTAL STAGE OF DIFFUSION. Literature reviewed above leaves hardly any doubt that the intensity-dependence of the time-course of visual response is determined primarily at the retina. Behavioral tasks (such as reaction

time) in which different retinal image intensities control specific, distinct motor responses may be confounded by a nonvisual intensity-dependent component (see Brauner & Lit, 1976), but for the illusions just named above, the (inferred) latency differences may be assumed to occur in the visual system and hence mostly at the retina. Thus, following approaches beginning with Ives and Kings-

bury (19l4) and Ives (1922), and later taken up by Veringa (1961, 1963), Chapman (l96l), Cone (l964b), Kelly (l969b, 1971), Penn and Begins (1969), and Kelly and Wilson (1978), it seems reasonable to postulate that the intensity-dependence of retinal latency should find its origin in a diffusion process. The present approach may be derived from the linear diffusion equation D
2

u(x, t) - u(x, t)

u(x,t) t

(17)

in which u(x, t) represents concentration of some diffusible substance in a small region around location x at time t. theory proposed will involve a spatially degenerate case of equation (17) to describe the initial visual response, with a nonlinearity (response centroid) determining later CNS processing ultimately leading to vernier alignment of the Hess-effect targets. The

J. M. Williams

Luminance and the Hess Effect


2

69

It should be explained briefly that in equation (17), the symbol, represents the laplacian differential operator

D represents the diffusion coefficient describing diffusivity (here assumed constant) in the region in which equation (17) is applied, and represents a loss coefficient describing the absorption, loss, or chemical binding of the substance of concentration u. The laplacian in general is zero in any region

in which the spatial gradient [u(x, t)]/x is constant or zero; i.e., wherever no new substance is being produced at a rate different from the rate in any neighboring region. It is possible to give a physical interpretation to the sign conventions in equation (17) above: The leftmost (lap-

lacian) term always must be opposed or neutralized by the middle (loss) term, as witness the minus sign. The rightmost term

represents the time-changing result of the balance of the two terms on the left, in any region near x. Various mathematical

properties of the linear diffusion equation are discussed in Webster (1955, pp. 166-173), Bracewell (1965, ch. 17), and Broman (1970, sections 8.2-8.4). Rashevsky (1938, introduction)

derives equation (17) in the context of cellular metabolic turnover, and Landahl (1962: pp. 295-302) applies related approaches to various psychological fields. Fick's second law in chemistry

(e.g., Reed-Hill, 1973) may be treated as a case of equation (17), and the Hodgkin and Huxley (1952) equations for axonal impulse propagation can be combined and rewritten as a non1inear case of equation (17). Nonlinear diffusion in biological systems has been

J. M. Williams

Luminance and the Hess Effect

70

reviewed extensively in Scott (1975). The data of the present study allow no measurement of any diffusion process which might occur in the retinal photoreceptors. Work by others, however, may be used to set boundary

conditions (geometrically highly degenerate) on equation (17) above which greatly simplify the problem. For turtle cones, at least, the response at the outer plexiform layer would seem to be spatially linear, meaning independent of spatial stimulus pattern, at least for low-Hz stimulation (Tranchina, Gordon, Shapley, & Toyoda, 1980). Lateral current

spread in rat rods has been found to be small compared with the radial (lengthwise) current response (Penn & Hagins, 1969; see also Arden, 1976). So, although phenomenal and physical measures of

human behavioral response suggest that lateral inhibition should be a useful concept (e.g., Ratliff, 1965; Kelly, l969a), such would seem unnecessary in quantifying photoreceptoral and other very distal retinal temporal responses which might usefully be describable in terms of diffusion-changes initiated by light, at least if those changes be as measured by electrical currents. The single most relevant study providing data on a possible distal diffusion process would seem to be that of Bay1or, Lamb, and Yau (1979), who studied light-initiated responses of single toad rod outer segments (ROS) drawn into a micropipette but otherwise apparently left undamaged. In that study, the response

amplitude (hyperpolarization) was found to be proportional, for

J. M. Williams

Luminance and the Hess Effect

71

uniform stimulation of the entire ROS, to the length of ROS left exposed to the surrounding medium. Furthermore, Baylor, Lamb,

and Yau (1979) found that stimulation with a microscopic image of a slit evoked the same amplitude of hyperpolarization regardless of where on the ROS the image was located. At the region

near the connecting cilium and protoplasmic bridge of the ROS-inner segment junction, however, the time-course of the response was more rapid than it was near the ROS distal tip. Such a situation

(diagrammed in Figure 4 below) might well be examined in terms of equation (17) above: The diffusion of some internal messenger-sub-

stance such as Ca2+ or a cyclic nucleotide, presumably proceeds from one or more sites near the light-responding chromophore to the plasma membrane of the ROS. The delay of this internal, rad-

ia1ly directed process presumably is negligible in the total time scale of intensity-dependent visual latency. A resultant ROS

plasma-membrane hyperpolarization then occurs relatively slowly and is what is measured in the experiment. It only need be

assumed that the longitudinal propagation of the messenger is considerably more rapid than its radial propagation to explain Baylor, Lamb, and Yau's (1979) results; equation (17) above then would be applied to diffusion of closed plasma membrane pores at the ROS plasma membrane. The required lengthwise distribution

of the initial photic response might be assumed to be related to the constant, depolarization-dependent Na+ inflow, which seems not impeded by the ROS discs; it might even be secondary to some

J. M. Williams

Luminance and the Hess Effect

72

mechanical change in the ROS plasma membrane in relation to the ROS discs; for example, a decrease in total ROS volume or a retraction of the ROS ciliary microtubules. Baylor, Lamb, and Yau (1979) suggest that lengthwise diffusion of some substance originating at any single ROS location probably could not explain their data, a conjecture which would rule out boundary conditions on equation (17) above such as those proposed by Kelly (1971, Appendix A). The problem

pivots on the obtained apparently uniform response amplitudes and the rapider response in the proximal ROS; it assumes no preparation artifact such as a change in probe-ROS capacitance with change in length of ROS in the micropipette. The problem

also depends upon constant diffusivity over the ROS length, a defensible assumption for the more distal 2/3 of a typical ROS as based upon electron and light micrography available in the literature. The author's suggestion in the second paragraph before the present is that, in effect, the phase velocity of interaction of some internal messenger must be extremely high along the length of the ROS and highest at the most proximal ROS, where diffusivity and/or velocity might increase to quicken the response as observed. Once ROS plasma membrane pores have closed, it is assumed in effect that they distribute themselves on the ROS in accordance with equation (17) above. In any event, either the slow ROS plasma-membrane hyperpolarization itself or some closely-related response of the

J. M. Williams

Luminance and the Hess Effect

73

photoreceptor will be presumed to cause the intensity-dependent visual latency differences which determine the magnitude of the Hess effect. assumed. Now, again consider the linear diffusion equation (17) above: The question is whether boundary conditions might be found An underlying diffusion mechanism will be

consistent with Baylor, Lamb, and Yau's (1979) data (assumed applicable to the human retina) and sufficient to explain the Hess effect. First of all, as the length x of the ROS (and possibly

the length of the inner segment) seems to be the only spatial dimension of importance, equation (17) above might as well be rewritten in one-dimensional form, with x a scalar: D 2u(x, t) x2 u(x, t) = u(x,t) t , (18a)

Secondly, as illustrated in Figure 4 below, a hypothetical depictation based on Baylor, Lamb, and Yau (1979), both before ("t1") and after ("t5") photic stimulation has generated its maximal response, the spatial gradient u/x is zero over the ROS. Thus,

either in the hyperpolarized or fully recovered (steady) state, the leftmost term in equation (l8a) above is found to be near zero for the Baylor, Lamb, and Yau (1979) preparation. But even in

intermediate response states, it seems possible to assume that the function u(x, t) be at most quadratic in x, so that u/x must be at most linear and thus 2u/x2 constant or zero. This

J. M. Williams

Luminance and the Hess Effect

74

approximation (subject to experimental test) makes it possible to reduce equation (17) to the form

(some constant or 0) -

u(x, t)

u(x,t) t

(l8b)

The approximation might also be applied piecewise for, say, distal and proximal halves of the ROS, assuming that the CNS response (below) depended on only one piece. Finally, the lack of a spatial derivative in (l8b) allows the result to be rewritten as an ordinary differential equation of first order--in homogeneous form, with the postponement until later of a system input: Thus, from equation (17) we have

[u(x,t)] t

+ u(x, t)

= 0; or, (19)

d u(t) dt

+ u(t)

= 0.

b. SMALL RETINAL REGIONS RESPOND. Equation (19) above supplies the single-unit response of the retina presumed to originate the latency difference of equations (2) above. The logarithmic form of the stimulus input

alluded to by Roufs (1974) is not inconsistent with data of Baylor, Lamb, and Yau (1979), the (human) retinal input being

J. M. Williams

Luminance and the Hess Effect

75

essentially the same at the single-unit photoreceptor level as that of the toad single ROS response. As shown in Baylor, Lamb,

and Yau's (1979) Figure 4, there is approximately a logarithmic dependence of ROS plasma membrane peak transconductance upon stimulus intensity for about a two log-unit intensity range above threshold but below response saturation. This means that to be

consistent with Baylor, Lamb, and Yau's (1979) results, the human retinal input should be considered as applied not to single photoreceptors but rather to small retinal regions containing several photoreceptors: Thus, a linear response (nonzero and nonsaturated)

by at least one photoreceptor may be assumed with amplitude proportional to log stimulus intensity in every such small retinal region; the time-course of this response will be assumed to be governed by the degenerate diffusion equation (19) above. conclusion is represented in Figure 5 below. This

c. GEOMETRY OF THE CENTRAL CENTROID. The system governing retinal latency having been established, it now is necessary to consider how the vernier match of the Hess task might depend upon it: This requires nothing more

than the assumption that the observer align the moving targets of the Hess apparatus according to the relative location of the target retinal images. The retina is known to project retino-

topically to more central regions at which the vernier alignment response must be originated. Based perhaps upon early learning

(including possible developmental factors), a human observer may

J. M. Williams

Luminance and the Hess Effect

76

be assumed to conform to the experimental instructions by recognizing a match whenever some two regions projecting ultimately from two vertically aligned retinal regions respond simultaneously to the two targets. The judgement of simultaneity is

assumed made by establishing equality of the two temporal centroids of some central response, one centroid corresponding to each target image. It might be mentioned that the centroid

(center of gravity) is a standard type of average which has been used to describe, in the temporal domain, blood flow and other metabolic measures (e.g., Po1issar & Rapaport, 1962, equation I.3.l). In any event, as the targets constantly are in

motion, any two retinal regions, one stimulated by each target image, could govern the vernier-match response; but presumably the closer such two regions were to being aligned vertically on the retina, the more precise would be the resulting vernier match. Let us call t the departure from simultaneity of the two centroids just mentioned; we also call the CNS locus of comparison R. Vernier alignment then will be a response made Let

by the observer which reduces t to zero at location R.

the central time-varying response to the image of the upper target be called gu(t) and that to the lower gl(t): These res-

ponses are referred to specific small, vertically-aligned retinal regions and are kept distinct by the retinotopic organization of the visual CNS (see Figure 5). The centroid for gu(t) will

be called t(gu) and that for gl(t) will be called t(gl).

J. M. Williams

Luminance and the Hess Effect

77

The result may be written,

t(gl) - t(gu) .

(20)

Now, for identical, aligned targets of equal luminance, as the moving images sweep across the retina the time-course at the two aligned regions respectively generating gu(t) and gl(t) will be the same, t(gu) will equal t(gl), and at locus R we shall have

t(gl) - t(gu)

0,

(21)

and the observer will report the targets to be in vernier alignment (= an alignment or match response). If the upper target have a higher luminance, so that a phenomenal Hess effect occur, for aligned targets we shall have

t(gl) - t(gu)

>

0 ,

(22a)

and no match response:

Here, given control of relative target

lateral displacement, the observer will displace the lower-illuminated target ahead of the upper until t(gl) be reduced (made earlier) to the point that

t'

t(gl') - t(gu')

again,

(22b)

as in equation (21) above, so that a match response occur. Now, consider the difference between the two situations

J. M. Williams

Luminance and the Hess Effect

78

described by equations (22a) and (22b). ures the Hess effect. we obtain

This difference meas-

Subtracting equation (22a) from (22b),

t' - t

[t(gl') - t(gl)] - [t(gu') - t(gu)]

>

0 .

(23)

Using the lower target as the standard, this is the same as

t' - t

t(gu) - t(gu')

>

0 .

(24)

And the immediately preceding moreover requires

t' - t

t(gl') - t(gu') ,

(25)

because, as above, t(gu) = t(gl) = t(gl'). Calling the measure of the Hess effect t' - t = t and applying the definition of centroid, from equation (25) we obtain in general

t(gl) - t(gu)

tglt glt

tgut gut

(26)

J. M. Williams

Luminance and the Hess Effect

79

d. DERIVATION OF THE THEORETICAL t. The final stage in the construction of the present theory is to set the distal time-varying diffusion response u(t) equal (within some linear relation) to the central time-varying vernier response input g(t): This means that equation (19) determines

the central system in the form (du/dt) + u a(dg/dt) + g ; (27)

furthermore, it then follows that the time-varying troland value E(t) of the input to the retinal diffusion process determines the central response by the logarithmic relation ln[E(t)] = a(dg/dt) + g . (28)

The theoretical system is diagrammed in Figure 5. Equation (28) is solved in Appendix 3 below by elementary methods to yield
t

g(t) =

1 k exp [t/a ]lnEd expt/a . a a 0

(29)

In equation (29), let us assume white-on-black stimulation so that E(t) is near 0 except while the target image is impinging upon the small retinal regions generating g(t). The target width and

speed then will determine some fixed exposure duration t0 assumed such that ln[E(t)] is identically - outside the interval 0 < t t0. This makes the limits of integration in

J. M. Williams

Luminance and the Hess Effect

80

solution (29) = 0 to = t0.

The response the cent-

roid of which is defined by (26) cannot begin before E rises above 0 but my persist indefinitely: Thus, the limits of integration

in expression (26) will be t = 0 and t = . Substituting the solution g(t) of equation (29) into the definition (26) yields in general

t ln [ E ] exp [1/at ] d
t(g) =
0 0 0 t 0

{ {

k expt /a dt . (30)

ln [ E ] exp [1/at ] d k expt/a dt

} }

The outer integration in expression (30) cannot be performed as given because the integral in (29) diverges for t beyond t0. Therefore,

it is necessary, at least by the present approach, to define the integral in expression (29) to be identically equal to its value on the domain 0 < t0 for any domain the upper bound of which exceeds t0. The rightmost term in numerator and denomin-

ator of equation (30), the decay or persistence term, then may be integrated separately out to t = infinity. Thus, for the lower

target, expression (30) must be evaluated in the form

J. M. Williams

Luminance and the Hess Effect

81

t(gl)=

t t t ln [ E ] exp a dkexp a dt 0 0
t0 t

t0

{ {
t 0 0

ln [ E ] exp

[ [

t t dkexp dt a a

] ]

} }

kt exp
t0

t dt a

kexp
t0

t dt a

(31)

The integrals added at the right side of numerator and denominator of expression (31) represent pure "decay" terms; these give responses of the system which persist after stimulation by light has ceased.

The value of t(gl) is given by expression (31) above for some particular location on the retina stimulated by the moving image of the lower target; the same expression also may be assumed to hold at any other similar location on the same retina (or on the retina of the other eye of the observer). Suppose all to be

the same at some other location except that the image of the upper target be the source of stimulation: Then equation (31)

should hold unchanged at the new location except that E() would be replaced by cE(), in which the constant c l is defined as the ratio of new to old target luminance. By defining c thus,

J. M. Williams

Luminance and the Hess Effect

82

we establish a convention of notation by which the less intensely illuminated target always generates the centroid (31) above and the more intensely illuminated target generates a similar centroid as follows:

t0

t
t(gu)=
0 t0 0

{ {
t 0 t 0

ln [ cE] exp

t t ln [ cE ] exp dkexp dt a a

[ [

t t dkexp dt a a

] ]

} }

k texp
t0
0

t dt a

t k exp a dt t

(32)

For the present experiment, the retinal image is of a sharply defined, uniformly illuminated target rod, as will be described below. The moving target-image thus will produce

(ignoring optical imperfections) a rectangular temporal pulse of illumination as the image sweeps across any given retinal point. Therefore, we may assume E() identically equal to some

constant value E whenever it is not zero; namely, for = t such that 0 t t0. Constant target thickness and con-

stant target speed always make t0 a constant, too.

J. M. Williams

Luminance and the Hess Effect

83

With these constraints, expressions (31) and (32) above may be integrated and simplified as follows:

From (31),

t(gl) =

ln(E) [t02/2 + at0 exp(-t0/a) + a2exp(-t0/a) -a2] + ka ; ln(E) [t0 + a exp(-t0/a) - a] + k

(33)

and, from (32),

t(gu) =

ln(cE) [t02/2 + at0 exp(-t0/a) + a2exp(-t0/a) -a2] + ka ; ln(cE) [t0 + a exp(-t0/a) - a] + k

(34)

Equation (26) above therefore becomes

t(gl) - t(gu)

K1 ln(E) + k a K2 ln(E) + k

K1 ln(cE) + k a K2 ln(cE) + k

(35)

in which

J. M. Williams

Luminance and the Hess Effect

84

K1 and K2

t02/2 + a t0 exp(-t0/a) + a2 exp(-t0/a) - a2 ,

(36)

t0 + a exp(-t0/a) - a .

(37)

Using the same abbreviations (36) and (37), equation (35) may be rewritten algebraically as t = k ln(c)(a K2 - K1) [K2 ln(E) + k]2 + K2 ln(c)[K2 ln(E) + k] , (38)

which has the form of a rectangular hyperbola in ln(c). Equation (35) also may be written as t = ln[EK1 exp(k a)] ln[EK2 exp(k)] ln[(cE)K1 exp(k a)] ln[(cE)K2 exp(k)] . (39)

The final theoretical expressions for t, (35), (38), and (39) above, show qualitatively the characteristics which might be expected of valid visual-latency formulas: Let t0 approach 0 in (36) and (37); then in (38) we have a K2 approaching K1, which makes t go to 0. Thus, as target

speed is increased or target thickness decreased, the theoretical t also decreases, as observed in data of Lit (l96Oa, b). As

discussed above on p. 20, the rather mild variation with v of the approximate linear relation between t and log Eu (= log10(cE)) is therefore, theoretically, the hyperbolic relation between t and t0 seen in equation (38). This hyperbolic relation, not at

all equivalent to a simple hyperbola derived from a conic section, can be seen to be difficult to handle beyond the qualitative level of the present discussion.

J. M. Williams

Luminance and the Hess Effect

85

Let t0 approach infinity; then dividing numerator and denominator in (35) by K1 shows that t also increases without bound (not inconsistent with Lit (l96Oa, b)). Let c approach 1; then as in (38), t approaches 0 (as expected). Let c approach infinity; then dividing numerator and denominator in (38) by ln(c) shows that t approaches k (a K2 - K1) , 2 K2 ln(E) + K2 k (40)

an asymptote which is the higher the lower the illuminance level E of the lower target, a result consistent with data of Lit (1949), Alpern (1968), Prestrude (1971), and others. Let the time constant a approach 0; then from (35), t approaches -k t0 ln(c) , 2[ln(E) + k] [ln(c E) + k] (41)

some value less than 0 which approaches 0 as k approaches 0, which increases negatively as c increases, and which finally becomes positive as E approaches 0. No obvious interpretation

is attached to this behavior, which derives from the singularities in (35) where the denominators go to 0. Let a approach infinity; then (35), after some manipulation, shows that t approaches 0.

J. M. Williams

Luminance and the Hess Effect

86

Finally, letting k approach either 0 or infinity in expressions (38) and (39) makes t approach 0. Some meaning

may be attached to this behavior for k interpreted as a persistence or decay factor, but k may as well be considered a scaling constant determining the base of the logarithmic stimulus input, as referred to the Napieran base e. This latter

meaning makes k interpretable as a free parameter of no significance beyond the fact that k is implied by the initial postulation of a first-order differential system in the proposed theory.

J. M. Williams

Luminance and the Hess Effect

87

EXPERIMENTAL METHOD 1. SUBJECTS. Two paid female undergraduates, AB and LD, were available an average of four days per week for sessions lasting up to two hours per day. Both had 20/20 uncorrected visual acuity or bet-

ter, normal color vision, and normal depth perception as determined by a Keystone-industrial eye examination and by performance in research apparatus during premployment testing. As LD was obliged to withdraw from the study before it was completed, the present author (JMW) gathered extensive data on himself. JMW's myopia was corrected by -2.25 D eyeglasses; small

pieces of photographer's tape on the artificial pupil(s) allowed the eyeglass-lenses to be pressed against the artificial pupils without being scratched. An eyeglass-lens often was taped directSes-

ly to the artificial pupil for Hess and reaction-time trials.

sions for scotopic data on JMW often lasted more than four hours.

2. APPARATUS. a. REACTION-TIME AND PULFRICH APPARATUS. The reaction times were measured using a standard Scientific Prototype Model GB three-channel tachistoscope. Lamps were Sylvania F8T51D (low-pressure mercury daylight fluorescent) advertised to operate at a correlated color temperature of about 6500 K. Flash duration at peak was checked using a photo-

voltaic cell and oscilloscope, so that the nominal 86.8 millisecond

J. M. Williams

Luminance and the Hess Effect

88

exposure duration set at the tachistoscope control panel was verified correct within a few ms. At one meter viewing distance,

the illuminated, aligned targets, made from a black-painted thin sheet of clear glass, had the same visual-angle dimensions as those in the Hess apparatus. A red fixation dot, subtending about

2 min and flanked by three white pinpoints (one above the fixation dot and one on each side about 1/2 away), was located l.8 directly above the gap between the targets. The fixation dot was preFilters

sented in a side channel and was visible constantly to S.

from the same set as used in the Hess apparatus were placed just before both targets to vary target troland value. Both targets

always had the same troland value in the reaction-time experiment. A 2.7 mm artificial pupil yielded source target troland value of 2.48 log td as calibrated using an Ilford SEI exposure meter (blue spot) at the start of each session. A low, 1000 Hz warn-

ing tone preceded each target presentation by a 3- to 6-second interval varied randomly by E using a digital clock and logic circuit. The Pulfrich apparatus was as described in Lit (1949) and in Lit, Finn, and Vicars (1972), and was operated at a speed of 4 degree/s to match the speed of the Hess apparatus described below. The illuminated target rods, viewed at 1 meter, were

the same visual-angle size as for the Hess apparatus, although the gap between the Pulfrich targets was only about 3 min high. The

same set of neutral filters as for the Hess experiment was used, the 2.5 mm artificial pupils yielding a source target troland

J. M. Williams

Luminance and the Hess Effect

89

value of 3.48 log td for AB trials and 3.58 log td for JMW trials. b. HESS TARGET-MOVEMENT MECHANISM. A darkened, sound-attenuating cubicle was provided for S and a somewhat larger space for E and the apparatus. The S sat

in a comfortable chair and gripped a mouthbite made of dental wax. A 2.4 mm artificial pupil was mounted in a viewing tube; an aperture stop a few cm before the artificial pupil reduced stray light and facilitated centering of the eye at the artificial pupil. A pushbutton buzzer system was used to signal beginning

and end of each trial; an intercom allowed other spoken communication. The S had a set of two pushbuttons which controlled

lateral separation of the oscillating targets by energizing of a small, reversible direct-current (DC) motor as described below. The two major components of the Hess apparatus were the target-movement mechanism (TMM) and the lightbox (see Figure 6 below). Both components rested on a sturdy, black-painted table

to which was bolted a rail made of 3" angle-iron, 2 meters long. This rail lay on top of the table and was parallel to S's line of sight (i.e., to the optic axis of the apparatus). Bolts

could be tightened in slots in the rail to position the TMM so that the targets moved back-and-forth in a plane parallel to the S's frontal plane at a distance of 100 cm from the artificial pupil. The space beneath the table was enclosed in plywood and held controls and power supplies for the apparatus. The lightbox

J. M. Williams

Luminance and the Hess Effect

90

was made of galvanized sheet metal in which a narrow groove was cut to receive the rail just described. The lightbox could be

slid along the rail and rested by its own weight at a distance of about 50 cm behind the targets; thus arranged (as, for photometry) the apparatus presented S with a view of the targets silhouetted against a featureless, uniformly-illuminated field. For the

present experiment, a black, velvet-covered partition was placed behind the targets to provide a dark background. The target-movement mechanism proper (TMM) consisted of a welded angle-iron frame about 63 cm high by 56 cm wide by 42 cm deep; this frame enclosed and supported several smooth, straight guide rails and two precision-milled brass Yankee jackscrews all running parallel (at various distances) to the S's frontal plane (see Figure 6). Motive power was provided by a 1/3-horsepower

l2O volt alternating-current (VAC) single-phase electric motor and a variable-speed transmission (Zero-Max Model JK1, 0-400 rpm), not shown in Figure 6, which drove a belt-and-pulley (a of Figure 6) turning the upper jackscrew (b). The upper jackscrew

drove the lower jackscrew (b') through a chain linkage (c) between gears (g) and (g') as shown in Figure 6. A movable gear

(m) could be displaced up or down under control of S by turning a screw driveshaft (s). The screw was turned by means of a

geared-down l/10-horsepower reversible DC motor (mentioned above) mounted on top of the angle-iron frame but not shown in Figure 6. When gear m moved up or down, the length of chain between jack-

J. M. Williams

Luminance and the Hess Effect

91

screw gears g and g' was changed, thus changing the spatial phase relationship between the jackscrew shafts. This phase change

rotated the jackscrews slightly relative to one another and thereby displaced the targets relative to one another in a lateral direction as viewed by S. Except actually during a change of

phase, the jackscrews rotated together at the same constant angular velocity, thus maintaining a constant target speed of 7.03 cm/s in the direction of target motion, which corresponded to about 4.02/s visual angle. The phase between the jackscrews, when

constant, thus maintained a constant target relative displacement. The takeup arm (T, in Figure 6) maintained chain tension adequate to keep the chain on the gears; this tension was not constant but varied with takeup arm position, as discussed below with the electrical circuitry. The two counterpitched grooves of each jackscrew were joined together at the ends of each jackscrew. The targets t and t'

shown in Figure 6 consisted of illuminated glass rods 6.1 mm in diameter mounted on the filter-boxes of two rod-illumination optical channels similar to those which have been described elsewhere (Lit, Finn, & Vicars, 1972). These filter boxes were positioned on

metal platforms (p and p' of Figure 6) which in turn were joined to guide-blocks containing spring-loaded steel followers. These

followers rode in the grooves of the jackscrews, thus providing the forces which moved the targets. Two precision, stainless

steel guide rails (not shown) were inserted in linear ball bearings on each target platform to prevent variation in direction of

J. M. Williams

Luminance and the Hess Effect

92

target motion.

Each jackscrew thread (groove) had the same pitch

and number of turns and, for constant speed of pulley a, was machined with precision adequate to move the targets at constant calibrated velocity. The followers could not cross threads except

at the ends of the jackscrews, so the position of the movable gear m (Figure 6), as above, determined a constant spatial phase relationship between the two targets as they moved together back-and-forth. A pointer mounted on the casing of movable gear m (Figure 6) cast a shadow on a metric scale (r) made of a white piece of steel measuring tape. The lever-arm of the gears and the pitch of the

jackscrew threads gave this scale a magnification factor of about 5.1 times actual target relative displacement. The scale r could

be read by E to within a fraction of a millimeter, yielding an accuracy of better than 0.04 mm actual target relative displacement in the lateral direction. This accuracy was achieved only

for one direction of target motion (left-to-right, as viewed by S) because of inherent imprecision in the operating relation of the followers and the jackscrew grooves. In principle, the differ-

ence in scale readings left-to-right versus right-to-left could have been subtracted out as a constant error, but only if each datum had been recorded in terms of the direction of target motion prevailing when S made each match-decision. Such a procedure

would have gained very little but made datum-records and interpretation of match variability needlessly complicated.

J. M. Williams

Luminance and the Hess Effect

93

Located a few cm in front of the targets and functioning as a fieldstop was a horizontal rectangular aperture measuring 23 cm wide by 3.6 cm high. This fieldstop, as many other components,

has been omitted from Figure 6 for c1arity's sake; S's side of it was painted flat black and the target side was lined with black velvet to minimize stray target light. The visible extent of

each target as viewed at 100 cm was 57 min x 20 min visual angle; the gap between the tips of the rods was about .35 cm (12 min) high and was centered in the vertical direction of the field of view. The gap could be seen easily parafoveally except at the

lowest illumination levels of the present study (e.g., Brown, 1972). Blackened sheet-metal plates (not shown in Figure 6) were bolted to the front face of the angle iron frame of the TMM to reduce stray light and hide all movement of mechanical parts. A piece

of black felt was draped over the top of the TMM to dampen sounds and intercept light from the lightbox aperture. A small fixation

light (12 V grain-of-wheat bulb operated at 2.0 to 3.8 VDC) was glued to a magnet and mounted at the edge of the upper sheet-metal plate; this arrangement placed the path of the gap between the tips of the targets about 1.8 below the fixation point as the targets passed the center of the field of view. The 2 mm high

by 1 mm wide, v-shaped filament of the fixation light was kept dim but clearly visible (l or 2 log units above dark-adapted foveal threshold) during experimental trials. It was assumed

that even at scotopic levels, fixation accuracy would be within 1/3 of the filament center (Kandel & Bedell, 1976).

J. M. Williams

Luminance and the Hess Effect

94

Apparatus electrical power was supplied by two single-phase 120 VAC mains. outlet. The two mains fed from a 20 ampere, fused service

One main fed the lightbox line regulator: the other main The other

fed a second line regulator for the TMM main motor.

main also fed a third line regulator for all TMM illumination and other electrical power. The cooling fans for the lightbox and

target illumination lamps were fed from the line regulator inputs; everything else was fed from the line-regulator outputs, each of which had its own 20-ampere fuse. The 1/3-horsepower main motor could be switched off independently at any time. Main transmission speed was mechanically

determined and manually regulated by a turn-counting dial; transmission speed was reduced by E to a low value each time before the main motor was started, thus reducing cumulative mechanical decalibrations and any risk of damage to moving parts. The reversible DC target lateral-displacement motor was run by a homemade motor controller incorporating a filtered bridge rectifier for the field and a filtered half-wave rectifier for the armature. A 6-volt relay controlled by E or S (at E's option)

reversed field polarity to reverse direction of target lateral displacement. An autotransformer, with secondary in series both

with motor armature and with a half-wave rectifier diode, was controlled directly by E to set speed of target lateral displacement at a convenient level. Field and armature circuits both

were open unless a relay was closed (either by E or S) to cause

J. M. Williams

Luminance and the Hess Effect

95

the targets to be displaced.

Because energy was stored whenever

the takeup arm T (Figure 6) was elevated against force of gravity, the speed of target displacement always was slower in one direction (namely, that retarding the bottom target relative to the top target in the direction of target motion) than in the other. In

actual operation, it took 8 to 10 seconds to advance the bottom target 6 mm in the direction of target motion; it took 12 to 14 seconds to retard the bottom target by the same distance. Prior exper-

ience with a standard thyristor "constant-speed" motor controller compatible with the target-displacement motor had shown that the tension required (with weight W, Figure 6, = about 20 pounds) was too great for constant regulated speed, at least without radical modifications to the controller. Thus, trial-to-trial counter-

balancing of direction of initial target relative lateral offset during the present experiment could be made so as to equalize average distance of initial offset, on the average, or to equalize average time from offset to S's point of subjective equality, but not both distance and time. All controls operated by S were fed from a standard 8 VAC (rms) doorbell-transformer secondary and series power resistor. The secondary was grounded, and the circuit, in use, supplied 5 or 6 VAC to the signal buzzer and/or control relays. Target illumination was by two manually regulated 28 VDC power supplies, one per target rod (see Lit, Finn, & Vicars, 1972). For the upper target, the supply was an Electro model EF; for the lower, an Epsco model EF. Both were operated in the "0-28 V" range

J. M. Williams

Luminance and the Hess Effect

96

and are advertised as having filtered output for 5-ampere continuous duty with less than l percent ripple. less than 2 amperes during actual operation. The lamps each drew Resolution of the

potentiometer controls for these supplies allowed repeatable settings within .2 VDC, as read from a voltmeter mounted on each power supply. Photoresistors (Clairex CL5P9M) permanently mounted in

heat- and light-insulated cells were connected to optic fiber bundles implanted in each of the target lamp housings; these were used to regulate target illumination by reading their resistances with an ohmmeter (Triplett, 20,000 ohm/volt, 7-inch scale) on R x l K scale. Another photoresistor of the same type was mounAll photoresistors were adjusted perman-

ted in the lightbox.

ently (by location, aperture, and neutral-density filter) to read near 5,000 ohms under operating conditions. All apparatus terminal boxes and other metal parts were grounded to building ground.

During the experiment, the room-space containing the apparatus received some unavoidable stray light from the lightbox (if turned on) or the illuminated targets. The room was provided with A 4-watt

black felt hung on walls and draped in strategic places.

red lamp mounted below the overhanging table-edge (Figure 6) provided light by which data were recorded and various meters read during the experiment.

J. M. Williams

Luminance and the Hess Effect

97

c. THE HESS-APPARATUS LIGHTBOX. This measured 54 cm high by 75 cm wide by 34 cm deep, depth being along the S's optic axis (see Figures 6 - 8). The

floor of the box was elevated to form a false bottom about 9 cm above the surface of the table (see Figure 7). This false

bottom allowed room for a groove cut in the lightbox sheet metal so that the box might be fit snugly onto the rail described above. The false bottom also contained electrical connections. A quiet, light-baffled fan with 1 m3/min capacity and a light-baffled vent were at opposite ends of the upper side walls of the lightbox to circulate cooling air. A voltmeter and 30 amp-

ere autotransformer mounted below the table were used to set voltage (in parallel) across the six lamp filaments; normal operating level was 124.0 VAC (rms). The inside of the lightbox was lined with white General Electric Textolite tile. Background 1uminance supplied by the

lightbox with 6 General Electric Soft-White 15-watt lamps installed measured about 25O mL; the tile gave a slightly reddish but nicely diffuse white about the same as that of the red spot of an Ilford (SEI) photometer. The wiring and cooling design would permit

operation with 200-watt lamps for an estimated 7500 mL background. The supported-filament 15-watt lamps used suggested little problem with any residual 120-Hz flicker caused by the 60-HZ line frequency (see Riggs, 1965, note 7; Ives & Kingsbury, 1914; Bouma, l94l). This conjecture was confirmed to the extent that no beats could be seen when the lightbox was examined using a stroboscope flash-

J. M. Williams

Luminance and the Hess Effect

98

ing near 60 Hz or near 120 Hz.

Other possible effects such as

emission polarization by the lamp filaments (Grum & Costa, l974) or residual specular reflection by the lightbox interior (Brandenberg & Neu, 1966) were not considered significant for the present applications. The approach used would seem applicable to the design of any lightbox, so this approach is discussed here in some detail; design considerations adaptable from room-illumination computations may be found in Kaufman and Christensen (1972, ch. 9). The following procedure is suggested in designing a lightbox: [1] The size of field needed must be determined: the present apparatus, this was 15 cm high by 45 cm wide. [2] The interior surfaces must be of material which is diffusely-reflecting. [3] The lamp filaments should be arranged fairly evenly spaced around the aperture (or optic axis) and must be out of sight of S. [4] The smaller the surface area of the box occupied by the viewing aperture, and the squarer the shape of the box, the more closely the visible reflecting surface will approximate that of an integrating sphere and hence of an ideal Lambertian reflector (see Wyszecki & Stiles, 1967, Figure 4.4). So, a large For

box with a small aperture should be used, if possible. The preceding four preliminary considerations allow accurate estimation of the needed height and width of a lightbox (in the present case, as shown in Figures 7 and 8). The remaining

dimension, depth of the box, is the most critical because it most

J. M. Williams

Luminance and the Hess Effect

99

strongly affects the intensity and evenness of illumination of the visible reflecting surface (VRS). [5] To estimate the depth of the box needed, it first should be recalled that the Weber fraction for intensity discrimination of the human eye will be .01 or greater for any but the most abrupt luminance gradients. servative figure. The value .01 thus is a con-

For simplicity, and as a further safety

factor, the calculations below will not consider the inevitable smoothing effect on luminance gradients of the sides of the lightbox lying parallel to the optic axis. then may be made as follows: First, apply the inverse-square law (see Figure 9) to estimate the distribution of i1luminance E on the VRS due to direct illumination by any arbitrary pair of lamps not too far from the optic axis. For example, the pair B2 and B6 yield the The depth-estimation

following result: For lamp B2, we have E2 = I2 (k/r22) . (42)

Similarly, for lamp B6 we have E6 = I6 (k/r62) . (46)

The constants k are arbitrary and assumed not to depend on location of the lamps vis-a-vis the VRS. Values I2 and I6 are

proportional to lamp intensities, and are equal for identical lamps. Distances r2 and r6 are to some fixed, arbitrary point P

J. M. Williams

Luminance and the Hess Effect

100

on the VRS. Now, from relations (42) and (43), the sum E of illuminances at the point P will be, for I = I2 = I6, E = E2+6 = k I (r2-2 + r6-2) . (44)

As shown in Figure 9, from equation (44), at y = 0, we must have r2 = r6 = r; this, for points P (as will be assumed always) in a sagittal plane of the box passing through lamp-filaments of B2 and B6. Thus, for y = 0, E2+6 = 2 k I r-2 . Second, from Figure 9 it can be seen that r2 and r6 may be expressed as functions of x and y: r22 = x2 + (9 - y + 6)2 ; and r62 = x2 + (9 + y + 6)2 , (47) Namely, (46) (45)

the constant 9 (in cm) being chosen as the maximum value of y at which a point on the VRS can be seen by the observer (Figure 6). The 6 in equations (46) and (47) represents the distance from the maximum visible y-value to a point on the VRS and also on a line parallel to the optic axis and passing through the lamp filaments. Substituting results (46) and (47) into equation (44), we obtain

J. M. Williams

Luminance and the Hess Effect

101

E2+6

k I {[x2 + (15 - y)2]-1 + [x2 + (15 + y)2]-1} .

(48)

By inspection of equation (48) for small values of x, E2+6 attains a local minimum at y = 0 as y approaches infinity. For small

x, then, when undesirable gradients would be expected to be worst, E2+6 will be expected to attain its greatest visible value directly opposite one of the lamps (or nearby), namely at about y = +9 or y = -9 cm. Third, from the computations above, a ratio of minimum to maximum visible value of E2+6 (in the vertical direction) can be composed which will express the maximum VRS illuminance gradient due to lamps B2 and B6 shown in Figure 9. This ratio, for small

values of x, may be formed using equation (48) with y = 0 in the numerator and y = 9 in the denominator. This ratio, which

equals 1 minus the expected approximate maximum Weber ratio, is: Emin/Emax = (x2 + 152)-1 + (x2 + 152)-1 (x2 + 62)-1 + (x2 + 242)-1 . (49)

As x increases, corresponding to increasing depth of the lightbox, approximation (49) becomes increasingly poor as the maximum E located near y = 9 shifts toward y = 0. A refinement of formula (49) may be made by including effects of reflected light from the front interior surface of the lightbox (i.e., the surface in which the viewing aperture is cut). For this refinement, and for standard-size bulbs, assume the filaments of B2 and B6 to be located 9 cm from the interior front

J. M. Williams

Luminance and the Hess Effect

102

surface.

If the lining be assumed to have reflectance of about

.8 and to subtend one steradian solid angle at the filament, equation (49) may be rewritten with additional terms expressing the directly-declining illumination-contribution of the front as distance of the front from the VRS be increased: the box front a constant 9 cm from the filaments, Emin/Emax = (x2 + 152)-1 + (x2 + 152)-1 + K (.8)(4)-1 (x + 9)-1 . (x2 + 62)-1 + (x2 + 242)-1 + K (.8)(4)-1 (x + 9)-1 (50) Thus, keeping

In expression (50), the parameter K has been included to weight changes in viewing-aperture size and interior front-surface reflectance: no front. The values entered in Table I were computed from formula (50) above. In that Table, ratios above 1.00 result from Small values of K represent a lightbox with little or

the shift in maximum mentioned above and should be considered artifacts of the approximation, in view of the safety factors mentioned above. The K = 0 column in Table I corresponds to Values

values which would be obtained from formula (49) above.

of K = 5 or K = l0 probably represent the situation with all factors considered, but no analytic proof of this has been attempted. Inspection of Table I suggests that any distance x above about 23 cm would yield no discriminable gradient of illumination on the VRS caused by added light from a typical lamp-pair B2 and B6.

J. M. Williams

Luminance and the Hess Effect

103

The even spacing of the lamps (Figure 8) suggests that this result for B2 and B6 would apply for any lamp-pair in the present apparatus. Furthermore, the frosted surfaces of the bulbs actually

used, the additional gradient-smoothing contributions of the four lamps not entering the calculations above, and the diffuse light from the four sides of the lightbox interior which lay parallel to the optic axis all further improve the design safety factor for the depth estimated by the approach above. x chosen in the present design was 25 cm. [6] The result of the preceding calculations and the comments on Lambertian surfaces of Wyszecki and Stiles (1967, pp. 386 ff.) alluded to above both suggest that a good rule of thumb of lightbox design would be that the distance x of lamp filament to visible reflecting surface always exceed the average spacing of the lamps around the optic axis. d. CALIBRATIONS. (1) PHOTOMETRY. The values assigned to E0 (3.56 log td) for the targets and for the lightbox were established in a series of measurements taken over a period of weeks. For each measurement, after Actual distance

15 minutes of warmup, the lightbox voltage was set to 124 .05 VAC (rms ) as read from a panel meter permanently mounted in the apparatus. The target power supplies, warmed up slightly below

normal operating voltage, then were offset at least l volt DC from their known approximate operating levels. Using the power

J. M. Williams

Luminance and the Hess Effect

104

supply control knobs, the experimenter (viz., the present author) then matched the targets to the (lightbox) background in luminance, repeatedly adjusting the match by touch as viewed repeatedly through the artificial pupil. A bracketing procedure was used, and the

target-target match was attended to more closely than the target-background matches (see Brown, 1952). An apparent color-tem-

perature difference existed between the targets and background, the targets appearing bluer (too blue for the Ilford red spot and far too red for the Ilford blue spot) than the background. The back-

ground was about the same color as the Ilford red spot, which is assumed to have been at about 2350 K. At match, the background

was measured first with the Ilford photometer (SEI exposure meter) and then with a MacBeth Illuminometer. Both measurements were

made from the TMM fieldstop position and typically agreed within 5 percent (in mL). Also at match, target power supply voltages

were recorded from the power supply panel meters, and a carefully zeroed ohmmeter (above) on 1 K ohm scale was used to measure photoresistor values for upper and lower target and for the lightbox. Early in the study, as read from the power-supply panel meters, the target lamp filaments for an older pair of lamps were operated at about 28 VDC. One of these burned out after

about n = 12 data under each condition had been gathered, and the newly-installed lamps were found to operate at about 24 VDC. These target lamps were 40-watt, #7079 Grimes reflectorized aircraft navigation lamps. The correlated color temperature of

the targets is estimated to have been 2500 K (not much different

J. M. Williams

Luminance and the Hess Effect

105

for either set of lamps), allowing the same troland-value units to be used throughout the photopic and scotopic range with error probably less than .1 log unit (see Wyszecki & Stiles, 1967, Table 2.11). During sessions, the lamp source voltages were set using calibrated photoresistor values described above; the voltages were monitored during trials, and the photoresistors routinely were reread at least once about halfway through each session. Read-

justments of target-lamp source voltage exceeding .l VDC rarely were necessary. It was determined that against an equal-lum-

inance background, a .l VDC source voltage change in one target power supply would cause a jnd change or more, but would produce a change in luminance far less than .05 log unit. All photopic

troland values for the Hess data reported below are believed correct within .05 log unit, and, for the JMW data, usually correct within .02 log unit. The Ilford photometer was used to set the Pulfrich and tachistoscope E0 to 3.48 and 2.48, respectively, log trolands photopic. The blue spot was used for the tachistoscope. These

values are believed accurate within .02 log unit for the Pulfrich apparatus (calibrated, set, and rechecked occasionally) and within .1 log unit (as independently reset each session) for the tachistoscope. For the later reaction-time trials with JMW as obser-

ver, the tachistoscope E0 was set at 2.48 log trolands, as above, but the Pulfrich-apparatus E0 was set to 3.48 log trolands.

J. M. Williams

Luminance and the Hess Effect

106

(2) OPTICAL FILTERS. The filters for the target-illumination channels were Fish-Shurman solid glass neutrals ranging from .ll to 2.91 log units neutral density (ND). These were calibrated to within

.01 ND by means of Wratten neutral-tint standard filters using a Martens polarizing photometer. A few of the obtained densities

were verified using a Beckman spectrophotometer versus air or versus the same set of standard Wrattens just mentioned. (3) HESS TARGET RELATIVE DISPLACEMENT. A front-surface mirror was mounted before the S's eyepiece filter box (just before the artificial pupil) so a cathetometer could be used to view the targets at three different target lateral displacements each within 4 cm of the center of the TMM fieldstop. The targets were vertically aligned at each

of these three locations, then, at each location, the target displacement motor-shaft was rotated manua11y in two-turn increments, always in the same direction. At each increment, the

relative lateral displacement of the two targets was read by means of the cathetometer as well as by the experimenter's readout scale (described above). At each location, increments

were repeated until target relative displacement reached 6.1 mm, the thickness of one target rod. The result was an estimated readout-scale magnification factor of 5.1, to be used in all computations converting scale readings to true target displacement. The precision of this

J. M. Williams

Luminance and the Hess Effect

107

calibration greatly exceeded its accuracy because of unknown effects (presumably constant) of the motion of the apparatus at any given speed during actual experimental trials. In any case, the precision

of the 5.1 was about l0-4, in terms of range of sample values. But the expected effect of magnification-factor inaccuracy in any case would be small: In computing t from equation (1) above,

the formula is t = s/v, the displacement as read from the scale being used to compute s. If the true magnification factor

m be considered to differ from the estimated one by some small error e, with e probably .1 or less, we have

s/v

(s/m)/v

(s/(m' + e))/v,

(51)

m' being the "5.1" value estimated from the data.

The expected

magnitude of the error in t resulting from any e different from zero thus will be, as a fraction of t,

d(t)/t

= = =

(t/e) (de/t) [(-s/v)(m' + e)-2] (de/t) (de)/(m' + e)2 (.1)/(5.1)2

(52)

(53)

= about 0.4%, ignoring the sign of e.

J. M. Williams

Luminance and the Hess Effect

108

(4) HESS TARGET SPEED. The original calibration was done using a microswitch mounted on the lower target platform and two c-clamps clamped to the angle-iron frame of the TMM. As the platform moved back-and-

-forth, the microswitch was depressed by the c-clamps, thus signalling the timer to start or stop as appropriate. Speed settings

read from the turn-counting dial ranged from "l5" to "55" in increments of 5. The digital timer, controlled by the microswitch,

recorded times of each of 20 traverses of the platform at each speed setting: 10 right-to-left and 10 left-to-right. Traver-

ses were timed over the same (approximate) 23-cm traverse as that over which the targets would be visible to the observer during the experiment. All times were averaged at each dial setting; then,

a linear regression yielded the equation estimated speed = .4ll628 (dial setting) - 4.851547, (54)

in cm/s, with coefficient of determination .993, for the averaged speed in both directions. As vernier alignments were made only for motion from left to right, the relevant data were used to compute the regression equation estimated speed = .4l2349 (dial setting) - 4.865529, (55)

in cm/s left-to-right, with coefficient of determination .992. Equation (55) was checked occasionally during the study, using

J. M. Williams

Luminance and the Hess Effect

109

a stopwatch; obtained speeds were not found to vary more than .1 cm/s from the calibrated (regression) value, an error within the expected error of the stopwatch as used.

3. PROCEDURE. A dental-wax mouthbite was used. Targets were viewed with

the right eye only; except for the Pulfrich experiment, in which both eyes were used and the left eye always presented the target at the lower troland value El. a. REACTION TIME. For each session, S was dark-adapted 30 minutes or more, after which the 12 stimulus conditions were presented in random order. Each condition consisted of 2 to 4 practise trials followed by 12 datum trials. Trials were spaced about 10 seconds apart and conNo-response trials simply were re-

ditions 1 to 5 minutes apart.

peated; trials for which S signalled a rejection were repeated after one additional practise trial. Only reaction times bet-

ween .100 and 1.0 seconds were considered genuine, others being considered false alarms. Sessions were repeated daily for each

S until the obtained mean reaction times (n = 12) became stable enough to be considered repeated measures over four consecutive sessions. A sample of size 48 thus was gathered under each stim-

ulus condition. b. PULFRICH EFFECT. After 20 minutes or more dark-adaptation, a few

J. M. Williams

Luminance and the Hess Effect

110

depth-settings were made with the targets at the lowest level of troland value to be used in that session. For these depth

settings, the upper target was left at rest and laterally displaced by a few mm from the path of the lower (adjustable) target. Then eight different stimulus conditions were presented in increasing order of target troland value. For each condition, after 30

seconds to two minutes of light-adaptation, 8 depth-settings were made, 4 to the apparent near path and 4 to the apparent far path of the oscillating (upper) target. The lower target was offset

randomly by a small distance by E between trials in the same direction. (1960a). Other details of the procedure are described in Lit Fixation was on the upper tip of the lower target for

AB or on the lower left edge of the lower target when S was JMW. There were two sets of eight photopic conditions, and the procedure was repeated once so that n = 16 settings were made under each stimulus condition by each S. c. HESS EFFECT. In each session, after 20 to 30 minutes dark adaptation (often more when JMW was both E and S), six stimulus conditions were presented in order of increasing target troland value. There was a pause of at least one minute between conditions when the TMM had to be stopped to change optical filters. Each con-

dition consisted of l2 vernier settings, with a one-minute rest after the first 6 settings. The illuminance increase between

conditions never exceeded 3.1 log units, and 20 to 60 seconds of

J. M. Williams

Luminance and the Hess Effect

111

light adaptation was allowed before the first setting was made at a new level of higher target illuminance. The lower, always less

intense, target was offset to lag about 6 mm behind the upper target before the first of the l2 settings under each new condition; this was to facilitate light adaptation while familiarizing S with the location and appearance of the lower target. Offsets for sub-

sequent settings were counterbalanced lead/lag by random distances from the previous setting, these distances usually being between 2 and 5 mm. The offset distances were established by displacing the

targets at various speeds but always for an 8-second interval, thus reducing timing cues to S. After offsetting the targets before

each tria1, E returned the speed control (armature autotransforner) to a standard position convenient for S and then signal1ed S to align the targets. For each vernier setting, Ss were instruc-

ted to signal when satisfied and to make their final judgement on each trial only when the targets were moving from left to right, while Ss were sure they were fixating the fixation point, and, if possible, using only the leading edges of the moving targets. Also, Ss were encouraged to make their decision while the targets were passing below the fixation point--but this instruction was not emphasized because of visibility problems at scotopic levels. The sessions were organized in pairs. Each pair included

some 7 to 10 conditions each of which had the same lower-target troland value (El) as parameter, the upper target troland values (Eu) being in an increasing series. The conditions defined by the

same El were divided between the sessions of the pair so that the

J. M. Williams

Luminance and the Hess Effect

112

spacing between successive Eu values in each session was irregular: This was to reduce possible response biases which might develop as the Hess illusion visibly became larger with the successively-increased Eu values during any one session. One or more high-El con-

ditions were tagged on at the end of each session to make up the total of six distinct stimulus conditions. Successive increases in

target troland value were spaced not to exceed 1 log unit except in the transition to the tagged-on conditions for which the increase sometimes exceeded 2 log units. A protocol sheet was made up listing each pair of sessions. Every other day, a pair was chosen at random and its sessions run on two consecutive days whenever S-scheduling made this possible. After all pairs on the protocol had been run once for a given S, the procedure was begun again and repeated until a total of n = 48 settings had been made at each stimulus condition. There were a total

of 15 pairs of sessions on the protocol sheet for each S at the start of the study; this number later was reduced to 7, the full 15 pairs having been completed to n = 12 for LD and n = 24 for AB. AB even-

tually reached n = 48 for 2 of the reduced set of 7 pairs, having completed n = 36 for all 7. The protocol for JMW both as E and S was the same as for the others after the reduction, with a few additional stimulus conditions being explored, but JMW's daily procedure was slightly different: JMW preserved dark-adaptation of the right eye by wearing

an eye-patch and by closing the eye and cupping it with the palm of the hand as necessary. The vernier readout-scale was read,

J. M. Williams

Luminance and the Hess Effect

113

the optical filters changed, and the data recorded using the left eye only. The counterbalanced offsets of targets could not be

random, of course, but were varied in distance by leaving the target lateral displacement speed control at one speed throughout the session, and by counting-off silently 8, 6, or 4 seconds duration of displacement before each trial. These times also were varied

whimsically occasionally by a second or so, and no silent counting was done while viewing the targets as S. None of the JMW data

were analyzed until all had been gathered; this was to prevent possible bias. The only exception was for a few higher-photopic

data added to the protocol after the n = 48 sample for 7 session-pairs had been completed and graphed. To validate the data for

JMW as E and S, a hired undergraduate student worker (JV) was trained as E and ran JMW as S for 3 protocol session-pairs for total sample of n = 24. JMW was blind as to which session was being run on any

day of the validation but knew which 3 pairs had been selected. d. STIMULUS INTENSITY RANGE EXPLORED. For reaction time, the stimuli ranged from -1.99 log td (near absolute threshold for continuous viewing; about .5 log unit below absolute threshold when flashed) to 2.48 log td. For the Pulfrich effect, El was either .44 log td (with Eu ranging up to 2.43 log td) or 1.49 log td (with Eu ranging up to 3.48 log td. For the Hess affect, El ranged from -2.30 log td

(probably within .5 log unit of parafoveal absolute threshold for the moving targets) to 2.48 log td; Eu ranged from -2.19 log td

J. M. Williams

Luminance and the Hess Effect

114

to 3.56 log td. The specific combinations of stimulus troland values used during each of the experimental sessions have been tabulated with the Results reported below.

J. M. Williams

Luminance and the Hess Effect

115

RESULTS 1. REACTION TIME. Means and standard deviations of obtained reaction time (RT) for each observer under each of the 12 levels of target troland value studied will be found in Tables II through IV below. Each

entry tabulated represents 12 reaction times obtained for the 12 trials of the session and under the stimulus condition specified. These tables have only 11 rows because no observer was able to detect the flashed targets when they were at the lowest level of -1.99 log trolands. The column-heading numerals for these tables

give the chronological order in which the data were gathered. Mean RTs for all 4 sessions for each observer are tabulated in Table V and are graphed in Figure 10 below. The datum trains in

Figure 10 below show the typical decline in RT with increasing intensity that has been discussed above and demonstrated in Appendix I below. Close inspection of the data in Figure 10 suggests a scotopic-photopic breakpoint at about -.5 log td for LD and JMW, but not for AB. JMW's data show greater variability in the figure (as JMW's false-alarm rate

well as in Table V) than do AB's or LD's.

(response to no-target) was much higher than AB's or LD's, based on unrecorded observations during the study.

J. M. Williams

Luminance and the Hess Effect

116

To compare RT with the Hess effect as they measure intensity-dependent latency, the means of Table V were themselves averaged over observers and may be found in the RT-datum columns of Table VI. These means then were used pairwise to interpolate

(or extrapolate, as appropriate) linearly to obtain expected RTs at the levels of log Eu at which Hess-effect data had been gathered during the study for log El = 0.27 and log El = 1.19. The

set of RT data and expected RTs then was placed in order of log E (or log Eu); for the resulting sequence, a running average RT then was computed at each level of log E for a span of 3 numbers both in increasing and decreasing order. Finally, the two sets of

running-average RTs were averaged at each level of log E (or log Eu), finally yielding an estimated RT as tabulated in Table VI. Table VII then was composed to compare RT latency-differences to Hess-effect latency-differences; this was done simply by taking differences between estimated RTs at the various pairs of target troland values at which each Hess-effect datum was obtained. Cursory inspection of Table VII shows that t predicted by RT increases more rapidly than the t of the Hess effect as log Eu is increased. This gradual departure of the RT prediction from

the Hess t is noticeable in Table VII at both levels of log El. The comparison is graphed in Figure 32 below. 2. PULFRICH EFFECT. Means and sample standard deviations for the near and far Pulfrich-effect settings are given in Tables VIII and IX for AB

J. M. Williams

Luminance and the Hess Effect

117

and in Tables X and XI for JMW. gathered for LD.

No Pulfrich-effect data were

The numerals at the heads of the columns of

the tables just identified give the order in which the data were gathered, each of the two sessions being repeated once. The

entries of these columns were averaged and the Pulfrich-effect geometric t then was computed for each near (N) and far (F) average setting using the formulas tN tF = = s/v s/v = (ipd/v)[CN/(d - CN)] , = (ipd/v)[CF/(d + CF)] , (56) (57)

in which ipd represents the observer's interpupillary distance (5.8 cm for AB; 6.7 cm for JMW), CN is the near setting in centimeters displacement from the true target path distance of 100 cm, d = 100 cm, CF is the far setting corresponding to CN just defined, and v is the target speed of lateral motion in cm/s. The geometry of formulas (56) and (57) has been treated in detail by Lit (1949). The means of the obtained tN and tF, under

each stimulus condition for each observer are given in Table XII. As indicated in the note below Table XII, observers often reported no Pulfrich illusion when the target images in the two eyes were nearly equal in retinal illuminance (troland value). NOTE: Beginning here, pagination is adjusted to accommodate a new paragraph following the end of the original p. 118. The Dissertation ms. pagination is reestablished on p. 120. The data of Table XII are graphed in Figure 12 and show a tendency of JMW's t to rise more rapidly as a function of log Eu than AB's t at the lower values of log El. This difference probably is related to fixation and is discussed below.

J. M. Williams

Luminance and the Hess Effect

118

The Pulfrich-effect t computed for AB is compared with the classical data of Lit (1949) in Figure ll; in this figure, AB's t seems to show a slight tendency to level off sooner for log El = .44 than does the t for the Lit (1949) data at log El = .98. A similar slight difference in the other direc-

tion occurs for AB's log El = 1.49 datum-train vis-a-vis the Lit log El = 1.38 train. All things considered (including the aver-

aging of the Lit data and the black-on-white Lit targets versus the white-on-black AB targets), the AB Pulfrich-effect data seem not to differ very much from those of Lit (1949). Figure 13 gives

a theoretical fit of the present proposed theory (formula (35) above) to the extensive Lit (1949) Pulfrich-effect data; the same fit would appear to hold for the AB data of Figure ll with no greater error. 3. HESS EFFECT. The data may be found for each observer in Tables XIII through XLVI. In all these tables, the numeral at each column head gives Each entry tabulated the order, left to right, in which settings for the t means and standard deviations tabulated were made. stimulus conditions specified. Each sample standard deviation (s. d.) given is a mean of the two (n = 6 each) sample standard deviations before and after the one-minute break between the 6th and 7th setting made under each condition. break. The given s. d.'s therefore do not represent any variance attributable to a drift in the mean setting because of the [This is the new paragraph] When data tabulated were gathered for more than 4 sessions, the 5th-session data are in column l in the row immediately below the represents 12 vernier setttings during the session and under the

J. M. Williams

Luminance and the Hess Effect

119

row of the (chronologically) preceding 4 sessions. Because of the procedure used to present stimuli, it happens to be convenient to record the Hess-effect data gathered for logE = .ll separately from all other data; thus, the first table in the set of tables for a given observer contains results at a single value of logE = .11; every other of Tables XIII through XLVI gives various increasing values of logE for single values (supraordinate column headings) of log El. Table XLVI collects a few odd settings made by each

observer at log El = 2.09. The Hess-effect data for LD may be found in Tables XIII through XX and in Table XLVI. Many data tabulated in the two

leftmost columns of each of these tables have been graphed in Figure 15 or have been averaged with AB's data in Figures 19, 21, 22, 23; many of these data also are averaged with both AB's and JMW's in Figures 27 and 28. The LD data in Figure 15 show the trains typical of those found for Hess or Pulfrich effects: t remains near 0 when E1

is about equal to Eu; as Eu is increased away from (the parameter value of) El, t increases. For LD, as for the other obser-

vers of the present study, a scotopic-photopic breakpoint (see Figures 15, 18, and 27) is evident near -1 log troland; below this breakpoint, t increases with Eu at a rapidly decreasing rate, and above this breakpoint, a steep and almost log-linear rise (see Figure 27) in t commences and continues until the lower target is lost in the increasing glare of the upper target, and settings

J. M. Williams

Luminance and the Hess Effect

120

no longer can be made.

When the lower target itself is increased

parametrically to slightly above the breakpoint (somewhere near -.4 or -1 log troland), the above-breakpoint increase in t with increasing log Eu no longer is endlessly linear, but suddenly seems to halt at some remarkably sharply-defined maximum value, as can be seen in the log El = -.69 and -.43 trains in Figure 15. When

both targets are above the breakpoint in troland value, the data seem to show an evenly-ordered gradually-decreasing rate of rise of t with increase in log Eu (see Figures 15, 16, and 25, and compare Figure 13). The three observers of the present study

show small individual differences in the photopic range, but all datum-trains (e. g., in Figures 15, 18, and 25) almost can be superimposed with deviations in average t very small compared with the shapes of the trains defining the functional dependence of t on stimulus intensity. These relatively small deviations

justify averaging the data of all three observers, the underlying assumption being that the same (normal) photopic human retina is being studied in all three observers of the present study. One

small photopic departure from the others in LD's data can be found: The log El = 2.48 rise in t seems steeper (Figure 15) than for JMW or AB (compare Figures 18 and 25). The Hess-effect data for AB are in Tables XXI through XXXVI and in Table XLVI. Many of these data have been graphed

in Figures 16, 17, and 18, and have been averaged with data of LD in Figures 19, 21, 22, and 23; these AB data also have been averaged both with data of LD and of JMW in Figures 27 and 28.

J. M. Williams

Luminance and the Hess Effect

121

The extensive data for AB show the same general features as those for LD just described. It might be noted here in passing that

in the scotopic range, Figures 15 and 18, especially for parameter value log El = -2.19, seem to show a decrease in the rate of rise of t with increasing log Eu; this decrease seems to become especially pronounced by about t = 80 or 90 ms. For the data

plotted in this range, the tables reveal occasional session means exceeding t = 95 or 100 ms for LD and AB (Figures 15 and 18). By contrast, the JMW data of Figure 24 show almost a linear rise to t = 130 ms at least. The levelling-off in AB's and LD's

scotopic data is believed to be an artifact of incomplete dark adaptation and guessing strategy. In Figure 17, the abscissa variable of Figure 16 has been used as parameter, an exchange made useful by AB's extensive n = 24 data. With the data replotted as in Figure 17, it is

apparent that for small differences logE, t remains about constant in the scotopic range, rises slightly between -1 and -.5 log td, then decreases rather sharply toward 0 in the photopic range. Also evident is the ragged but systematic tendency of the This

small triangles to lie above the small open circles plotted: demonstrates that a target troland-value difference of .11 log unit consistently yields a t greater than 0 for the mean of a sample of size 24. t

The large-symbol data in Figure 17 seem

to show a peak at log El between -1 and 0, the apparent peak being at lower values of log El for greater values of logE.

J. M. Williams

Luminance and the Hess Effect

122

Figure 17 also demonstrates that there is a region in the log El-t upper half-plane in which ts seem not to be obtained. Although some data difficult to parametrize meaningfully have not been replotted from Figure 16, this border does seem to be defined by a curve with negative slope and approaching t = 0 somewhere near log El = 3. This may represent saturation of the scotopic

mechanism (Wyszecki & Stiles, 1967, pp. 580-81) for this observer, who consistently showed a negative constant error of several milliseconds in vernier matching at photopic levels during the first n = 24 settings of the experiment (Figure 17. small circles to right). AB's extensive Hess-effect data, like those of JMW be1ow, allowed use of the sample standard deviation to estimate vernier acuity for the moving Hess-effect targets. These standard-devia-

tion data have been averaged over sessions for AB and JMW and are collected in Tables XLVII, XLVIII, and XLIX for target differences in log troland value of 0.00, .ll, and .91 respectively. The

same data are graphed for AB in Figure 29 and for JMW in Figure 30. These figures show that AB's vernier acuity for the laterally moving Hess-effect targets of the present study, as measured by the sample standard deviation of settings, decreases from about 8 ms at log El = -2 to about 2 ms for log El above about 0. Similar

data for JMW show a decrease from about 15 ms at log El = -2 to near 3 ms above log El = 0. Because of the geometric nature of

the t used here, these acuities represent localization thresholds of 35 to 50 seconds of arc in the photopic range.

J. M. Williams

Luminance and the Hess Effect

123

The Hess-effect data for JMW are given in Tables XXXVII through XLVI below. Most of these data have been graphed in

Figures 24, 25, 26 and have been averaged with LD's and AB's data in Figures 27 and 28. The major trends in the datum-trains

graphed have been pointed out above, but it might be mentioned again that the JMW data do not seem to show the high-level deceleration of t with increasing log Eu for scotopic values of log El. The validation data of JMW gathered by experimenter JV are given in Figure 26 and are interpreted to validate the data of Figures 24 and 25. For log El = -1.40 in Figure 26, the datum-train

seems to display a breakpoint between log Eu = -.5 and -1 which does not seem to appear in the previously-gathered log El = -1.40 data (Figure 24). However, a distinct breakpoint is evident in

Figure 25 for the log El = -2.19 trains; the slightly different shape of the validation train vis-a-vis the previous log El = -1.40 data thus is attributable to random factors justifying the averaging of the previous and the validation data in Figures 27 and 28.

4. LEAST-SQUARES APPROXIMATE FITS. Hess-effect mean obtained t-values for each observer were averaged at log El = -2.19, -1.40, 0.27, 1.19, and 2.48 and are given in Tables L and LI. Inspection of the graphs of the

low-level data (Figure 27) suggests that they be broken into scotopic-photopic subsets as coded in Table L: The [101] - [107]

J. M. Williams

Luminance and the Hess Effect

124

data in Table L are assumed to represent t for scotopic response to both targets. The [201] - [211] data presumably were gathered

for scotopic response to the lower target and photopic response to the upper. JMW's fixation is assumed to have been consistent

enough to have prevented gross light-adaptation of the retina in the regions traversed by the image of the lower Hess-effect target. It might be mentioned that for the [101] - [107] and [201] - [211] data of Table L, when the upper target was relatively intense, the lower target, when visible, was phenomenally the same blue color as that of the blue-arc entoptic rays (Moreland, 1969) which can be created by imaging a small luminous spot on the parafoveal region of the retina. This blue contrast color of the lower target was

replaced with a reddish or brownish contrast color for trials in which log El was high enough to enter the low photopic range. By inspection of Figure 27, the [200]-series data of Table L seen to fall along (log) straight lines. Mathematica11y,

the proposed theory cannot account for precisely straight lines, which would be expressible entirely in the first two terms of a Taylor series expansion of the theoretical function: The Taylor

expansion of formula (35) must be infinite and so, for the [200]-ser ies data, cannot be expected to converge nicely for any well-defined, unique parameter value of the system time-constant a. Thus, it was decided to use all the log El = -l.40 data as a unit for the fit in this region; this, in spite of a slight slope discontinuity evident (Figure 27) in the train separating the [105] [lO7] data from the [206] - [211] data. Even so, the log El =

J. M. Williams

Luminance and the Hess Effect

125

-1.40 unit still does not seem to depart much from a straight line. For the log El = -2.19 train (Figure 27), it was decided

to fit only the lower segment (data [101] - [104] of Table L). As can be seen from Figure 28, the upper segment of the log El = -2.19 train is well fit by a straight line which, incidentally, locates the presumptive scotopic-photopic breakpoint in the train at about -.9 log td. Data in Figure 27 and in Table L were chosen

to exclude any trains (of those studied in the entire experiment) the log El parameter of which might itself fall near the scotopic-photopic breakpoint which, as described above, seems to occur under the present conditions somewhere between -.5 and -l log troland. The procedure for determining normal equations for a least-squares fit of the various theories of interest has been described in Appendix II below. Upon obtaining these equations as in Appen-

dix II, it was found that they usually were less tractable than the sums of squares of deviations (SS) themselves. So, the least-

-squares equations were not used but rather the minimum SS was approximated (usually to three significant digits in the parameters) empirically by iteration and successive inspection of results on a hand-held programmable calculator. For the -1.40

datum-trains, the convergence was quite poorly defined in the case of the present theory proposed: Values of the time constant a

ranging between .77 and 1.3 could be fit practically with no difference in SS. Thus, a = 1.0 was chosen arbitrarily and is tabThe problem with

ulated with the other results in Table LII.

Alpern's power function with n free to fit was different but also

J. M. Williams

Luminance and the Hess Effect

126

was solved arbitrarily:

For the log El = -1.40 train, any increase

in n (with concomitant increase in K and T0) reduces the SS minimum obtainable for that value of n; thus, after a rather unrealistic n = 9.7 was reached, the algorithm was terminated and the result given in Table LII. The results for the approximate least-squares fit for each theory are given in Table LII: The names associated with the var-

ious formulas in the Selected Theories section above are listed in the leftmost column, the respective parameter-labels are to the right of the names, and the results for each datum-train fitted are in the five columns to the right of the parameter-labels. Alpern's formula was fit both for n = 4 and for n (continuously) free to fit, because of that formula's theoretical significance. In the tabulations for any given theory, the fitted parameter values for each datum-train are given, then, below, in parentheses, is a descriptive statistic which is being called the deviation ("Dev.") of the fit and is defined as deviation = [SS/(n - 1)]1/2 , (58)

in which SS is the sum of squares of deviations for the given parameter value and n is the number of data in the train being fit (see Tables L and LI). The unit of measurement of this

deviation is the millisecond, the same unit as for t. A quantitative comparison of the theories chosen was made in terms of the mean of the deviations for the respective theories as given ("Dev.") in Table LII. Beginning with Exner's, these

J. M. Williams

Luminance and the Hess Effect

127

mean deviations, in the order of Table LII, were: 3O,11, 4,O9, 3.77, 2.47, 1.85, and 2.75.

9.41, 3.60,

The power functions of

Charpentier (3.60) and Alpern (3.77, n fixed at 4) by this measure thus score quite closely; they are surpassed by the present proposed theory (2.75). but not if Alpern's parameter n is allowed to vary continuous1y (2.47 or less). Enright's difference-equation

fit (1.85) seems remarkably and uniformly (Table LII) close, although its interpretation in terms of possible underlying mechanisms is obscure. Based on the mean deviations just described, Figure 28 below was used to sketch a comparison of the present proposed theory (formula (35) above) with Enright's. Alpern's theoretical fit

to Pulfrich effect data may be found in Alpern (1968) and in Prestrude (1971). The curves in Figure 28 were drawn by substituting

the appropriate parameters and stimulus values into the formulas and plotting a few resulting points. The points plotted then

were fit with a flexible template and a smooth curve inked along the template. The data in Figure 28 are replotted from Figure 27.

For the fits sketched, it was considered desirable to group the appropriate approximate least-squares estimates and use the resulting averaged parameter estimates to generate the curves plotted. The assumption was that the usefulness of a theoretical time-constant should be decided by whether it can be considered constant. The averages used, taken from Table LII, are explained in the caption to Figure 28.

J. M. Williams

Luminance and the Hess Effect

128

It should be noticed in Figure 28 that the fit for a = .154 seems reasonably close both for the lower scotopic segment of the log El = -2.19 datum-train and for the data in the photopic range. Figures 13, 22, and 23 give similar results for fits made by eye in which a = .16 was chosen. Although the parameter k of the

present theory can be interpreted as a persistence (or relaxation or decay) weighting factor, k also happens to be equivalent to a free parameter scaling the base of the logarithmic input to the theoretical first-order system of the proposed theory. Thus,

the value of k has less of a definite meaning than that of the time-constant a in describing the mechanisms possibly underlying the Hess and other latency dependent effects. further-analysis of k has been made. Therefore, no

5. LATENCY COMPARISONS FOR INDIVIDUAL SUBJECTS. For reaction time, the trains of data in Figure 10 can be seen to be reasonably parallel for the individual observers. The slightly rapider downward trend which seems to occur in JMW's RT at the higher intensities, if real, can be attributed to the higher variability probably underlying JMW's higher false-alarm rate as noted above. The averaged data given in Table VII and

Figure 32 therefore were considered representative of the individual cases. so no comparisons (see above) of RT with other latencies were made on individual bases. For observer AB, the Pulfrich-effect and Hess-effect data are graphed together in Figure 20. There would seem to be a

J. M. Williams

Luminance and the Hess Effect

129

shallower slope for the AB Pulfrich effect than for the Hess effect in Figure 20. AB's Pulfrich-effect performance is comThe small parametric dif-

pared with that of JMW in Figure 12.

ference of log El = .44 for AB versus .54 for JMW in Figure 12 seems to make a considerable difference in the datum-trains at these levels; the higher-level data for log El = 1.49 for AB and 1.59 for JMW seem only slightly different. Both AB and JMW,

as has been pointed out above, tend to display essentially equivalent performance on the Hess effect. Figure 31 gives data for JMW comparable to the Figure 20 data for AB. For ts less than about 20 ms, the Pu1frich-ef-

fect and Hess-effect trains appear only to reflect the differences in parameter value. In Figure 31 for ts above about 20 ms,

the Pulfrich and Hess ts seem to converge to run parallel and nearly along the same paths; this apparent break downward of the .27-parameter Hess-effect data in Figure 31 seems complimented by a similar break upward in the Pulfrich-effect .54-parameter train, and similarly for the Hess-effect 1.19-parameter train complimentary to the Pulfrich-effect 1.59-parameter train. It should be recalled that JMW's fixation was about .8 below the path in the visual field of the tip of the laterally-moving Pulfrich target for the data of Figure 31, and of Figure 12 as well. 6. MISCELLANEOUS OBSERVATIONS. For Hess-effect sessions, at the highest levels of target

J. M. Williams

Luminance and the Hess Effect

130

illumination (log El or log Eu above 3 log trolands), JMW was able to see a brief intrusion into the visual field beginning with each target traverse about .5 second after the targets first became visible. The intrusion was as by a circular black

stopping-down (or constriction) which reduced the visual field by up to perhaps 5% of its linear extent. The stopping-down

became maximum and then seemed to recede before the targets had reached the point in their path below the fixation light. This

effect is interpreted to have been due to JMW's iris becoming visible because the small visual-angle subtense of the illuminated targets allowed a shadow of the iris's edge to be imaged on the retina with negligible blur. Using de Groot and Geb-

hard's (1952) formula, a pupil diameter of some 2.46 mm would be expected for a large-angle 250 mL target viewed through a 2.4 mm artificial pupil; thus it is not inconsistent with known population data that the phenomenal intrusion should have been that of JMW's iris. Because the intrusion distance was phenomenally

small, it is not certain that the iris diameter needs ever to have been smaller than that of the artificial pupil; therefore, no correction to the photometrically-determined target troland values which have been given in the present work has been applied. Another observation might be worth mentioning: For certain

combinations of Hess-effect target intensities, with the lower target leading the upper target, the two targets were not seen as being separate but rather as though joined by a luminous ramp (or pseudopodium) extending from the lower target to the upper. This effect

J. M. Williams

Luminance and the Hess Effect

131

was noticed by JMW as observer and experimenter and was noted to have been most obvious when the values of log Eu and log El respectively were -.05 and -1.40. The effect also was observed and

so noted at log Eu = .43 with log El at -1.40, and also at log Eu = 3.18 with log El at .27. This ramp effect resembled the appar-

ent filling-in of color sometimes observed when matching in a small bipartite field: The colors of the distinct halves often are seen

to flow together and mix uniformly although a considerable difference in dominant wavelength might be present. The ramp obser-

ved appeared to have the brightness of the lower target.

J. M. Williams

Luminance and the Hess Effect

132

DISCUSSION The most important finding of the present study with respect to the original goal is that the Hess effect yields a latency measure equivalent to the Pulfrich effect within the theoretical context developed: The time constants for the theor-

etical underlying linear systems for the Hess effect and for the Pulfrich effect both would seem to lie between .15 and .16 second under the experimental conditions investigated. This, combined

with the monocular basis of the Hess effect, further suggests that both the Hess effect and the Pulfrich effect depend upon latency differences determined mainly at the retina. A related finding is that the present study seems to show that reaction time (RT) does not predict the same intensity-dependence of visual latency that the Hess or Pulfrich effects do: One

conclusion possible from this is that RT involves a central component of latency which itself depends upon stimulus intensity. The findings of Bernhard (l940) discussed above have shown that over a two log-unit range of intensities, the dependence of RT upon intensity can be attributed to retinal processes. The im-

portance of the present study in this regard is that the present findings confirm those of Brauner and Lit (1976) in that if a sufficient1y wide range of stimulus intensities are investigated, the RT-intensity relation departs from the Pulfrich or Hess-effect latencies. But further study of this question would seem to be

J. M. Williams

Luminance and the Hess Effect

133

indicated in designs allowing direct measurement of the retinal and the other-CNS components of latency as functions of visual stimulus intensity (see Appendix I). One possible inference regarding the Hess effect consistent with the present data would be that the effect depends upon a mechanism with predominantly scotopic (perhaps specifically a rod-photoreceptoral) temporal response characteristics: Figure

28 clearly illustrates the similarity in shape of the certainly-scotopic lowest level Hess effect latencies studied in the experiment, as seen at left, to the shape of the high-mesopic or photopic branches of the latency response at right. This similarity,

with the identity (apparently) of the system time-constants for the theoretical fit, suggests the same underlying mechanism. Also, it has been noted above that when the Hess-effect data are plotted as functions of log El, the t datum-trains decline considerably as the visual system enters a range in which the rod mechanism becomes saturated (unable to respond differentially to small or brief changes in stimulation). This level of rod sat-

uration, consistent with Wyszecki and Stiles (1967, section 7.8), appears in Figure 17 to occur somewhere near 3 log trolands. It has been mentioned already that the color temperature of the illuminated Hess-effect targets allows the same scale to be used, with little error (Weaver, 1949), both for troland value in the photopic range and in the scotopic range. A photopic-scotopic breakpoint can be defined using the Hess effect either in terms of the lower target (parametric) level

J. M. Williams

Luminance and the Hess Effect

134

by which the present data usually have been plotted, or in terms of the upper target: The two breakpoints obtained (see Figures

24 and 25) may possibly be located at the same level near -.9 log td (Figure 28, line segment), and they certainly lie between -1.3 and -.5 log trolands under the present experimental conditions. It might be mentioned for comparison here, that the

classical study of Shlaer (1938) locates the resolution acuity scotopic-photopic breakpoint for gratings at about -1.5 log td as compared with the similar breakpoint for Landolt C's at about -.5 log td. If one considers the exposure duration at individual

small retinal regions for the moving targets of the present study, the energy summed in those regions would be expressible as about .011 td-s; for Shlaer (1938), assuming a summation interval of about .l second for continuous viewing of stationary targets, the breakpoint would be defined near .03 td-s for the Landolt C's and near .003 td-s for the gratings. Thus, the scotopic-photopic

breakpoint found in the present study reflects a sensitivity of the photopic mechanism in the present vernier localization task somewhere intermediate between the similar sensitivity for resolution of a grating versus resolution of a Landolt C. Figure 31, showing the apparent convergence of intensity dependent ts for the Hess and Pulfrich effects, remains unexp1ained: It would seem possible that the increasing intensity

allowed JMW to fixate more precisely for the Pulfrich data at the higher levels, thus yielding a t disproportionally larger than that which would be predicted from the trends at the lower levels.

J. M. Williams

Luminance and the Hess Effect

135

It would be interesting if the small residual t difference between Hess-effect and Pulfrich-effect data in Figure 31 could be attributed solely to a CNS function mediating stereopsis. But

latency-analysis, as discussed in the Introduction above, becomes difficult to apply validly beyond the first visual synapses of the midbrain and cortex; it is responses of these regions which would seem to be involved, at least as inputs, in central systems governing stereoscopic depth perception. In terms of a sample standard deviation, vernier localization threshold for moving targets in the Hess effect has been given as a function of intensity for AB and JMW in Figures 29 and 30. There do not seem to be any published data on moving vernier acuity in the literature. The .91 log-unit difference in target troland

values seems to have had a small effect for JMW (Figure 30) and little or none for AB (Figure 29). The obtained thresholds, ranging

down to 30 seconds of arc or so, are above those usually found for vernier acuity for stationary targets, which latter thresholds may be 4 to 10 seconds of arc under similar conditions (Berry, 1948; Berry, Riggs, & Duncan, 1950; Baker, 1949). 0n the other hand,

the present localization thresholds are considerably lower than the corresponding 120-second resolution thresholds for moving Landolt C's (Brown, 1972). AB's excellent uncorrected Keystone acuity as

well as JMW's corrected myopia both seem to be mediating typical hyperacuity performance under the present moving-vernier conditions.

J. M. Williams

Luminance and the Hess Effect

136

It seems fitting to conclude here with one final inference as to retinal function based on the Hess-effect data obtained during the present study: This would be in reference to the seeming-

ly endless linear increase in t with increasing log Eu for the JMW data in Figure 24, an increase apparently obtained whenever the lower target lies in the scotopic range and the upper target in the photopic range. As has been mentioned already in the

Results, the theoretical formula (35) above does not fit this kind of increase with a well-defined time-constant a. One pos-

sible explanation would be that for the linear train, the successive increases of the upper target troland value successively shorten the summation interval for the response of the lower target; thus, in effect, absolute latencies for the lower target are constantly being increased by successive reduction in energy input to retinal regions stimulated by the lower target image. One might guess that this interaction only would occur when the two targets were being responded to by different (viz., scotopic versus photopic) mechanisms, the response of the same mechanism being unable to shorten its own summation interval. This

inference would seem to be consistent with the approach of Matin (1968) and with the data of Gouras and Link (1966), which latter indicate that the earlier-responding mechanism can preempt the ordinary stimulus-specific output of a given ganglion cell depending on shorter latencies governed by higher intensities of effective stimulation.

J. M. Williams

Luminance and the Hess Effect


TABLE I DEPTH OPTIMIZATION FOR A LIGHTBOX WITH VISIBLE REFLECTING SURFACE 15 x 45 cm

137

DISTANCE x (cm) 10 15 20 22 24 25 28 30 50 100

APERTURE CONSTANT K 0.00 .697 .875 .964 .984 .998 1.003 1.014 1.018 1.020 1.007 0.01 .698 .876 .965 .984 .998 1.003 1.014 1.018 1.020 1.007 0.10 .708 .881 .967 .985 .998 1.003 1.013 1.017 1.018 1.006 1.00 .780 .918 .979 .991 .999 1.002 1.007 1.009 1.008 1.002 5.00 .895 .965 .992 .997 1.000 1.001 1.003 1.003 1.002 1.000 10.00 0.94 .980 .995 .998 1.000 1.000 1.001 1.002 1.001 1.000

J. M. Williams

Luminance and the Hess Effect


TABLE II REACTION-TIME DATA (MILLISECONDS) LD SESSIONS (n = 12) IN CHRONOLOGICAL ORDER SESSION

138

Log E Mean -1.51* -1.22 -.82 -.43 .04 .49 .92 1.40 1.69 2.09 2.48 525.8 364.8 373.5 345.3 323.1 287.4 299.9 276.4 302.7 257.3

1 s. d. Mean

2 s. d. Mean 631.2 151.9 33.9 36.0 22.3 21.4 39.0 19.6 23.7 32.0 19.8 373.2 428.2 411.1 367.1 321.9 307.7 303.0 263.6 269.4 282.8 48.3 76.3 69.1 57.0 24.1 32.5 28.7 28.9 24.0 31.5 538.9 412.2 375.4 345.8 298.1 310.3 271.0 283.9 274.6 279.8

3 s. d. 163.6 109.6 42.7 20.5 38.7 38.1 30.5 38.5 26.7 24.9 18.8 384.5 384.7 396.7 337.5 335.0 284.6 285.5 281.4 274.4 266.2 Mean

4 s. d.

49.6 27.7 102.7 44.0 52.4 31.8 28.3 54.1 20.8 28.4

*Note:

S could not see the flashed targets except during Session 3.

Underlined data for Session 1 were run inadvertently on the day of Session 2, and vice-versa.

J. M. Williams

Luminance and the Hess Effect


TABLE III REACTIDN-TIME DATA (MILLISECONDS) AB SESSIONS (n = 12) IN CHRONOLOGICAL ORDER SESSION

139

Log E Mean -1.51 -1.22 -.82 -.43 .04 .49 .92 1.40 1.69 2.09 2.48 515.3 441.0 372.2 325.8 302.3 273.7 256.2 255.7 248.1 240.7 250.3

1 s. d. 76.8 54.8 18.1 27.8 22.6 20.7 14.9 34.5 16.6 22.5 17.2 Mean 573.4 455.7 385.3 350.3 310.7 283.3 283.1 260.7 242.4 246.3 245.4

2 s. d. 101.0 70.1 28.9 31.3 19.2 15.1 27.9 18.6 26.9 24.3 21.6 Mean 548.8 419.9 385.1 337.8 300.4 300.6 258.8 267.3 236.6 253.3 256.8

3 s. d. 134.5 73.7 25.6 35.7 28.2 68.9 21.9 22.7 9.9 24.7 29.0 Mean 482.7 445.4 370.6 339.1 318.8 295.9 263.3 253.1 258.6 254.8 245.0

4 s. d. 89.1 89.4 7.0 17.9 24.6 27.5 19.2 17.4 21.6 17.7 21.7

J. M. Williams

Luminance and the Hess Effect


TABLE IV REACTIDN-TIME DATA (MILLISECONDS) JMW SESSIONS (n = 12) IN CHRONOLOGICAL ORDER

140

SESSION Log E Mean -1.51 -1.22 -.82 -.43 .04 .49 .92 1.40 1.69 2.09 2.48 490.0 441.0 340.8 346.8 279.3 281.0 267.1 212.0 222.5 226.9 198.8 1 s. d. 55.7 57.4 61.0 47.6 33.9 65.5 50.3 16.0 35.7 35.4 20.4 Mean 528.0 452.2 385.6 343.8 289.8 286.5 248.2 238.8 216.8 232.0 189.4 2 s. d. 48.5 82.3 58.4 71.6 52.6 74.8 35.9 53.8 30.9 38.0 22.0 Mean 553.1 432.0 367.3 367.2 273.2 245.8 283.2 216.5 226.5 200.8 208.3 3 s. d. 104.2 117.2 61.0 84.2 41.3 38.7 84.5 51.4 26.1 37.0 41.5 Mean 483.9 469.0 353.7 310.4 264.1 266.0 241.8 234.6 208.1 196.6 208.3 4 s. d. 104.7 91.9 32.8 36.5 32.6 43.8 38.3 36.6 35.1 20.8 42.5

J. M. Williams

Luminance and the Hess Effect


TABLE V COMBINED REACTIDN-TIME DATA (MILLISECONDS)

141

Log E -1.51 -1.22 -.82 -.43 .04 .49 .92 1.40 1.69 2.09 2.48

Mean 631.2* 455.6 397.4 389.2 348.9 319.5 297.5 289.9 276.3 280.3 271.5

LD

s. d. 163.6* 89.9 45.2 57.1 40.5 34.0 33.5 28.8 33.4 25.4 24.6

Mean 530.1 440.5 378.3 338.3 308.1 288.4 265.4 259.2 246.4 248.8 249.4

AB

s. d. 100.4 72.0 19.9 28.2 23.7 33.1 21.0 23.3 18.8 22.3 22.4

Mean 513.8 448.6 361.9 342.1 276.6 269.8 260.1 225.5 218.5 214.1 201.2

JMW

s. d. 78.3 87.2 53.3 60.0 40.1 55.7 52.3 39.5 32.0 32.8 31.6

* Note:

These data for n = 12 responses.

J. M. Williams

Luminance and the Hess Effect


TABLE VI AVERAGE 0F THREE OBSERVERS' REACTION TIMES ESTIMATED AT LEVELS OF ILLUMINANCE STUDIED FOR THE HESS EFFECT (MILLISECONDS) Hess-effect Log El 0.27 1.19 RT Datum 311.2 RT Est. 306.52 300.69 .49 292.6 294.66 288.15 279.41 .92 274.3 274.27 266.59 1.40 258.2 259.42 253.58 1.69 2.09 247.1 247.7 249.57 247.02 245.42 242.57 2.48 240.7 239.04 235.52 231.40 2.79 3.18 2.10 2.39 2.48 240.7 1.86 2.09 247.7 1.62 1.69 247.1 1.19 1.38 1.40 258.2 Hess Log Eu RT Log E .92 RT Datum 274.3

142

Hess Log Eu

RT Log E .04

RT Est. 270.22 265.27 260.88 256.23 251.85 249.54 247.84 246.65 245.58 242.65 239.12 235.52 231.40

.27

.54 .80

1.08

1.62

2.10 2.39

2.79 3.18

J. M. Williams

Luminance and the Hess Effect


TABLE VII INTENSITY-DEPENDENT DIFFERENCES IN REACTION TIME COMPARED WITH HESS-EFFECT ts AVERAGED FOR THE SAME THREE OBSERVERS Hess-effect Log El 0.27 1.19 RT (ms) 0.00 12.54 19.84 34.10 47.11 55.27 58.12 65.17 69.29 Log Eu 1.19 1.38 1.62 1.86 2.10 2.39 2.79 3.18 Hess Datum 0.78 3.30 9.30 14.14 17.18 20.20 21.05 24.87

143

Log Eu 0.27 0.54 0.80 1.08 1.62 2.10 2.39 2.79 3.18

Hess Datum 0.15 12.89 19.56 30.27 41.23 44.67 45.13 45.75 44.76

RT (ms) 0.00 4.39 13.42 17.43 19.69 22.62 29.75 33.87

J. M. Williams

Luminance and the Hess Effect


TABLE VIII PULFRICH-EFFECT DATA DISPLACEMENT IN CENTIMETERS AB SESSIONS (n = 4) IN CHRONOLOGICAL ORDER Log El = 0.44 1 2 Far s. d. .312 .166 .124 .147 .116 .257 .433 .718 Mean -.240* .668 1.335 1.995 1.878 2.380 2.973 3.820 s. d. .212 .108 .263 .303 .245 .023 .842 .345 Mean -.170* -.300* -.133 .305 1.292 1.927 2.350 3.562 Near s. d. .126 .189 .129 .179 .144 .177 .345 .428 Mean .255* .233* .480 1.418 1.565 2.313 2.445 2.860 Far

144

Log E Mean 0.00 .19 .43 .67 .91 1.20 1.61 1.99 .032 .187 .347 .670 .660 1.032 1.432 1.012

Near

s. d. .178 .083 .067 .083 .201 .386 .127 .275

*Note: No effect reported: No phenomenal Pulfrich effect (difference in apparent distance, left-to-right vs. right-to-left) was reported.

J. M. Williams

Luminance and the Hess Effect


TABLE IX PULFRICH-EFFECT DATA DISPLACEMENT IN CENTIMETERS AB SESSIONS (n = 4) IN CHRONOLOGICAL ORDER

145

Log El = 1.49 1 2 Far s. d. .091 .062 .149 .164 .143 .308 .084 .272 Mean .003* .678 1.260 1.728 2.113 2.830 3.305 3.930 s. d. .073 .132 .198 .238 .084 .191 .364 .323 Mean -.078* -.093* .267 .460 1.857 2.205 2.922 3.170 Near s. d. .165 .102 .177 .186 .082 .307 .140 .547 Mean .183* .668* 1.345 1.628 2.063 2.495 2.480 3.083 Far s. d. .057 .090 .238 .194 .227 .108 .358 .366

Log E Mean 0.00 .19 .43 .67 .91 1.20 1.61 1.99

Near

-.060* .370 .730 1.122 .872 1.242 2.077 2.255

*Note: No effect reported.

J. M. Williams

Luminance and the Hess Effect


TABLE X PULFRICH-EFFECT DATA DISPLACEHENT IN CENTIMETERS JMW SESSIONS (n = 4) IN CHRONOLOGICAL ORDER Log El = 0.54 1 2 Far s. d. .407 .414 .477 .444 .728 .674 1.146 2.599 Mean .960* -.792* 1.368 1.763 2.813 2.635 3.940 4.273 s. d. 1.524 .649 1.046 .673 .799 .792 .583 1.877 Mean .630* 1.11 1.855 2.452 2.740 2.992 4.187 4.967 Near s. d. .289 .577 .524 .375 .190 .436 .584 .952 Mean -.732* -.015* .685 1.135 1.725 1.403 2.633 3.380 Far

146

Log E Mean 0.00 .19 .43 .67 .91 1.20 1.61 1.99

Near

s. d. .221 .236 .381 .132 .759 .634 .738 .682

-.173* .265* .890 1.845 2.252 3.047 4.347 3.552

*Note: No effect reported.

J. M. Williams

Luminance and the Hess Effect


TABLE XI PULFRICH-EFFECT DATA DISPLACEHENT IN CENTIMETERS JMW SESSIONS (n = 4) IN CHRONOLOGICAL ORDER Log El = 1.59 1 2 Far s. d. .293 .312 .259 .242 .473 .640 .876 .823 Mean -.555* -.405* -.590* -.342 .590 2.193 1.085 2.048 s. d. .916 .390 1.282 1.797 .079 2.582 .378 .640 Mean .790* .762* 1.925 2.390 2.155 2.702 3.632 3.950 Near s. d. .178 .344 .289 .134 .647 .330 .446 .617 Mean -.432* -1.222* -.387 -.115 .443 .285 .790 1.250 Far

147

Log E Mean 0.00 .19 .43 .67 .91 1.20 1.61 1.99 .540*

Near

s. d. .203 1.433 .153 .240 .216 .671 .561 .847

1.200* 1.735* 2.247 3.025 3.207 2.990 4.340

*Note: No effect reported.

J. M. Williams

Luminance and the Hess Effect


TABLE XII COMBINED PULFRICH-EFFECT DATA t IN MILLISECONDS AB JMW Log El 1.49 0.1* 3.3* 7.4 10.1 14.2 18.0 22.2 25.5 0.54 0.6* 1.4* 11.5 17.3 22.8 24.3 36.4 38.8 1.59 0.8* 0.8* 6.5* 10.2 12.9 19.1 22.4 27.3

148

Log E 0.44 0.00 .19 .43 .67 .91 1.20 1.61 1.99 -0.3* 1.6* 4.2 8.9 11.1 15.7 18.9 23.0

Log El

*Note: These averages include at least one session datum for which no effect was reported.

J. M. Williams

Luminance and the Hess Effect


TABLE XIII HESS-EFFECT DATA t IN MILLISECONDS LD SESSIONS (n = 12) IN CHRONOLOGICAL ORDER log E = .11

149

Log El Mean -2.30 -1.89 -1.51 -1.26 -1.02 -.80 -.54 -.25 -.10 .16 .38 .57 .67 .81 1.08 1.58 1.98 2.37 7.6 4.1 3.6 5.8 5.5 2.0 3.5 1.6 -3.0 -1.7 1.9 0.5 -2.6 -1.7 1.6 0.6 -.6 0.3

1 s. d. 25.74 10.54 5.66 7.25 6.90 6.34 4.88 4.73 5.45 3.95 5.87 5.45 3.61 4.53 4.32 3.68 2.37 5.88 4.1 -1.9 2.8 12.1 12.6 10.5 Mean

2 s. d. Mean

3 s. d. Mean

4 s. d.

14.08

6.80 6.53 9.0 8.53

9.38 4.17 6.48 7.5 5.35 8.3 8.55

J. M. Williams

Luminance and the Hess Effect


TABLE XIV HESS-EFFECT DATA t IN MILLISECONDS LD SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El -1.4

150

Log E

-2.19

-1.78 1 2 s. d. 4.48 6.44 7.53 7.32 7.29 12.08 17.51 10.21 33.81 Mean 9.9 25.4 12.8 42.9 52.0 80.2 73.1 96.0 95.4 s. d. 7.91 9.84 13.15 13.18 13.14 14.18 15.92 12.65 18.18

Mean 0.00 .27 .53 .91 1.35 1.83 2.12 2.52 2.91 -1.8 -1.2 11.6 35.1 31.3 70.8 47.9 95.4 105.3

s. d. 18.87 17.69 7.43 13.90 14.96 13.82 15.85 11.32 22.90

Mean -1.0 20.8 23.2 27.7 43.1 64.4 70.3 72.8 68.9

s. d. 11.10 8.30 8.53 8.46 16.91 15.67 8.63 17.47 40.48

Mean 2.5 13.1 23.6 29.9 54.3 60.3 79.9 66.5 78.6

J. M. Williams

Luminance and the Hess Effect


TABLE XV HESS-EFFECT DATA t IN MILLISECONDS LD SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El

151

Log E Mean 0.00 .27 .53 .91 1.35 1.83 2.12 2.52 2.91 3.6 17.9 21.9 31.4 43.1 58.6 57.7 61.3 77.8

-.69 -1.15 -.91 1 s. d. 4.42 9.19 3.95 7.33 7.07 11.21 6.40 5.86 22.46 Mean -.6 9.3 17.1 34.2 42.5 52.0 61.0 69.1 66.2 s. d. 3.88 9.15 4.17 5.87 4.55 9.70 12.44 10.81 9.70 Mean -1.1 14.9 20.7 34.6 60.6 63.1 68.9 87.1 91.2 s. d. 8.02 6.14 5.65 6.79 6.90 9.24 12.09 11.99 14.22 Mean 4.3 20.1 31.1 44.7 58.5 82.6 69.3 81.8 75.5 2 s. d. 6.36 6.86 5.49 5.68 9.75 14.07 17.82 17.04 19.80

J. M. Williams

Luminance and the Hess Effect


TABLE XVI HESS-EFFECT DATA t IN MILLISECONDS LD SESSIONS (n = 12) IN CHRONOLOGICAL ORDER

152

Log El

Log E Mean 0.00 .27 .53 .91 1.35 1.83 2.12 2.52 2.91 -.7 11.0 16.7 28.1 34.7 59.1 52.1 51.9 75.0

-0.43 -.14 1 s. d. 5.91 6.14 5.15 5.86 7.33 6.58 3.96 9.76 11.75 Mean 0.8 19.2 26.1 38.3 49.3 67.3 69.7 67.7 68.8 2 s. d. 6.90 6.67 5.55 10.21 8.14 7.61 13.53 19.79 12.34 Mean -.4 21.3 20.8 37.9 46.1 65.7 73.2 60.9 69.1 3 s. d. 5.13 6.07 7.79 4.43 9.05 4.22 13.15 13.11 12.24 Mean -4.2 7.0 16.5 25.8 41.9 49.5 58.4 51.4 59.0 s. d. 5.05 4.38 6.26 7.35 5.23 8.83 10.28 9.44 14.61

J. M. Williams

Luminance and the Hess Effect


TABLE XVII HESS-EFFECT DATA t IN MILLISECONDS LD SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El

153

Log E Mean 0.00 .19 .27 .43 .53 .67 .91 1.20 1.35 1.60 1.83 1.99 2.12 2.52 2.91 33.4 47.2 21.0 28.8 33.1 9.5 -4.5 6.3

.27 0.01 1 s. d. 4.28 6.08 10.0 7.66 16.7 6.78 7.42 5.83 35.4 7.35 50.7 10.46 41.6 62.2 61.6 6.15 9.73 10.53 51.7 53.5 41.5 9.83 7.21 10.83 4.57 40.6 11.07 40.1 2.90 6.89 50.6 6.68 32.6 7.10 25.0 7.28 38.4 9.73 5.69 31.4 7.63 17.9 17.8 27.7 4.28 4.36 6.58 5.35 20.0 5.45 10.6 4.92 Mean -1.7 s. d. 6.32 Mean 2.5 2 s. d. 6.21 Mean 1.4 5.1 s. d. 2.64 4.21 .49

J. M. Williams

Luminance and the Hess Effect


TABLE XVIII HESS-EFFECT DATA t IN MILLISECONDS LD SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El

154

Log E Mean 0.00 .19 .29 .43 .67 .69 .91 1.08 1.20 1.60 1.99 22.3 8.2 4.4 -4.7

.78 .68 1 s. d. 5.52 Mean 0.7 -3.7 3.53 7.8 17.9 6.82 15.0 4.02 20.4 35.5 37.3 5.69 5.54 6.39 33.4 10.76 5.59 22.5 9.15 18.9 6.02 2.71 3.19 9.1 4.39 s. d. 4.31 4.43 0.9 4.07 Mean 2 s. d. Mean -.4 s. d. 2.75 .92

J. M. Williams

Luminance and the Hess Effect


TABLE XIX HESS-EFFECT DATA t IN MILLISECONDS LD SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El

155

Log E Mean 0.00 0.00* .19 .29 .40 .43 .67 .69 .79 .91 1.08 1.20 1.60 1.99 17.4 18.1 25.5 3.82 4.84 5.30 14.8 2.82 8.4 7.1 3.21 4.16 2.1 2.62 -1.3 1 s. d. 2.87

1.19 2 Mean 3.6 s. d. 5.69 Mean 0.9 7.3 -.5 3.93 -1.3 2.3 9.6 14.5 7.60 6.04 6.9 4.5 17.9 10.82 8.2 20.5 24.2 24.1 7.68 4.48 8,78 3.54 4.55 4.10 4.25 5.20 1 s. d. 3.51 5.77

1.69 2 Mean -5.1 3.7 s. d. 4.60 9.50

*Note:

Two additional sessions were run for 1.69 condition.

J. M. Williams

Luminance and the Hess Effect


TABLE XX HESS-EFFECT DATA t IN MILLISECONDS LD SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El = 2.37

156

Log E Mean 0.00 5.3

1 s. d. 4.31 Mean 6.4

2 s. d. 7.82 Log El = 2.48

1 Mean 0.00 .29 .69 1.08 -3.2 2.8 7.9 8.2 s. d. 5.05 6.02 5.08 5.55 Mean -3.7 9.7 17.9

2 s. d. 3.60 9.87 7.28 Mean 0.0

3 s. d. 8.24 Mean 8.0

4 s. d. 9.06

Log El = 3.56 Mean 0.00 1.1 s. d. 5.79

J. M. Williams

Luminance and the Hess Effect


TABLE XXI HESS-EFFECT DATA t IN MILLISECONDS AB SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log E = .11

157

Log El Mean -2.30 -1.89 -1.51 -1.26 -1.02 -.80 -.54 -.25 -.10 .16 .38 .57 .67 -.2 15.4 5.0 6.5 4.8 0.3 12.4 8.6 1.8 8.6 10.3 5.0 4.0

1 s. d. 9.68 6.87 6.61 7.35 5.76 3.54 2.40 4.85 2.01 1.90 1.02 1.95 1.69 Mean 2.0 1.4 -.9 2.2 8.3 -.7 4.3 10.9 9.2 -2.5 7.3 4.8 -.7

2 s. d. 7.88 9.45 5.31 5.74 4.02 4.56 1.78 2.01 2.06 1.81 1.70 1.70 2.05 6.6 1.5 0.0 -6.0 -7.1 Mean 12.8

3 s. d. 13.71 Mean

4 s. d.

5.28

6.61

4.89

2.31

-.4

1.91

3.18

(Continued next page)

J. M. Williams

Luminance and the Hess Effect


TABLE XXI (Continued) HESS-EFFECT DATA t IN MILLISECONDS AB SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log E = .11

158

Log El Mean .81 1.08 1.58 1.58* 1.98 2.37 2.6 -.2 -.3 -1.2 -.8 -2.9

1 s. d. 2.43 2.29 1.76 2.52 2.36 1.90 3.0 -1.1 Mean 2.2 2.5 -1.4

2 s. d. 1.46 2.20 1.24 5.1 -4.6 Mean

3 s. d. Mean

4 s. d.

4.13 2.51

0.7 -1.4

3.12 2.27

2.38 1.83

2.4

3.67

-.7

2.77

*Note:

One additional session was run for 1.58 condition.

J. M. Williams

Luminance and the Hess Effect


TABLE XXII HESS-EFFECT DATA t IN MILLISECONDS AB SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El = -2.19

159

Log E Mean 0.00 .27 .53 .91 1.35 1.83 2.12 2.52 2.91 3.7 15.5 16.3 24.6 36.5 49.8 61.6 69.6 97.6

1 s. d. 8.55 8.35 6.29 5.16 11.22 10.53 9.83 11.55 12.31 Mean 4.5 15.1 12.3 18.5 28.3 44.9 48.9 51.2 33.8

2 s. d. 10.18 7.59 7.88 9.06 5.62 8.06 10.05 8.51 12.81 Mean 7.4 10.3 6.8 15.4 23.5 43.8 50.0 61.8 60.5

3 s. d. 13.32 17.35 11.25 7.89 9.80 14.32 10.35 18.64 20.13

J. M. Williams

Luminance and the Hess Effect


TABLE XXIII HESS-EFFECT DATA t IN MILLISECONDS AB SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El = -1.78

160

Log E Mean 0.00 .27 .53 .91 1.35 1.83 2.12 2.52 2.91 5.2 13.7 21.1 27.1 39.1 60.5 69.7 68.7 73.7

1 s. d. 5.09 5.15 9.73 5.52 6.58 5.98 6.75 14.99 6.67 Mean -8.6 7.4 5.5 16.2 23.9 49.5 54.1 52.4 87.2

2 s. d. 7.22 6.55 7.96 7.24 11.17 6.26 11.92 9.59 6.71

J. M. Williams

Luminance and the Hess Effect


TABLE XXIV HESS-EFFECT DATA t IN MILLISECONDS AB SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El = -1.40

161

Log E Mean 0.00 .27 .53 .91 1.35 1.83 2.12 2.52 2.91 -6.6 10.6 20.9 28.4 47.0 63.8 75.8 87.6 96.5

1 s. d. 6.72 5.73 4.49 4.49 5.24 7.39 3.79 7.75 7.86 Mean -6.9 10.6 17.4 25.7 32.5 52.8 51.0 67.3 74.7

2 s. d. 9.16 5.27 6.36 6.67 4.98 8.88 7.50 8.80 14.13 Mean -6.8 -3.7 9.1 26.1 34.2 56.6 55.8 64.5 68.3

3 s. d. 8.95 3.89 7.49 4.27 7.84 5.06 9.47 14.67 16.69

J. M. Williams

Luminance and the Hess Effect


TABLE XXV HESS-EFFECT DATA t IN MILLISECONDS AB SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El = -1.15

162

Log E Mean 0.00 .27 .53 .91 1.35 1.83 2.12 2.52 2.91 0.4 9.7 22.5 27.5 42.4 56.2 71.4 77.1 78.0

1 s. d. 3.63 3.71 3.65 3.30 4.88 4.81 4.42 5.22 7.77 Mean -5.6 7.1 15.8 20.3 34.9 51.8 63.6 66.4 66.3

2 s. d. 5.65 5.55 4.34 3.99 6.76 6.99 4.77 10.32 8.49

J. M. Williams

Luminance and the Hess Effect


TABLE XXVI HESS-EFFECT DATA t IN MILLISECONDS AB SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El = -.91

163

Log E Mean 0.00 .27 .53 .91 1.35 1.83 2.12 2.52 2.91 4.9 10.4 19.7 35.9 49.3 62.6 66.3 74.2 81.8

1 s. d. 5.26 4.67 2.98 3.30 5.37 5.44 4.82 5.88 7.52 Mean -1.7 16.1 20.3 40.6 48.5 71.0 79.1 80.1 97.4

2 s. d. 3.36 3.44 5.83 2.91 5.05 5.88 4.70 5.41 6.79 Mean -7.1 9.3 14.8 23.8 35.5 54.1 62.1 58.2 66.9

3 s. d. 6.73 4.70 3.44 3.57 3.78 5.51 4.84 8.13 8.17

J. M. Williams

Luminance and the Hess Effect


TABLE XXVII HESS-EFFECT DATA t IN MILLISECONDS AB SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El = -.69

164

Log E Mean 0.00 .27 .53 .91 1.35 1.83 2.12 2.52 2.91 0.2 9.0 21.0 33.9 47.1 68.7 73.7 83.6 84.2

1 s. d. 3.44 4.53 2.72 3.33 2.31 5.63 4.22 6.75 3.40 Mean -5.0 10.5 16.8 28.2 44.3 52.0 63.7 63.0 78.7

2 s. d. 4.74 1.98 3.57 3.10 2.41 4.48 3.65 4.69 5.09

J. M. Williams

Luminance and the Hess Effect


TABLE XXVIII HESS-EFFECT DATA t IN MILLISECONDS AB SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El = -.43

165

Log E Mean 0.00 .27 .53 .91 1.35 1.83 2.12 2.52 2.91 2.7 20.8 24.0 44.8 52.5 67.6 69.4 70.8 56.8

1 s. d. 2.12 3.07 3.53 2.71 2.83 3.93 4.04 3.14 5.15 Mean -.7 10.3 11.2 30.0 36.1 53.1 53.8 55.4 56.7

2 s. d. 3.51 4.32 2.82 4.56 2.79 2.94 5.37 3.23 4.87 Mean -1.3 13.9 12.9 29.4 36.4 53.1 55.6 60.9 54.4

3 s. d. 5.35 3.51 3.42 7.29 2.66 5.03 3.08 5.01 6.26

J. M. Williams

Luminance and the Hess Effect


TABLE XXIX HESS-EFFECT DATA t IN MILLISECONDS AB SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El

166

Log E Mean 0.00 .19 .27 .43 .53 .67 .91 1.20 1.35 1.60 1.83 1.99 62.5 3.42 47.7 2.54 37.4 2.47 22.4 2.12 16.5 2.61 6.5 1 s. d. 3.08

-.14 2 Mean -5.1 s. d. 2.82 Mean 3.8 10.4 13.1 2.41 19.0 14.9 2.69 28.3 29.7 2.69 34.8 44.6 40.3 4.00 49.7 50.7 5.44 57.1 3.81 1.76 3.47 3.12 3.67 2.37 1 s. d. 3.63 1.53

.01 2 Mean 8.2 2.8 s. d. 2.27 2.51

19.1

0.95

23.9 38.6 39.7

2.36 2.57 3.10

52.5

2.86

49.5

3.28

J. M. Williams

Luminance and the Hess Effect


TABLE XXX HESS-EFFECT DATA t IN MILLISECONDS AB SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log E = .27

167

Log El Mean 0.00 .27 .53 .91 1.35 1.83 2.12 2.52 2.91 4.4 14.8 21.7 32.8 47.1 47.0 45.9 49.8 47.5

1 s. d. 1.78 2.44 2.77 2.19 2.61 3.40 2.22 2.69 4.69 Mean -2.7 10.1 18.4 33.5 38.2 40.6 46.8 41.7 39.4

2 s. d. 2.72 1.81 2.31 4.48 3.17 3.26 2.90 3.72 3.71 Mean 1.6 10.9 15.2 25.9 37.0 45.1 42.9 44.4 43.3

3 s. d. 3.08 1.98 2.38 2.08 1.58 2.12 3.19 2.45 5.01 Mean -5.8 2.2 10.9 20.8 31.6 35.4 36.9 35.4 36.4

4 s. d. 2.71 2.52 3.71 2.76 3.04 2.44 3.50 2.50 2.40

J. M. Williams

Luminance and the Hess Effect


TABLE XXXI HESS-EFFECT DATA t IN MILLISECONDS AB SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El

168

Log E Mean 0.00 .19 .29 .43 .67 .69 .91 1.08 1.20 1.60 1.99 38.7 45.2 40.7 2.77 3.30 2.05 35.0 1.31 18.2 25.5 2.43 3.68 9.2 13.8 1 s. d. 2.05 1.59

.49 2 Mean -1.3 8.7 s. d. 3.65 1.95 11.0 17.6 24.3 2.62 3.12 24.5 30.8 1.94 30.6 36.4 40.2 41.0 2.61 4.09 2.71 2.40 1.98 1.06 Mean -1.5 1 s. d. 1.84

.68 2 Mean -.3 s. d. 2.09

6.9

2.09

20.0

2.54

28.5

1.28

J. M. Williams

Luminance and the Hess Effect


TABLE XXXII HESS-EFFECT DATA t IN MILLISECONDS AB SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El = .78

169

Log E Mean 0.00 .19 .43 .67 .91 1.20 1.60 1.99 2.1 10.3 16.7 26.6 28.4 33.2 30.7 28.7

1 s. d. 1.94 1.44 1.98 3.65 2.93 3.56 2.65 3.03 Mean -1.8 1.0 12.6 15.9 27.3 26.0 27.7 30.6

2 s. d. 3.00 1.74 2.18 1.66 2.05 1.67 6.40 3.90 23.6 25.7 25.3 28.7 Mean -2.3

3 s. d. 3.72

3.03 4.75 3.28 3.81

J. M. Williams

Luminance and the Hess Effect


TABLE XXXIII HESS-EFFECT DATA t IN MILLISECONDS AB SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El = .92

170

Log E Mean 0.00 .29 .69 1.08 -1.6 10.7 22.4 28.1

1 s. d. 2.82 1.51 1.51 3.46 Mean -.3 10.6 20.5 25.5

2 s. d. 2.05 1.94 1.84 3.61

J. M. Williams

Luminance and the Hess Effect


TABLE XXXIV HESS-EFFECT DATA t IN MILLISECONDS AB SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log E = 1.19

171

Log El Mean 0.00 .19 .43 .67 .91 1.20 1.60 1.99 -2.5 -.1 10.0 12.2 19.8 23.1 19.1 24.7

1 s. d. 2.57 2.58 1.67 1.32 1.63 0.93 2.45 2.62 Mean -1.8 5.1 9.5 18.8 18.1 17.8 18.2 21.6

2 s. d. 2.20 1.69 1.56 1.72 1.87 1.87 2.47 1.92 Mean 0.7 1.4 8.0 11.6 17.4 17.7 17.1 21.1

3 s. d. 4.35 2.80 3.33 1.95 2.82 1.44 3.08 3.21 Mean -3.4 -.4 4.8 11.2 12.3 15.9 19.7 24.3

4 s. d. 2.94 3.70 2.38 2.89 2.64 1.98 1.84 2.36

J. M. Williams

Luminance and the Hess Effect


TABLE XXXV HESS-EFFECT DATA t IN MILLISECONDS AB SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El

172

Log E Mean 0.00 0.00* 0.00* .29 .40 .69 .79 1.08 -3.7 -6.9 0.1 5.6 0.7 11.0 4.7 12.1 1 s. d. 1.44 2.52 3.39 3.65 2.09 2.57 1.66 2.44

1.69 2 Mean -2.9 -3.8 s. d. 2.98 2.44 Mean -6.7 1 s. d. 2.05

2.37 2 Mean -2.3 s. d. 2.31

2.1 -.5 5.8 3.7 11.7

2.59 3.70 1.41 2.27 2.11

*Note:

Five sessions were run for the 1.69 condition.

J. M. Williams

Luminance and the Hess Effect


TABLE XXXVI HESS-EFFECT DATA t IN MILLISECONDS AB SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log E = 1.19

173

Log El Mean 0.00 0.00 .18 .29 .29 .69 .69 1.08 -5.4 2.0 -1.3 -3.0 1.3 1.1 2.9 0.4

1 s. d. 1.51 6.67 1.16 2.86 2.41 1.34 3.07 1.38 Mean -5.1 0.8 0.6 -2.0 -2.5 -1.5 -2.2 0.1

2 s. d. 2.37 3.53 3.23 2.20 3.21 1.76 1.67 2.45 5.5 0.5 0.9 Mean -3.5 -1.8

3 s. d. 2.65 3.93 Mean -3.3 -4.0

4 s. d. 2.13 4.18

1.62

2.4

2.33

1.70

3.1

3.86

3.30

3.7

2.71

Log El = 3.56 1 Mean 0.00 -6.5 s. d. 2.91 Mean -3.4 2 s. d. 2.50

*Note: The Dissertation ms says "Six sessions each run for these conditions". However, there are eight for the "0.00" condition and six for the ".29" and ".69". It seems that the count in this Note was wrong.

J. M. Williams

Luminance and the Hess Effect


TABLE XXXVII HESS-EFFECT DATA t IN MILLISECONDS JMW SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log E = .11

174

Log El Mean -2.30 -1.51 -1.51 (Vald.) -1.02 -.54* .16 .67 1.08 1.08 (Vald.) 1.58 1.98 1.98 (Vald.) 2.4 -9.7 -2.2 -3.2 -2.5 1.5 6.8 -.1 5.5 -4.0 -3.2 3.3

1 s. d. 30.55 10.39 48.52 5.49 6.79 4.57 3.26 4.25 3.19 2.72 1.94 3.14 Mean -5.1 -2.0 -13.4 -2.2 -.3 9.2 7.1 6.0 1.9 5.0 2.6 1.9

2 s. d. 31.08 13.27 5.76 5.17 4.71 3.11 2.38 2.48 2.77 4.02 2.89 3.70 4.8 4.2 4.4 9.7 10.7 1.4 9.3 Mean 2.0 9.9

3 s. d. 37.26 13.66 Mean 4.4 8.3

4 s. d. 19.77 5.59

4.75 5.94 3.75 5.49 2.51

6.2 5.5 11.6 10.8 5.7

5.77 6.30 3.78 3.99 3.04

2.76 2.26

6.3 3.6

2.48 3.92

*Note: The Dissertation ms included a Note saying, "A fifth session was run with mean == 9.7 ms, s. d. = 5.93 ms". This should have been fourth; the changes to be made are now in the Table Session 3 (these were discovered from raw data after the Dissertation had been approved). The notation "Vald." in the Log El column means that these data were validation data; they were gathered by paid student worker JV, who operated the apparatus as E, with JMW as S.

J. M. Williams

Luminance and the Hess Effect


TABLE XXXVIII HESS-EFFECT DATA t IN MILLISECONDS JMW SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El = -2.19

175

Log E Mean 0.00 .27 .53 .91 1.35 1.83 2.12 2.52 2.91 -3.2 8.2 18.6 21.2 39.0 54.9 76.9 122.3 137.2

1 s. d. 21.50 13.16 12.52 11.31 13.15 9.13 19.13 20.41 11.62 Mean -2.5 7.1 21.1 15.7 39.8 68.0 83.4 121.3 148.0*

2 s. d. 22.74 18.91 13.01 17.11 9.37 14.67 11.73 7.00 18.94* Mean -1.2 15.4 15.9 40.4 37.5 68.3 77.7 111.6 137.5

3 s. d. 14.6 9.27 9.50 18.49 8.45 12.10 12.37 15.23 7.71 Mean -.5 12.9 26.5 38.0 49.7 77.4 103.3 112.1 132.3

4 s. d. 15.98 14.57 11.63 10.29 15.16 8.14 8.94 12.47 10.10

*Note: Only 5 of the 12 settings could be made in this session because target displacement repeatedly exceeded measuring-scale limits. Data in the 1.35 row, Session 1, have been corrected according to a recheck of the raw data after the Dissertation ms was approved.

J. M. Williams

Luminance and the Hess Effect


TABLE XXXIX HESS-EFFECT DATA t IN MILLISECONDS JMW SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El = -1.40

176

Log E Mean 0.00 0.00 (Vald.) .27 .27 (Vald.) .53 .53 (Vald.) .91 .91 (Vald.) 1.35 1.35 (Vald.) 1.83 1.83 (Vald.) -11.6 -5.0 -4.4 15.7 5.3 26.9 19.3 40.9 41.5 61.3 56.4 76.4

1 s. d. 10.30 4.31 9.09 6.87 7.70 8.30 6.69 4.22 5.26 5.81 10.95 7.03 Mean -1.9 4.3 14.7 22.7 15.0 14.1 34.6 31.0 48.6 41.2 73.0 67.5

2 s. d. 14.08 3.51 5.52 5.74 8.99 7.98 10.58 5.42 5.01 7.13 5.91 5.17 82.1 49.5 34.8 18.0 6.2 Mean -1.4

3 s. d. 9.77 Mean 2.1

4 s. d. 6.94

11.03

10.1

4.78

7.07

25.8

8.17

9.70

34.5

5.28

4.87

57.7

6.01

8.03

77.3

5.33

(Continued next page)

J. M. Williams

Luminance and the Hess Effect


TABLE XXXIX (Continued) HESS-EFFECT DATA t IN MILLISECONDS JMW SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El = -1.40 (Continued)

177

Log E Mean 2.12 2.12 (Vald.) 2.52 2.52 (Vald.) 2.91 2.91 (Vald.) 74.4 93.5 96.9 110.8 127.8* 131.1

1 s. d. 9.27 5.61 12.63 8.44 10.85* 8.73 Mean 84.9 78.2 104.2 106.2 113.7 113.6

2 s. d. 6.82 8.42 10.19 9.64 13.27 12.97 122.8 112.9 Mean 81.9

3 s. d. 7.95 Mean 92.6

4 s. d. 5.01

6.05

116.7

6.76

14.42

126.1

10.43

*Note: The target at lower luminance was visible only occasionally on the day this session was run; many of the settings of the 12 made were recorded without certainty of whether the vernier alignment was to the lower target or to a streak of stray light.

J. M. Williams

Luminance and the Hess Effect


TABLE XL HESS-EFFECT DATA t IN MILLISECONDS JMW SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El = -.91

178

Log E Mean 0.00 .27 .53 .91 1.35 1.83 2.12 2.52 2.91 -12.6 6.2 11.6 25.2 43.0 66.7 81.2 90.6 100.7

1 s. d. 3.67 6.07 4.41 4.63 3.90 5.97 7.43 6.07 11.20 Mean -5.1 4.4 11.2 31.0 51.2 73.8 82.1 106.9 114.0

2 s. d. 10.51 8.56 4.67 7.24 5.55 9.70 5.98 6.09 8.20 Mean -1.3 15.7 23.3 39.8 67.1 83.6 98.6 113.8 127.6

3 s. d. 5.72 7.72 6.64 7.32 3.64 6.43 4.48 6.92 5.59 Mean 0.8 24.1 29.1 41.7 65.4 82.9 94.9 115.5 116.6

4 s. d. 7.52 6.50 5.34 5.08 4.28 5.66 5.44 4.95 6.82

J. M. Williams

Luminance and the Hess Effect


TABLE XLI HESS-EFFECT DATA t IN MILLISECONDS JMW SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El = -.43

179

Log E Mean 0.00 0.00* .27 .53 .91 1.35 1.83 2.12 2.12** 2.52 2.91 -8.2 -1.4 4.3 9.7 24.7 42.2 57.9 65.0 83.2 73.3 67.7

1 s. d. 6.07 2.77 5.38 8.20 5.66 5.62 7.01 9.08 4.48 7.40 8.95 81.7 79.3 Mean -.3 -5.7 13.2 14.0 38.9 51.1 68.0 81.4

2 s. d. 6.07 5.90 5.79 3.57 4.62 3.04 7.92 3.44 Mean 1.0 -2.3 14.6 21.1 37.9 53.5 70.4 79.0

3 s. d. 7.6 5.79 3.89 5.74 5.61 5.17 6.36 6.14 Mean 1.0 -8.0 18.5 23.8 42.4 53.5 ** 84.3

4 s. d. 3.78 4.75 3.93 4.69 1.88 3.47 ** 4.38

6.57 9.54

84.1 80

4.22 7.04

85.1 85.4

4.74 5.44

*Note :

Additional 4 sessions run under this condition.

**Note: An additional 2.12 condition was run inadvertently in place of a 1.83 condition.

J. M. Williams

Luminance and the Hess Effect


TABLE XLII HESS-EFFECT DATA t IN MILLISECONDS JMW SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El = .27

180

Log E Mean 0.00 .27 .53 .91 1.35 1.83 2.12 2.52 2.91 3.29 0.6 7.9 20.0 30.3 37.3 47.7 44.1 44.6 47.9 47.1

1 s. d. 5.72 2.72 3.95 3.17 5.31 3.33 3.47 4.24 4.42 3.61 Mean -.3 14.7 15.8 29.2 41.1 41.4 42.6 29.5 30.5 47.9

2 s. d. 4.28 4.48 2.54 2.16 4.39 3.46 2.86 4.28 3.74 5.37 Mean 2.2 19.3 23.7 35.6 48.4 48.9 48.8 47.5 46.8 45.2

3 s. d. 2.93 3.25 3.14 3.01 3.12 4.10 3.49 3.65 4.69 4.45 Mean 0.7 19.0 21.8 31.2 45.6 49.3 50.0 48.9 52.7 52.4

4 s. d. 3.96 3.00 3.33 2.83 3.14 2.59 3.63 3.57 5.61 4.63

J. M. Williams

Luminance and the Hess Effect


TABLE XLIII HESS-EFFECT DATA t IN MILLISECONDS JMW SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El = .78

181

Log E Mean 0.00 .52 .91 .91 (Vald.) 1.20 1.20 (Vald.) 1.60 1.99 2.78 2.1 18.4 22.4 23.8 27.5 27.5 27.3 29.4 38.5

1 s. d. 4.82 2.19 3.53 4.59 3.63 2.15 3.86 3.47 3.42 Mean 1.5 19.0 22.7 24.6 26.0 31.3 28.3 29.7 38.9

2 s. d. 4.28 2.75 4.28 3.76 2.04 3.32 3.81 5.45 3.40 34.6 36.7 36.3 32.2 Mean 1.2 16.1 26.9

3 s. d. 4.48 2.45 2.87 Mean 3.5 17.0 26.2

4 s. d. 4.10 3.03 4.21

3.53

30.1

3.18

1.81 2.65 3.51

33.2 33.2 39.0

3.05 2.13 4.07

J. M. Williams

Luminance and the Hess Effect


TABLE XLIV HESS-EFFECT DATA t IN MILLISECONDS JMW SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El = 1.19

182

Log E Mean 0.00 0.00 (Vald) .19 .19 (Vald.) .43 .43 (Vald.) .67 .67 (Vald.) .91 .91 (Vald.) 1.20 1.20 (Vald.) -.7 5.1 -.8 5.9 7.8 12.4 12.2 16.2 15.6 18.1 20.1 25.6

1 s. d. 3.32 5.12 3.30 3.19 1.72 3.65 2.47 2.75 3.51 4.57 4.46 7.92 Mean 1.2 1.8 6.8 3.7 8.1 11.5 16.1 14.3 14.0 21.9 17.1 20.5

2 s. d. 5.05 2.33 3.40 2.44 4.16 2.68 2.52 2.12 3.43 1.78 4.25 3.63 22.9 18.0 17.7 11.3 9.0 Mean 4.6

3 s. d. 2.31 Mean 2.1

4 s. d. 2.11

2.62

7.4

2.73

3.40

10.2

2.65

2.33

17.8

3.05

2.01

18.3

2.31

1.76

23.8

2.08

(Continued next page)

J. M. Williams

Luminance and the Hess Effect


TABLE XLIV (Continued) HESS-EFFECT DATA t IN MILLISECONDS JMW SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El = 1.19 (Continued)

183

Log E Mean 1.60 1.60 (Vald) 1.99 1.99 (Vald.) 2.37 22.3 25.2 23.8 30.0 30.5

1 s. d. 2.59 4.06 3.56 3.65 2.01 Mean 19.8 22.4 21.5 28.0 32.0

2 s. d. 2.80 2.96 2.69 3.63 4.38 28.8 24.5 Mean 22.0

3 s. d. 2.72 Mean 24.5

4 s. d. 2.24

1.99

29.3

4.10

2.34

33.0

1.74

J. M. Williams

Luminance and the Hess Effect


TABLE XLV HESS-EFFECT DATA t IN MILLISECONDS JMW SESSIONS (n = 12) IN CHRONOLOGICAL ORDER Log El = 2.48

184

Log E Mean 0.00 0.00* 0.00 (Vald) .18 .18 (Vald.) .29 .29* .29 (Vald.) .69 .69* .69 (Vald.) 1.08 1.08 (Vald.) 0.4 2.7 1.4 0.2 2.3 1.0 0.2 3.8 4.1 7.6 6.9 6.0 10.3

1 s. d. 2.05 1.52 2.16 1.42 2.43 2.08 2.09 1.67 1.70 1.73 2.15 3.46 2.69 Mean -2.5 2.7 1.5 -.7 1.5 3.0 -.6 1.9 2.2 8.9 5.1 9.6 9.6

2 s. d. 3.30 2.08 3.17 3.14 1.69 2.29 1.97 1.03 1.81 1.38 1.98 3.39 2.86 12.7 6.2 10.4 5.0 7.0 Mean -1.3 -.5 0.8 5.4

3 s. d. 2.47 2.48 1.69 1.55 Mean -2.7 0.3 1.9 3.6

4 s. d. 2.30 2.31 2.87 3.05

2.37 1.34

6.5 3.6

1.91 1.84

2.82 1.46

4.0 7.3

2.08 1.60

1.48

13.1

2.29

*Note:

An additiona1 4 sessions were run for this condition.

Also, in the 0.00 validation, Session 4, the data are corrected slightly because of recalculation from the raw data after the Dissertation ms had been approved.

J. M. Williams

Luminance and the Hess Effect


TABLE XLVI HESS-EFFECT DATA t IN MILLISECONDS LD, AB, and JMW SESSIONS (n = 12) IN CHRONOLOGICAL ORDER FOR Log El = 2.09 1 2 s. d. Mean s. d. LD Mean 3 s. d. Mean 4

185

Log E

Mean

s. d.

0.00 .39

-1.8 3.0

2.37 3.19

-5.4 4.6

6.47 6.48 AB

0.00 .39

-3.7 -1.3

2.15 1.65

1.4 4.6

2.38 1.77 JMW

2.0 3.6

1.78 3.28

-3.8 -2.5

2.04 2.69

0.00 0.00 (Vald.) .39 .39 (Vald.)

-4.0 0.7 0.3 5.9

2.55 2.59 2.43 2.84

0.0 -1.0 4.1 3.2

1.78 2.50 1.53 2.71

2.0

1.55

1.1

2.13

6.3

3.21

5.7

2.52

J. M. Williams

Luminance and the Hess Effect


TABLE XLVII HESS~EFFECT AVERAGED SAMPLE STANDARD DEVIATIONS OF SETTINGS FOR SELECTED VALUES OF LOG El t IN MILLISECONDS Log El -2.19 -1.78 -1.40 -1.15 -.91 -.69 -.43 -.14 .01 .27 .49 .68 .78 .92 1.19 1.69 2.37 2.48

186

Log E = 0.00
AB 10.683 6.155 8.277 4.640 5.117 4.090 3.660 2.950 2.950 2.573 2.850 1.965 2.887 2.435 3.015 2.210 2.180 3.371 2.364 2.364 4.420 4.223 5.341 6.855 8.152 JMW 18.705

J. M. Williams

Luminance and the Hess Effect


TABLE XLVIII HESS~EFFECT AVERAGED SAMPLE STANDARD DEVIATIONS OF SETTINGS FOR SELECTED VALUES OF LOG El t IN MILLISECONDS Log El -2.30 -1.89 -1.51 -1.26 -1.02 -.80 -.54 -.25 -.10 .16 .38 .57 .67 .81 1.08 1.58 1.98 2.37

187

Log E = 0.11
AB 10.423 8.160 5.733 6.545 5.463 4.050 3.023 3.430 2.035 1.983 1.360 1.825 2.307 1.945 2.935 2.060 2.795 1.865 3.040 2.995 2.975 3.780 3.803 5.356 5.295 16.198 JMW 29.665

J. M. Williams

Luminance and the Hess Effect


TABLE XLIX HESS~EFFECT AVERAGED SAMPLE STANDARD DEVIATIONS OF SETTINGS FOR SELECTED VALUES OF LOG El t IN MILLISECONDS

188

Log El -2.19 -1.78 -1.40 -1.15 -.91 -.69 -.43 -.14 .01 .27 .49 .78 1.19

Log E = 0.91
AB 7.370 6.380 5.143 3.645 3.260 3.215 4.853 2.580 2.845 2.878 1.625 2.670 2.240 3.873 2.935 2.793 4.443 6.068 6.982 JMW 14.300

J. M. Williams

Luminance and the Hess Effect


TABLE L HESS-EFFECT LOW-LEVEL DATA FOR THEORETICAL FIT t IN MILLISECONDS Log El

189

Log E Datum 0.00 .27 .53 .91 1.35 1.83 2.12 2.52 2.91 3.29 -1.85 10.90 20.53 28.83 41.78 67.15 85.33 116.83 138.75 ----

-2.19* Code [101] [102] [103] [104] [201] [202] [203] [204] [205] Datum -2.25 10.83 17.52 32.52 49.97 72.12 84.25 107.95 122.52 ----

-1.40* Code [105] [106] [107] [206] [207] [208] [209] [210] [211] 0.00 0.15 12.89 19.56 30.27 41.23 44.67 45.13 45.75 44.76 48.15

0.27** Code [301] [302] [303] [304] [305] [306] [307] [308] [309] [310]

*Note:

JMW data

**Note: Combined LD, AB, and JMW data. The "Code" values refer to inferred scotopic and photopic conditions and are explained in the text.

J. M. Williams

Luminance and the Hess Effect


TABLE LI HESS-EFFECT HIGH-LEVEL DATA FOR THEORETICAL FIT t IN MILLISECONDS Log El

190

Log E Datum 0.00 .18 .19 .29 .43 .67 .69 .91 1.08 1.20 1.60 1.99 2.37 0.78 ---3.30 ---9.30 14.14 ---17.18 ---20.20 21.05 24.87 31.08

1.19* Code [311] Datum -0.61 1.45 [312] ---2.28 [313] [314] ------5.13 [315] ---7.20 [316] [317] [318] [319] ----------

2.48* Code [320] [321]

[322]

[323]

[324]

*Note:

Combined LD, AB, and JMW data.

J. M. Williams

Luminance and the Hess Effect


TABLE LII THEORETICAL FITTED PARAMETER VALUES AND DEVIATIONS FROM HESS-EFFECT DATA

191

Theory a Exner b Dev. a Charpentier n Dev. k Dev. T Dev. K Alpern T0 4 = n Dev. K Alpern T0 n Dev. Y Enright S Dev. a (present) k Dev.

Hess-Effect Data: -2.19 .131 .319 -1.40 .134 .202 .27 .077 .0002

Log El 1.19 .006 .009 2.48 .0171 .0001

4.16
8.38 2.70

34.83
65 4.57

6.19
67 2.0

1.55
83 3.1

.33
81 3.1

1.20
-.16 18.77 14.7

13.13
.05 70.02 17.74

1.73
.11 37.71 8.075

1.59
3.2 19.45 6.075

.34
100000 4.64 3.02

Hecht Roufs

2.42
20.3 79.0 4

3.90
86 151 4

10.53
67.9 66.0 4

3.18
75.2 44.4 4

.41
-.5 .7 4

2.38
20.4 79.5 3.97

10.09
204 265 9.7

4.49
65.8 49.2 1.10

1.58
72.1 42.1 3.8

.32
-2.5 .7 1.4

2.25
42.8 4.82

6.52
42.4 6.58

1.73
42.2 -15.0

1.57
26.9 -6.10

.29
24.9 -6.10

2.67
.150 .28

1.80
1.0 .196

2.59
.149 .023

1.82
.177 .130

.38
.140 .104

1.55

5.69

4.37

1.85

.31

Note: Dev. = deviation defined as [SS/(n - 1)]1/2, in which SS represents the sum of squares of deviations, and n the number of datum-points fit as given in Tables L and LI. Parentheses around the Dev. values appeared in the Dissertation ms but are omitted here. A typographical error in the Dissertation ms has been corrected: column for log El = .27 was labelled incorrectly as "-.27". The

J. M. Williams

Luminance and the Hess Effect

192

FIG. 1. Visual perception lag in an eye sketched as though seen from above. The eye is fixated at some motionless point. A luminous rectangular object moves as shown, and the object Eo creates an image Ei on the retina at any given instant in time. The earliest perceptual response Pi to that image by assumption occurs at some retinal location previously stimulated by Ei; at the instant shown, the percept thus is localized in object-space at Po. The time t required for the luminous rectangle to move to Eo from its previous location (of Po) is the absolute visual latency of perception in this example. For a rectangle moving at constant speed v, measurement of the distance s between Po and Eo (or between Pi and Ei) in principle allows t to be computed as t = s/v.

J. M. Williams

Luminance and the Hess Effect

193

FIG 2. The Hess effect. On the right, pairs of vertically aligned, luminous rectangular objects are depicted moving from left to right. The target-objects are shown trailed by their corresponding percepts (dashed outlines) at left. In the upper row A, both rectangles have the same luminance and hence the same troland value, so they are perceived veridically as being aligned directly one above the other. In row B, the luminance of the lower rectangle has been reduced, so the visual response is retarded and an apparent relative lag is reported, the lag being given the phenomenal magnitude "s." This lag is the phenomenal Hess effect. In row C, the same 1uminance-difference as in B exists, but the upper rectangle has been displaced physically so that the apparent lag is just cancelled. The rectangles are reported as aligned in C, but the Hess effect still occurs, but now with measurable magnitude s (= s1 - s2 in the context of Figure l above). The spatial "extents" of the percepts in A, B, and C have not been kept the same, because there is no good reason why these extents should be the same.

J. M. Williams

Luminance and the Hess Effect

194

FIG. 3. Sketch of hypothetical pathways for the optic eyeblink reflex, possibly the shortest-latency (50 - 110 ms) overt visualbehavioral response known. The pontine nuclei of cranial nerves V and VII are shown caudal to their normal locations for illustrative clarity. A retinal ganglion cell at right is shown projecting to the superior colliculus (S), and another retinal ganglion cell is shown projecting to the lateral geniculate nucleus (L). OR = optic radiations, ICT = internal corticotectal tract, pt = pretectum, t = tectum. Brodmann's cerebral-cortical Areas 17, 18, 19, and 22 are represented at upper left; an efferent path of cranial nerve VII also is depicted. The tactile eyeblink reflex to a touch of the cornea probably is mediated via cranial nerve V, as shown. After scheme of Robinson (1968) and comments of Duke-Elder and Wybar (1961, pp. 771 f.). The importance of the retinocollicular direct afferent pathway in humans has been questioned (see Warwick and Williams, 1973. p. 887).

J. M. Williams

Luminance and the Hess Effect

195

FIG. 4. Upper panel A: Hypothetical time-course of rod outer segment (ROS) plasma-membrane hyperpolarization in response to light, based on text of Baylor, Lamb, and Yau (1979). As conjectured by the present author, the concentration u(x, t) of some diffusible substance or structural configuration is graphed in the upper panel as a function of distance x along the ROS sketched below the abscissa. Hyperpolarization is assumed linearly related to u. Stimulation is shown beginning at time t1 and the (normalized) peak response is shown achieved at time t5. In the lower panel B, the black dots represent plasma membrane pores closed in response to light.

J. M. Williams

Luminance and the Hess Effect

196

FIG. 5. Theoretical system accounting for intensity-dependent latency as discussed in the text. The curved limiting membranes of the retina enclose squares representing ganglion cell single units or small groups. The dashed lines enclose small local retinal regions responding essentially independently of one another. The target images of illuminance El and Eu sweep across the retina in a direction perpendicular to the plane of the figure. No image (top of figure) leads to a no-effect time-varying retinal output g0(t). The centroids t(gl) and t(gu) determine any perceived target relative displacement as described in the text.

J. M. Williams

Luminance and the Hess Effect

197

FIG. 6. The Hess-effect apparatus. The lightbox (LB) is at left and the target-movement mechanism (TMM) at right; the artificial pupil for the observer would be still farther to the right but is not shown. a = main-motor belt-and-pulley, b & b' = jackscrew shafts, g & g' = jackscrew gears, c = chain, m = movable gear, s = screwshaft for target relative lateral displacement, T = takeup arm, W = counterweight, r = readout scale, p & p' = target platforms with followers (not shown) engaging jackscrew grooves, f & f' = optical filter-boxes in target illumination channels, t & t' = upper and lower targets. F shows the location of the light-baffled cooling fan for the LB. Not shown, among other components, are: The parallel stainlesssteel guide-rails for m and for p & p': light-sources or cooling fans for f & f'; main motor or variable-speed transmission; the targetdisplacement motor at top of s; plexiglas safety shield covering the entire closest side of the target-movement mechanism; target display fieldstop: light-proof front cover plates of the target-movement mechanism; angle-iron rail (far side of sketch) which is used to hold the LB and target-movement mechanism at a fixed operating distance from the artificial pupil.

J. M. Williams

Luminance and the Hess Effect

198

FIG. 7. Isometric sketch of the lightbox showing the viewing aperture (15 cm x 45 cm) and the slot which positions the lightbox on the apparatus table. The dashed line locates the front of the false bottom mentioned in the text.

J. M. Williams

Luminance and the Hess Effect

199

FIG. 8. Arrangement of lamps within the lightbox as seen from inside, looking along the optic axis toward the observer's artificial pupil. With standard commercial l00-watt bulbs (A-19 glass envelope; medium bases), the filaments are located about 9 cm from each of the two nearest inside surfaces of the lightbox. The lamp-bases are simple white porcelain and are arranged symmetrically in pairs around the viewing aperture, with bases holding lamps #2 and #3 about 15 cm apart. Bases holding #5 and #6, of course, also are 15 cm apart. The glass envelop of an A-19 bulb is about 9 cm tall and 6 cm in diameter. The lightbox was used for Hess target luminance calibration with 15-watt A-19 bulbs.

J. M. Williams

Luminance and the Hess Effect

200

FIG. 9. View of the interior of the lightbox in a median plane passing through lamps #2 and #6. The geometric construction is explained in the text and allows maximum tolerable luminance gradient on the visible reflecting surface (to right) to be estimated as a function of design depth of the lightbox. The depth of the box as shown in the construction is equal to x + 9 cm

J. M. Williams

Luminance and the Hess Effect

201

FIG. 10. Reaction times for all observers as a function of target troland value. Each symbol represents a mean of 48 trials except the leftmost of LD's, which represents a mean of 12.

J. M. Williams

Luminance and the Hess Effect

202

FIG. 11. Pulfrich-effect t (= s/v) for AB as compared with averaged results for Lit's (1949) observers AH and CGM. The parameter is log troland value for the target of lower illuminance; the variable on the abscissa is the log troland value for the target of upper (higher) illuminance. Each Lit (1949, Table I) datum is averaged over 12 or 24 settings; the AB data represent averages over 16 settings.

J. M. Williams

Luminance and the Hess Effect

203

FIG. 12. Pulfrich-effect data for AB of Figure ll above replotted for comparison with present data of JMW. Each symbol represents a mean of 16 settings with parameter target log troland value for the lower-illuminated eye; the variable on the abscissa is the log troland value for the upper- (= more highly) illuminated eye.

J. M. Williams

Luminance and the Hess Effect

204

FIG. 13. Theoretical fit of Lit's (1949, Table I) Pulfricheffect data. The vertical scale for this figure is expanded as compared with previous figures above. The dots represent data many of which were given in Figure ll above; the smooth curves were fit by eye and generated from equation (35) of the text, using parameters a and k of .16 and .10 respectively.

J. M. Williams

Luminance and the Hess Effect

205

FIG. 14. Key to Hess-effect plotting symbols used in the figures below. To the right of each symbol is the target log troland parameter value for data represented by that symbol.

J. M. Williams

Luminance and the Hess Effect

206

FIG. 15. Hess-effect t as a function of upper target log troland value; LD's mean response for 24 settings. In this illustration, the parameter values represented by the plotting symbols are provided at the right or top of each datum-train. The value of the datum-parameter also can be determined approximately by reading the target log troland value on the abscissa at which t is about O.

J. M. Williams

Luminance and the Hess Effect

207

FIG. 16. Hess-effect t as a function of upper target log troland value; AB's mean response for 24 settings. Plotting symbols as in Figure 14 above, for lower-target log troland value (log El) as parameter.

J. M. Williams

Luminance and the Hess Effect

208

FIG. 17. Hess effect t as a function of lower target log troland value; AB's mean response for 24 settings. Parameter is log E. Plotting symbols not given in the legend may be found in Figure 14 above. These are many of the same data as in Figure l6 replotted in a different way. The log E = .69 symbols locate data obtained by linear interpolation between neighboring datumtrains many of which latter are not shown. Only those .69 symbols with a horizontal line bisecting them represent data gathered solely at a log E of .69.

J. M. Williams

Luminance and the Hess Effect

209

FIG. 18. Hess-effect t as a function of upper target log troland value; AB's mean response for 36 settings. Parameter is log El; plotting symbols as given in Figure 14 above.

J. M. Williams

Luminance and the Hess Effect

210

FIG. 19. Average Hess-effect t for AB and LD as a function of upper target log troland value. Parameter is log El for plotting symbols given in Figure 14 above. Each average is an unweighted mean of the appropriate two mean ts, one for LD's n = 24 sample and one for AB's n = 36 sample.

J. M. Williams

Luminance and the Hess Effect

211

FIG. 20. Comparison of Hess-effect and Pulfrich-effect tvalues for the same observer AB. Parameter is log troland value of the lower- -illuminated target; Hess-effect plotting symbols are as in Figure 14 above. The Pulfrich-effect data are from Figures 11 and 12 above, the Hess-effect data from Figure 18 above.

J. M. Williams

Luminance and the Hess Effect

212

FIG. 21. Across-studies comparison of various data on visual latency. Parameter value (log El) is given to the right of each datum-train--except for the open triangles, which are labelled on the left. The variable on the abscissa is log troland value of the target of upper (higher) luminance. The data for the Hess effect for AB and LD are as in Figure 19 above; the Pulfrich-effect data are from Figure l3 above. The Wilson and Anstis (1969) data were read from their Figure 2 (JW), with stimulus conditions converted from luminance to illuminance at the retina (troland value) using de Groot and Gebhard's (1952) formula for pupil size; the estimated pupil size in the Wilson and Anstis study was adjusted further for convergence miosis (which, at 45 cm viewing distance, is presumed to have reduced maximum pupil dilation at lower luminances) and for the small target size (which is presumed to have reduced maximum pupil contraction at highest luminances) by using the mean estimated (de Groot & Gebhard) pupil size over all levels of (higher) luminance studied. This overall mean was averaged individually with each separate de Groot and Gebhard pupil size at each level studied by Wilson and Anstis (1969), and the resultant pupil size was used to convert target luminance to troland value on the assumption that the target of lower luminance had no effect on the pupil.

J. M. Williams

Luminance and the Hess Effect

213

FIG. 22. Theoretical fit of the Hess-effect data plotted in Figure 19 above. Theoretical time constant a and persistence-factor k are the parameters defining the smooth curves, which latter were fit to the data by eye.

J. M. Williams

Luminance and the Hess Effect

214

FIG. 23. Comparison of the theoretical fits for Hess-effect and Pulfrich-effect data in the photopic range. The parameter for each datum-train is given in the legend at lower right, and theoretical parameters a and k of equation (35) above are shown for the curves fit by eye. The data were cascaded by the method of Alpern (1968), so that the variable on the abscissa is log E referenced to the respective datum-parameters given.

J. M. Williams

Luminance and the Hess Effect

215

FIG. 24. Low-level (scotopic and low-photopic) Hess-effeet data for JMW gathered by JMW as E and S. Plotting symbols (Figure 14 above) each represent a mean t (= s/v) for 48 settings. The variable on the abscissa is log troland value of the upper target; parameter is log troland value of the lower target. These data overlap with those of Figure 25 below.

J. M. Williams

Luminance and the Hess Effect

216

FIG. 25. Hess-effect data for JMW gathered by JMW as E and S. This figure omits some of the scotopic data near the top of Figure 24 above. Each plotting symbol (see Figure l4 above) represents a mean t for 48 settings. The variable on the abscissa is log troland value of the upper target; the parameter is log troland value of the lower target.

J. M. Williams

Luminance and the Hess Effect

217

FIG. 26. Validation of the Hess-effect data plotted in Figures 24 and 25 above. With a change in ordinate scale, data from Figures 24 and 25 have been replotted; the corresponding validation means are for samples of 24 settings each gathered by experimenter JV, as explained in the text. On the abscissa is log troland value of the upper target; the parameter is log troland value of the lower target.

J. M. Williams

Luminance and the Hess Effect

218

FIG. 27. Averaged Hess-effect data for theoretical fit. The two leftmost datum-trains are JMW data replotted from Figures 24 and 26 above on a reduced ordinate scale. Each circle of the second train from the left represents a mean of 72 settings obtained by averaging the validation data with the other data of Figure 26 above. The three rightmost datum-trains were obtained by equalweight averaging (mean) of the three available mean values of t for the three observers under each respective stimulus condition. Each circle plotted in the rightmost three trains represents from 48 to about l40 settings. The variable on the abscissa is log troland value of the upper target; parameter values (log troland value of lower target) are, from the leftmost datum-train to the rightmost, -2.91, -1.40, 0.27, 1.19, and 2.48.

J. M. Williams

Luminance and the Hess Effect

219

FIG. 28. Theoretical fit to averaged Hess-effect data for three observers. The solid smooth curves represent the theory proposed in the text (formula (35)). The interrupted curves are from Enright's (1970) formula for the Pulfrich effect, which formula, according to Table LII of the present work, fits the obtained data best in the least-squares sense. Time-constant a for the present theory was .154 s, an average computed from Table LII and used for all but the curve labelled a = 1.0. The Enright curves were generated with (average) parameter Y = 42.5 for the three left-most branches and Y = 25.9 for the two rightmost; the Enright parameter S was, for each branch from the left, 5.70, 5.70, -l5, -6.10, and -6.10, based on Table LII. The straight line sketched through the upper 5 data of the leftmost train is not theoretical but locates the apparent scotopic-photopic breakpoint for this train; this line's equation was obtained by (log) linear regression of data given in the leftmost column ("[20l]" to"[205]") of Table L. The variable on the abscissa is log troland value of the upper target; data are replotted from Figure 27.

J. M. Williams

Luminance and the Hess Effect

220

FIG. 29. Vernier acuity of observer AB for alignment of two illuminated targets on a black background, for targets moving laterally at 4/s. These were the Hess-effect targets of the present study. On the ordinate scale at left is sample standard deviation of the geometric t (= s/v) for the samples of size 24 to 48 available. The variable on the abscissa is log troland value of the lower target, and the parameter is difference in target log troland value. On the right are labels for the horizontal dotted lines locating values of the sample standard deviation of s in seconds of arc.

J. M. Williams

Luminance and the Hess Effect

221

FIG. 30. Vernier acuity of observer JMW for alignment of two illuminated targets on a black background, the targets moving laterally at 4/s. The ordinate scale at left gives the sample standard deviation of t for a sample size of 48. On the abscissa is log troland value of the lower target; datum parameters identified in the legend are for three different differences in target log troland value. Horizontal dotted lines, labelled at right, give values of the sample standard deviation of s in seconds of arc.

J. M. Williams

Luminance and the Hess Effect

222

FIG. 31. Comparison of Hess-effect and Pulfrich-effect latencies (t) for the same observer JMW. The parameter is log troland value of the lower-illuminated target, and the variable on the abscissa is log troland value of the upper(more highly) illuminated target, the illumination difference being between images in the eyes in the Pulfrich effect and between images in the same eye in the Hess effect. Parameter values for the Pulfrich-effect data are given to the right of the datum-train: the Hess-effect data are from Figure 25 above.

J. M. Williams

Luminance and the Hess Effect

223

FIG. 32. Intensity-dependent latency-differences predicted by reaction time compared with latency-differences inferred from the Hess effect. Average data for three observers. The lower panel gives Hess-effect ts replotted from Figure 27. The small dots in the lower panel are predicted using estimated RTs as explained in the text and given in Table VII. The differences between the lower-panel RTs and corresponding Hess- -effect ts are plotted in the upper panel; in the upper panel, the small open circles represent differences taken at log Eu values for the log E1 = 1.l9 data below. On the abscissas are log troland value of the upper target; the parameter is log troland value of the lower target.

J. M. Williams

Luminance and the Hess Effect

224

BIBLIOGRAPHY
Adrian, E. D., & Matthews, R. The action of light on the eye. Part I.

The discharge of impulses in the optic nerve and its relation to the electric changes in the retina. 414. (a) Adrian, E. D., & Matthews. R. The action of light on the eye. Part II. Journal of Physiology, 1927, 63, 378-

The processes involved in retinal excitation. 1927, 64. 279-301. (b) Allport, F. H. York:

Journal of Physiology,

Theories of perception and the concept of structure.

New

John Wiley & Sons, 1955. A note on visual latency. Psychological Review, 1968, 75,

Alpern, M. 260-264.

Ansbacher, H. L.

Distortion in the perception of real movement.

Journal

of Experimental Psychology, 1944, 34, 1-23. Arden, G. B. rat retina. Voltage gradients across the receptor layer of the isolated Journal of Physiology, 1976, 256, 333-360. Some properties of components of the cat Journal of

Arden, G. B., & Brown, K. T.

electroretinogram revealed by local recording under oil. Physiology, 1965, 176, 429-461. Arden, G. B., & Weale, R. A.

Variations of the latent period of vision.

Proceedings of the Royal Society (B Series), 1954, 142, 258-267. Arden, G. B., Bridges, C. D. B., Ikeda, H., & Siegel, I. M. generation of the early receptor potential. 3-24. Mode of

Vision Research, 1968, 8,

J. M. Williams

Luminance and the Hess Effect Transmission of visual signals to bipolar

225

Ashmore, J. F., & Falk, G.

cells near absolute threshold. Vision Research, 1979, 19, 419-423. Baker, K. E. Some variables influencing vernier acuity. II. Wave-length of illumination. I. Illumination Journal of the

and exposure time.

Optical Society of America, 1949, 39, 567 -576. Barlow, H. B. 4, 47-58. Barlow H. B., & Fatt, P. (Eds. ) Vertebrate photoreception. Academic Press, 1977. Barlow, H. B., Derrington, A. M., Harris, L. R., & Lennie, P. The effects New York: Dark-adaptation: A new hypothesis. Vision Resarch, 1964,

of remote retinal stimulation on the responses of cat retinal ganglion cells. Journal of Physiology, 1977, 269, 177-194. Component analysis of the Vision

Baron, W. S., Boynton, R. M., & Hammon, R. W.

foveal local electroretinogram elicited with sinusoidal flicker. Research, 1979, Bartlett, N. R. 19, 479-490.

A comparison of manual reaction times as measured by Psychological Record, 1963, 13, 51-56. Equipotentiality quantified: The anatomical

three sensitive indices. Bartlett, F., & John, E. R. distribution of the engram.

Science, 1973, 181, 764-767.

J. M. Williams

Luminance and the Hess Effect

226

Bartlett, N. R., & MacLeod, S. human reaction time. 44, 306-311. Bartley, S. H.

Effect of flash and field luminance upon

Journal of the Optical Society of America, 1954,

Relation of intensity and duration of brief retinal

stimulation by light to the electrical response of the optic cortex of the rabbit. American Journal of Physiology, 1934, 108, 397-408. Optic nerve response to retinal Proceedings of the Society for Experimental

Bartley, S. H., & Bishop, G. H. stimulation in the rabbit. Biology, 1940, 44, 39-41. Baylor, D. A., & Hodgkin, A. L. turtle photoreceptors.

Changes in time scale and sensitivity in

Journal of Physiology, 1974, 242, 729-758. The membrane current of single

Baylor, D. A., Lamb, T. D., & Yau, K. W. rod outer segments. von Bekesy, G.

Journal of Physiology, 1979, 288, 589-611.

The smallest time difference the eyes can detect with Proceedings of the National Academy of Sciences,

sweeping stimulation. 1969, 64, 142-147. Berlucchi, G.

Anatomical and physiologcal aspects of visual functions of Brain Research, 1972, 37, 371-392.

corpus callosum. Bernhard, C. G. pathway.

Contributions to the neurophysiology of the optic

Acta Physiologica Scandinavica, 1940, 1 (Supplement 1), 1-94.

J. M. Williams Berry, R. N.

Luminance and the Hess Effect Quantitative relations among vernier, real depth, and

227

stereoscopic depth acuities. 38, 708-721.

Journal of Experimental Psychology, 1948,

Berry, R. N., Riggs, L. A., & Duncan, C. P. depth discriminations to field brightness. Psychology, 1950, 40, 349-354. Bills, M. A.

The relation of vernier and Journal of Experimental

The lag of visual sensation in its relation to wave lengths Psychological Monographs, 1920, 28 (Whole

and intensity of light. #l27). l-101.

Bixler, E. O., Bartlett, N. R., & Lansing, R. W. reflex and stimulus intensity. 559-560. Blake, R. The visual system of the cat.

Latency of the blink

Perception and Psychophysics, 1967, 2,

Perception and Psychophysics,

1979, 26, 423-448. Blakemore, C. Binocular depth perception and the optic chiasm. Vision

Research, 1970, 10, 43-47. Boring, E. G. A history of experimental psychology (2nd ed.). New York:

Appleton-Century-Crofts, 1950. Bortoff, A. retina. Bouma, P. J. Localization of slow potential responses in the Necturus Vision Research, 1964, 4, 627-635. The flickering of sources of electric light. Philips

Technical Review, 1941, 6, 295-302. Bowen, R. W., Pola, J., & Matin, L. luminance, duration and energy. Visual persistence: Effects of flash

Vision Research, 1974, 14, 295-303.

J. M. Williams Bracewell, R.

Luminance and the Hess Effect The Fourier transform and its applications. New York:

228

McGraw-Hill, 1965. Brandenberg, W. M., & Neu, J. T. imperfectly diffuse surfaces. America, 1966, 56, 97-103. Brauner, J. D., & Lit, A. intensity discrimination. 105-114. Breitmeyer, B. G. Unmasking visual masking: A look at the "Why" behind The Pulfrich effect, simple reaction time, and American Journal of Psychology, 1976, 89, Unidirectional reflectance of Journal of the Optical Society of

the veil of the "How."

Psychological Review, 1980, 87, 52-69. Somatically

Brenner, D., Lipton, J., Kaufman, L., & Williamson, S. J. evoked magnetic fields of the human brain. Breton, M. E. Hue substitution:

Science, 1978, 199, 81-83. Vision

Wavelength latency effects.

Research, 1977, 17, 435-443. Briggs, G. E., & Haldane, J. B. S. action. A note on the kinetics of enzyme

Biochemical Journal, 1925, 19, 338-339. Central pathways of vision. (a) Annual Review of

Brindley, G. S.

Physiology, 1970, 32, 259-268. Brindley, G. S.

Physiology of the retina and the visual pathway. London: Edward

Monograph #6 of the Physiological Society. (2nd ed.) Arnold, 1970. (b)

Brindley, G. S., & Westheimer, G. electroretinogram.

The spatial properties of the human

Journal of Physiology, 1965, 179, 518-537.

J. M. Williams Broman, A .

Luminance and the Hess Effect Introduction to partial differential equations from Fourier Reading, MA:

229

series to boundary-value problems. Brown, B.

Addison-Wesley, 1970.

Resolution thresholds for moving targets at the fovea and in Vision Research, 1972, 12, 293-304. In C. H. Graham (Ed.), Vision and visual John Wiley, 1965, ch. 17. A new receptor potential of the monkey Nature, 1964, 201, 626-628. (a)

the peripheral retina. Brown, J. L. perception. Afterimages. New York:

Brown, K. T., & Murakami, M.

retina with no detectable latency. Brown, K. T., & Murakami, M.

Biphasic form of the early receptor Nature, 1964, 204, 739-740. (b)

potential of the monkey retina. Brown, W. R. J.

The effect of field size and chromatic surroundings on Journal of the Optical Society of America, 1952,

color discrimination. 42, 837-844.

Bunt, A. H., Minckler, D. S., & Johanson, G. W.

Demonstration of

bilateral projection of the central retina of the monkey with horseradish peroxidase 619-630. Burke, W., & Cole, A. M. geniculate nucleus. Extraretinal influences on the lateral Review of Physiology, Biochemistry, and neuronography. Journal of Comparative Neurology, 1977, 171,

Pharmacology, 1978, 80, 106-166. Burr, D. C. Acuity for apparent vernier offset. Vision Research, 1979,

19, 835-837.

J. M. Williams Burr, D. C., & Ross, J. depth?

Luminance and the Hess Effect

230

How does binocular delay give information about

Vision Research, 1979, 19, 523-532. The influence of intensity of the stimulus on the length Brain, 1885, 8, 512-515.

Cattell, J. M.

of reaction time. Chapman, J. M.

Spectral sensitivity of single neural units in the Journal of the Optical Society of America, 1961, 51,

bullfrog retina. 1102-1112. Charpentier, A. visuelles. Charpentier, A.

L'inertie rtinienne et la thorie des perceptions Archives d'Ophtalmologie, 1887, 6, 114-116. Dmonstration directe de la diffrence de temps perdu Archives de physiologie normale et pathologique,

suivant les couleurs.

1893, 5 (Series 5), 568-570. Cobb, W. A., & Dawson, G. D. The latency and form in man of the occipital Journal of Physiology, 1960, 152,

potentials evoked by bright flashes. 108-121. Coltheart, M.

Iconic memory and visible persistence.

Perception and

Psychophysics, 1980, 27, 183-228. Cone, R. A. The rat electroretinogram. I. Contrasting effects of Journal of

adaptation on the amplitude and latency of the b-wave. General Physiology, 1964, 47, 1089-1105. Cone, R. A. The rat electroretinogram. II.

Bloch's law and the latency

mechanism of the b-wave. 1116. Cone, R. A.

Journal of General Physiology, 1964, 47, 1107-

Early receptor potential of the vertebrate retina.

Nature,

1964, 204, 736-739.

J. M. Williams Cone, R. A.

Luminance and the Hess Effect The early receptor potential of the vertebrate eye. Cold

231

Spring Harbor Symposium on Quantitative Biology, 1965, 30, 483-491. Conner, J. D. The rod outer segment stores an adaptation signal. ARVO,

1980, p. lO3, Abstract #7.

(Announcement of the annual meeting of the

Association for Research in Vision and Ophthalmology at Orlando, Florida, in May, 1980). Conner, J. D., & MacLeod, D. I. A. Rod photoreceptors detect rapid

flicker. Science, 1977, 195, 698-699. Coren, S., & Ward, L. M. Levels of processing in visual illusions: The Journal

combination and interaction of distortion-producing mechanisms. of Experimental Psychology: 324-335. Creed, R. S., & Granit, R.

Human Perception and Psychophysics, 1979, 5,

Observations on the retinal action potential

with especial reference to the response to intermittent stimulation. Journal of Physiology, 1933, 78, 419-441. Crescitelli, F. The electroretinogram of the antelope ground squirrel.

Vision Research, 1961, 1, 139-153. Dodwell, P. C. Human pattern and object perception. In R. Held, H. W.

Leibcwitz, & H. L. Teuber (Eds.), Handbook of sensory physiology, vol. VIII: Perception. New York: Springer-Verlag, 1978, ch. 15. Optokinetic reactions in man elicited Vision Research, 1979, 19, 1105-

Dubois, M. F. W., & Collewijn, H.

by localized retinal motion stimuli. 1115.

J. M. Williams

Luminance and the Hess Effect System of ophthalmology. St. Louis: Mosby, 1961. vol. 2: The

232

Duke-Elder, S., & Wybar, K. C. anatomy of the visual system. Durup, G., & Fessard, A.

L'1ectroencphalogramme de l'homme.

Observations psycho-physiologiques relatives 1'action des stimuli visuals et auditifs. Durup, G., & Fessard, A. Annee Psychologique, 1935, 36, 1-32. L'1ectroencphalogramme de l'homme. Donnes

quantitatives sur 1'arrt provoqu par des stimuli visuals ou auditifs. Comptes Rendues de la Socit de Biologie, 1936, 122, 756-758. Dvorak, V. ber Analoga der persnlichen Differenz zwischen beiden Augen Sitzungsberichte der

und den Netzhautstellen desselben Auges.

Bhmischen Gesellschaft der Wissenschaften in Prag, 1872, Jan-Jun, 65-74. (Presented by Prof. Mach) Edmund, C. On the duration of the luminous impression. Acta

Ophthalmologica, 1928, 6, 414-424. Enright, J. T. Stereopsis, visual latency, and three-dimensional moving

pictures. American Scientist, 1970, 58, 536-545. van Essen, D. C. Visual areas of the mammalian cerebral cortex. Annual

Review of Neuroscience, 1979, 2, 227-263. Fiorentini, A. Mach band phenomena. In D. Jameson & L. M. Hurvich

(Eds.), Handbook of sensory physiology, vol. VII/4, Visual psychophysics. New York: Springer-Verlag, 1972, ch. 8.

J. M. Williams Foley-Fisher, J. A. light.

Luminance and the Hess Effect

233

Measurements of vernier acuity in white and coloured

Vision Research, 1968, 8, 1055-1065. The effect of target line length on vernier acuity in Vision Research, 1973, 13, 1447-1454. Diagnostic uses of

Foley-Fisher, J. A.

white and blue light.

Frisen, L., Hoyt, W. F., Bird, A. C., & Weale, R. A. the Pulfrich phenomenon. Lancet, 1973, 2, 385-386.

Fuortes, M. G. F., & Hodgkin, A. L.

Changes in the time scale and Journal of Physiology, 1964,

sensitivity in the ommatidia of Limulus. 172, 239-263. Goldstein, E. B., & Berson, E. L. early receptor potential. Gordon, B. functions.

Rod and cone contributions to the human

Vision Research, 1970, 10, 207-218. Structure, physiology and possible

Superior colliculus:

In C. C. Hunt (Ed.), Neurophysiology (vol. 3 of Physiology, Baltimore:

Series 1 of MTP International Review of Science). University Park Press, 1975, ch. 5. Gordon, J., Shapley, A. M., & Kaplan, E. and spectral mechanisms. 138. Gouras, P.

The eel retina: Receptor classes

Journal of General Physiology, 1978, 71, 123-

Rod and cone independence in the electroretinogram of the Journal of Physiology, 1966, 187,

dark- -adapted monkey's perifovea. 455-464. Goures, P., & Link, K. ganglion cells.

Rod and cone interaction in dark-adapted monkey

Journal of Physiology, 1966, 184, 499-510.

J. M. Williams Graham, C. H.

Luminance and the Hess Effect Visual perception. In S. S. Stevens (Ed.), Handbook of John Wiley & Sons, 1951, ch. 23. Temporal

234

experimental psychology.

New York:

van de Grind, W. A., Grusser, 0. J., & Lunkenheimer, H. U. transfer properties of the afferent visual system.

In R. Jung (Ed.),

Handbook of sensory physiology, vol. VII/3, Central processing of visual information A: Integrative functions and comparative data. New York:

Springer-Verlag, 1973, ch. 7. de Groot, S. G., & Gebhard, J. W. luminance. 495. Grossman, S. P. A textbook of physiological psychology. New York: John Pupil size as determined by adapting

Journal of the Optical Society of America, 1952, 42, 492-

Wiley & Sons, 1967. Grum, F., & Costa, L. F. Detemination of polarization in optical Applied Optics, 1974,

instruments and its metrological implications. l3, 2228-2232. Grusser, O. J.

Cat ganglion-cell receptive fields and the role of In F. 0. Schmitt & F. G. Worden Cambridge, MA: MIT

horizontal cells in their generation. (Eds.), The neurosciences: Press, 1979, ch. 15. Guth, S. L.

Fourth study program.

The effect of wavelength on visual perceptual latency.

Vision Research, 1964, 4, 567-578.

J. M. Williams

Luminance and the Hess Effect

235 In F. 0. Schmitt

Hagins, W. A. Excitation in vertebrate photoreceptors. I & F. G. Worden (Eds.), The neurosciences: Cambridge, MA: MIT Press, 1979, ch. ll. Physics, part II. (2nd ed.)

Fourth study program.

Halliday, D. & Resnik, R. Wiley & Sons, 1962. Harmon, L. D., & Lewis, K. 1966, 46, 513-591. Hartline, H. K.

New York: John

Neural modeling.

Physiological Reviews,

Intensity and duration in the excitation of single Journal of Cellular and Comparative Physiology,

photoreceptor units. 1934, 5, 229-247. Hartline. H. K.

The response of single optic nerve fibers of the In F. Ratliff (Ed.), New York:

vertebrate eye to illumination of the retina.

Studies on excitation and inhibition in the retina. Rockefeller University Press, 1974.

(Originally published in American

Journal of Physiology, 1938, 121, 400-415). Hebb, D. O. The innate organization of visual activity. III.

Discrimination of brightness after removal of the striate cortex in the rat. Journal of Comparative Psychology, 1938, 25, 427-437. The effect of early and late brain injury upon test scores, Proceedings of the

Hebb, D. 0.

and the nature of normal adult intelligence.

American Philosophical Society, 1942, 85, 275-292.

J. M. Williams Hecht, S.

Luminance and the Hess Effect The relation of time, intensity and wavelength in the

236

photosensory system of Pholas. 657-672. Heinemann, E. G.

Journal of General Physiology, 1928, 11,

Discriminability and ratio scaling.

In J. C.

Armington, John Krauskopf, and B. R. Wooten (Eds.), Visual psychophysics and physiology: A volume dedicated to Lorrin Riggs. New York:

Academic Press, 1978, ch. 13. Helson, H. Adaptation-level as a basis for a quantitative theory of Psychological Review, 1948, 55, 297-313. Delay in visual perception in Brain, 1974, 97,

frames of reference.

Heron, J. R., Regan, D., & Milner, B. A.

unilateral optic atrophy after retrobulbar neuritis. 69-78. Hess, C.

Untersuchungen ber den Erregungsvorgang im Sehorgan bei Kurz Pflger's Archiv fr die gesamte

und bei Lngerdauernder Reizung.

Physiologie des menschen und der tiere, 1904, 101, 226-262. Hobson, J. A., & Scheibel, A. B. The brainstem core: Sensorimotor

integration and behavioral state control. Program Bulletin, 1980, 18 (Whole #1).

Neurosciences Research

Hochstein, S., & Shapley, R. M. Y cat retinal ganglion cells. 284.

Linear and nonlinear spatial subunits in Journal of Physiology, 1976, 262, 265-

J. M. Williams

Luminance and the Hess Effect A quantitative description of membrane

237

Hodgkin, A. L., & Huxley, A. F.

current and its application to conduction and excitation in nerve. Journal of Physiology, 1952, 117, 500-544. van der Horst, G. J. C., & Bouman, M. A. phenomena in color vision. Hubbell, W. L., & Bownds, M. D. photoreceptors. On searching for Mach band-type

Vision Research, 1967, 7, 1027-1029. Visual transduction in invertebrate

Annual Review of Neuroscience, 1979, 2, 17-34. Brain mechanisms of vision. Scientific

Hubel, D. H., & Wiesel, T. N.

American, 1979, 241(3), September, 150-162, 250. Hyvarinen, J. CNS: Afferent mechanisms with emphasis on physiological and Annual Review of Physiology, 1973, 35, 243-

behavioral correlations. 272.

Ingling, C. R. Jr., & Drum, B. A.

Retinal receptive fields:

Correlations

between psychophysics and electrophysiology. 1151-1163. Ives, H. E. Visual diffusivity.

Vision Research, 1973, 13,

Philosophical Magazine, 1917, 33

(Series 6, #249), 18-33. Ives, H. E. The resolution of mixed colors by differential visual

diffusivity. Philosophical Magazine, 1918, 35 (Series 6, #25l), 413-421. Ives, H. E. A theory of intermittent vision. Journal of the Optical

Society of America, 1922, 6, 343-361.

J. M. Williams

Luminance and the Hess Effect

238

Ives, H. E., & Kingsbury, E. F.

The theory of the flicker photometer. I.

Philosophical Magazine, 1914, 28 (Nov.), 708-728. Ives, H. E., & Kingsbury, E. F. II. Unsymmetrical conditions. 290-321. Johnson, E. P., & Bartlett, N. R. Effect of stimulus duration on Journal of the Optical The theory of the flicker photometer. Philosophical Magazine, 1916, 29 (April),

electrical responses of the human retina. Society of America, 1956, 46, 167-170. Johnson, R. E., & Kiokemeister, F. L. Boston: Julesz, B. Allyn & Bacon, 1964. Global stereopsis:

Calculus, with analytic geometry.

Cooperative phenomena in stereoscopic

depth perception.

In R. Held, H. W. Leibowitz, & H. L. Teuber (Eds.), Perception. New York:

Handbook of sensory physiology, vol VIII: Springer-Verlag, 1978, ch. 7. Julesz, B., & White, B. phenomenon. Kaas, J. H. system. science.

Short term visual memory and the Pulfrich

Nature, 1969, 222, 639-641. Subdivisions and interconnections of the primate visual

In S. J. Cool & E. L. Smith III (Eds.), Frontiers in visual New York: Springer-Verlag, 1978, 557-563. Eccentricity of the scotopic fovea about a Vision Research, 1976,

Kandel, G. L., & Bedell, H. E.

photopic point as revealed by acuity contours. 16, 357-362.

J. M. Williams Kaneko, A.

Luminance and the Hess Effect Physiology of the retina. Annual Review of Neuroscience,

239

1979, 2, 169-191. Kaufman, J. E., & Cristensen, J. F. New York: Keller, M. IES Lighting Handbook (5th ed.).

Illuminating Engineering Society, 1972. The relation between the critical duration and intensity in Journal of Experimental Psychology, 1941,

brightness discrimination. 28, 407-418. Kelly, D. H.

Flickering patterns and lateral inhibition.

Journal of the

Optical Society of America, 1969, 59, 1361-1370. (a) Kelly, D. H. Diffusion model of linear flicker responses. (b) Uniform Journal of

the Optical Society of America, 1969, 59, 1665-1670. Kelly, D. H. fields. -546. Kelly, D. H., & Wilson, H. R. retinal diffusion. Kimble, G. A. New York:

Theory of flicker and transient responses, I.

Journal of the Optical Society of America, 1971, 61, 537

Human flicker sensitivity:

Two stages of

Science, 1978, 202, 896-899.

Hilgard and Marquis' conditioning and learning. (2nd ed.) Appleton-Century-Crofts, 1961. Adaptation properties of intracellularly recorded gekko In H. Langer (Ed.), Biochemistry and New York: Springer-Verlag, 1973, 219-

Kleinschmidt, J.

photoreceptor potentials.

physiology of visual pigments. 224.

J. M. Williams

Luminance and the Hess Effect

240

Kleitman, N., & Pieron, H.

Sur la vitesse d'tablissement de la sensation

lumineuse et la grandeur de 1'ondulation de prquilibre pour des excitations monochromatiques d'intensit variable. Comptes rendues des

sances de l'acadmie des sciences, 1925, 180, 393-396. Kong, K. L., & Wasserman, G. S. Two linear rules relate the latencies of Sensory Processes, 1978,

visual responses to their critical durations. 2, l-8. Krantz, D. H. Visual scaling.

In D. Jameson & L. M. Hurvich (Eds.), Visual psychophysics. New

Handbook of sensory physiology, vol. VII/4: York: Springer-Verlag, 1972, ch. 26.

Kratz, K. E., Webb, S. V., & Sherman, S. M.

Electrophysiological

classification of X- and Y-cells in the cat's lateral geniculate nucleus. Vision Research, 1978, 18, 1261-1264. Kruger, J. McCol1ough effect: A theory based on the anatomy of the

lateral geniculate body. 179. Kuffler, S. retina.

Perception and Psychophysics, 1979, 25, 169-

Discharge patterns and functional organization of mammalian Journal of Neurophysiology, 1953, 16, 37-68. A bioelectrical model for the human visual system:

Kumar, V. S. N.

Quantitative analysis using reaction time and evoked potential information. Unpublished doctoral dissertation, Southern Illinois

University at Carbondale, 1975.

J. M. Williams Landahl, H. D.

Luminance and the Hess Effect Miscellaneous applications of neural net theory to In N. Rashevsky (Ed.), Aspetti London:

241

psychological phenomena.

fisicomatematici della Biologia ("Enrico Fermi" XVI Corso). Academic Press, 1962, 295-302. Lennox, M. A. cat.

Geniculate and cortical responses to colored light flash in

Journal of Neurophysiology, 1956, l9, 271-279. what the

Lettvin, J. Y, Maturana, H. R., McCulloch, W. S., & Pitts, W. H. frog's eye tells the frog's brain.

Proceedings of the Institute of

Radio Engineering, 1959, 47, 1940-1951. Levinson, J. Z. Interpretation of generator potentials. In M. G. F. Physiology of

Fuortes (Ed.), Handbook of sensory physiology, vol. VII/2: photoreceptor organs. New York:

Springer-Verlag, 1972, ch. 8. The control of phosphodiesterase in rod

Liebman, P. A., & Pugh, E. N. Jr.

disc membranes--kinetics, possible mechanisms and significance for vision. Lit, A. Vision Research, 1979, 19, 375-380, 354. The magnitude of the Pulfrich stereophenomenon as a function of

binocular differences of intensity at various levels of illumination. American Journal of Psychology, 1949, 62, 159-181. Lit, A. Magnitude of the Pulfrich stereophenomenon as a function of Journal of the Optical Society of America, 1960, 50,

target thickness. 321-327 (a)

J. M. Williams Lit, A.

Luminance and the Hess Effect

242

The magnitude of the Pulfrich stereophenomenon as a function of Journal of Experimental Psychology, 1960, 59, 165-175.

target velocity. (b) Lit, A.

Effect of target velocity in a frontal plane on binocular spatial Journal of the

localization at photopic retinal illuminance levels. Optical Society of America, 1960, 50, 970-973. Lit, A. (c)

Equidistance settings at photopic retinal-illuminance levels as a Journal of the Optical

function of target velocity in a frontal plane. Society of America, 1964, 54, 83-88. Lit, A.

Depth-discrimination thresholds for targets with equal retinal American Journal of

illuminance oscillating in a frontal plane.

Optometry and Archives of the American Academy of Optometry, 1966, 43, 283-298. Lit, A. Illumination effects on depth discrimination. The Optometric

Weekly, 1968, 59, 42-55. Lit, A. Spatio-temporal aspects of binocular depth discrimination. In New

S. J. Cool & E. L. Smith III (Eds.), Frontiers in Visual Science. York: Springer-Verlag, 1978, 396-424.

Lit, A., & Finn, J. P.

Variability of depth-discrimination thresholds as Journal of the Optical Society of

a function of observation distance. America, 1976, 66, 740-742.

J. M. Williams Lit, A., & Hamm, H. D.

Luminance and the Hess Effect

243

Depth-discrimination thresholds for stationary and Journal

oscillating targets at various levels of retinal i1luminance. of the Optical Society of America, 1966, 56, 510-516. Lit, A., & Vicars, W. M. of

The effect of practice on the speed and accuracy American Journal of Psychology, 1966, 79,

equidistance-settings.

464-469. Lit, A., & Vicars, W. M. Stereoacuity for oscillating targets exposed Perception and

through apertures of various horizontal extents. Psychophysics, 1970, 8, 348-352. Lit, A., Finn, J. P., & Vicars, W. M.

Effect of target-background

contrast on binocular depth discrimination at photopic levels of illumination. Vision Research, 1972, 12, 1214-l-51. Simple reaction time as a function

Lit, A., Young, R. H., & Shaffer, M.

of luminance for various wavelengths. Perception and Psychophysics, 1971, 10, 397-399. Lockhart, R. D., Hamilton, G. F., & Fyfe, F. W. (2nd ed.) Ludvigh, E. 64. Lukas, F. X. J., Tulunay-Keesey, U., & Limb, J. O. Thresholds at Journal of the New York: J. B. Lippincott, 1969. Science, 1948, 108, 63Anatomy of the human body

The visibility of moving objects.

luminance edges under stabilized viewing conditions. Optical Society of America, 1980, 70, 418-422.

J. M. Williams Lythgoe, R. l41, 474.

Luminance and the Hess Effect Some observations on the rotating pendulum.

244 Nature, 1938,

McCann, J. J., Savoy, R. L., & Hall, J. A. Jr. frequency sine-wave targets: parameters. McDougall, W. the eye. McDougall, W.

Visibility of low-

Dependence on number of cycles and surround

Vision Research, 1978, 18, 891-894. The sensations excited by a single momentary stimulation British Journal of Psychology, 1904, 1, 78-115. The illusion of the "fluttering heart" and the visual British Journal of Psychology,

functions of the rods of the retina. 1905, 1, 428-434. McIlwain, J. T.

Receptive fields of optic tract axons and lateral Peripheral extent and barbiturate sensitivity.

geniculate cells:

Journal of Neurophysiology, 1964, 27, 1154-1173. MacLeod, D. I. A. Visual sensitivity. Annual Review of Psychology,

1978, 29, 613-645. Magoon, E. H., & Robb, R. M. and tract. Development of myelin in human optic nerve (Announcement of the annual

ARVO, 1980, p. 3, Abstract #l0.

meeting of the Association for Research in Vision and Ophthalmology at Orlando, Florida, in May, 1980). Mansfield, R. J. W. Latency functions in human vision. Vision Research,

1973, l3, 2219-2234. Mansfield, R. J. W., & Daugman, J. G. latency. Retinal mechanisms of visual

Vision Research, 1978, 18, 1247-1260.

J. M. Williams Matin, L.

Luminance and the Hess Effect Critical duration, the differential luminance threshold, A theoretica1

245

critical flicker frequency and visual adaptation: treatment. May, M. J.

Journal of the Optical Society of America, 1968, 58, 404-415. A new method for studying visual latency. Vision Research,

l964, 4, 515-516. Memendez, A. Simple visual reaction time as a function of target-flash

luminance at various target-annulus border separations, annulus areas, and annulus configurations. Unpublished master's thesis, Southern

Illinois University at Carbondale, 1979. Michaelis, R. R. (Ed.) (Text ed.) New York: Funk and Wagnall's Standard College Dictionary. Harcourt, Brace & World, 1963. Binocular depth perception and the

Mitchell, D. E. , & Blakemore, C. corpus callosum.

Vision Research, 1970, 10, 49-54. Accuracy of stereoscopic localization of

Mitchell, D. E., & O'Hagen, S.

small line segments that differ in size or orientation for the two eyes. Vision Research, 1972, 12, 437-454. Monnier, M. Retinal time, retinocortical time, alpha blocking time and Electroencephalography and Clinical

motor reaction time.

Neurophysiology, 1949, 1, 515 -517. Monnier, M. in man. Retinal, cortical, and motor responses to photic stimulation Retino-cortical time and optomotor integration time. Journal

of Neurophysiology, 1952, 15, 469-486.

J. M. Williams

Luminance and the Hess Effect Changes in critical

246

Montellese, S., Sharpe, L. T., & Brown, J. L. duration during dark-adaptation. Moreland, J. D.

Vision Research, 1979, 19, 1147-1153. Vision

Retinal topography and the blue-arcs phenomenon.

Research, 1969, 9, 965-976. Morgan, M. J. Differential visual persistence between the two eyes: Journal of Experimental A

model for the Fertsch-Pulfrich effect. Psychology:

Human Perception and Performance, 1977, 3, 484-495 S-potentials from luminosity-units in Journal of Physiology, 1966, 185, 587-

Naka, K. I., & Rushton, W. A. H. the retina of fish (cyprinidae). 599. Nissen, M. J., & Pokorny, J.

Wavelength effects on simple reaction time.

Perception and Psychophysics, 1977, 22, 457-462. Norren, D. V., & Vos, J. J. media. Spectral transmission of the human ocular

Vision Research, 1974, 14, 1237-1244. Studies of the optic nerve of the rhesus Vision

Ggden, T. E., & Miller, R. F. monkey:

Nerve fiber spectrum and physiological properties.

Research, 1966, 6, 485-506. Pankeve , J . I. 1975. Payne, B. R., Elberger, A. J., Berman, N., & Murphy, E. H. Binocularity Optical processes in semiconductors. New York: Dover,

in the cat visual cortex is reduced by sectioning the corpus callosum. Science, 1980, 207, 1097-1099.

J. M. Williams

Luminance and the Hess Effect

247

Penn, R. D., & Hagins, W. A.

Signal transmission along retinal rods and Nature, 1969, 223, 201-

the origin of the electroretinographic a-wave. 204. Penn, R. D., & Hagins, W. A. rods.

Kinetics of the photocurrent of retinal

Biophysical Journal, 1972, 12, 1073-1094. Delayed responses of ganglion cells in the

Pickering, S. G., & Varju, D. frog retina: deiay time. Pieron, H.

The influence of stimulus parameters upon the length of the Vision Research, 1969, 9, 865-879.

Loi de la vitesse d'tablissement des processus chromatiques

fondamentaux en fonction de l'intensit de l'excitation lumineuse. Comptes rendus des sances de l'acadmie des sciences, 1922, 174, 12941296. Pieron, H. De la variation des intervalles limites de masquage d'une

excitation lumineuse par une excitation conscutive trs intense en fonction de l'intensit de la premire. Biologie, 1923, 88, 736-739. Pieron, H. (a) Comptes rendus de la socite de

Du retard rductible de franchissement des synapses dans la

propagation de l'excitation lumineuse de la rtine l'corce crbrale. Comptes rendus des sances de l'acadmie des sciences, 1923, 176, 711714. (b)

J. M. Williams Plateau, J.

Luminance and the Hess Effect

248

Bibliographie analytique des principaux phnomnes subjectifs Premire section. Persistance des impressions sur la

de la vision. rtine.

Mmoires de l'Acadmie Royale des sciences, des lettres et des

beaux-arts de Belgique, 1878, 42, i-iv, 1-59. Poffenherger, A. T. Reaction time to retinal stimulation with special

reference to the time lost in conduction through nerve centers. Archives of Psychology, 1912 (Whole #23). Poggio, G. F. Central neural mechanisms in vision. In Mountcastle, V. C. V. Mosby, 1974,

B. (Ed.), Medical physiology (13th ed.). ch. 16. Polissar, M. J., & Rapaport, E. human heart.

St. Louis:

A mathematical pulsatile model of the In N.

Application to analysis of indicator curves.

Rashevsky (Ed.), Aspetti fisicomatematici della Biologia ("Enrico Fermi" XVI Corso). Pollack, J. D. luminances. London: Academic Press, 1962, 354-449.

Reaction time to different wavelengths at various Perception and Psychophysics, 1968, 3, 17-24. Neuronal mechanisms in

Poppel, E., Held, R., & Dowling, J. E. (Eds.) visual perception. Whole #3. Prestrude, A. M. illuminance.

Neurosciences Research Program Bulletin, 1977, 15,

Visual latencies at photopic levels of retinal Vision Research, 1971, 11, 351-361.

J. M. Williams

Luminance and the Hess Effect New method of measuring visua1-perceptual

249

Prestrude, A., & Baker, H. latency differences.

Perception and Psychophysics, 1968, 4, 152-154. Light adaptation and visual latency.

Prestrude, A. M., & Baker, H. D.

Vision Research, 1971, 11, 363-369 Pulfrich, C. Photometrie. Die stereoskopie im Dienste der isochromen und heterochromen Die Naturwissenschaften, 1922, 10, 553-564. 559-574.

[also posted at the SIU Pulfrich web site]. Rashevsky. N. of biology. Ratliff, F. retina. Ratliff, F. York: Mathematical biophysics: Chicago: Mach bands: San Francisco: Physicomathematical foundations

University of Chicago Press, 1938. Quantitative studies on neural networks in the Holden-Day, 1965. New

Studies on excitation and inhibition in the retina.

Rockefeller University Press, 1974. Physical metallurgy principles. (2nd ed.) New York:

Reed-Hill, R. E.

Van Nostrand, 1973. Regan, D. Recent advances in electrical recording from the human brain.

Nature, 1975, 253, 401-407. Rietveld, W. J., & Tordoir, W. E. M. The influence of flash intensity Acta Physiologica

upon the visual evoked response in the human cortex. et Pharmocologica Neerlandse, 1965, 13, 160-170.

J. M. Williams Riggs, L. A.

Luminance and the Hess Effect Light as a stimulus for vision. New York: In C. H. Graham (Ed.),

250

Vision and visual perception. 1. Riggs, L. A., & Wooten, B. R. on human vision.

John Wiley & Sons, 1965, ch.

Electrical measures and psychophysical data

In D. Jameson & L. M. Hurvich (Eds.), Handbook of Visual psychophysics. New York:

sensory physiology, vol. VII/4: Springer-Verlag, 1972, ch. 27. Robinson, D. A. 1219-1224. Rodieck, R. W. 1973. Rodieck , R. W. 2, 193-225. Rodieck, R. W., & Rushton, W. A. H. Visual pathways.

Eye movement control in primates.

Science, 1968, 161,

The vertebrate retina.

San Francisco:

W. H. Freeman,

Annual Review of Neuroscience, 1979,

Cancellation of rod signals by cones, Journal of Physiology,

and cone signals by rods in the cat retina. 1976, 254,775-785. Rodieck, R. W., & Stone, J. moving visual patterns.

Response of cat retinal ganglion cells to Journal of Neurophysiology, 1965, 28, 819-832. Intensity versus adaptation and the Yision Reseach, 1972, 12, 909-928. Eye movements and the Pulfrich

Rogers, B. J., & Anstis, S. M. Pulfrich stereophenomenon.

Rogers, B. J., Steinbach, M. J., & Ono, H. phenomenon.

Vision Research, 1974, 14, 181-185.

J. M. Williams Roufs, J. A. J.

Luminance and the Hess Effect Perception lag as a function of stimulus luminance.

251

Vision Research, 1963, 3, 81-91. Roufs, J. A. J. Dynamic properties of vision--I. Experimental Vision Research,

relationships between flicker and flash thresholds. 1972, 12, 261-278. Roufs, J. A. J. Dynamic properties of vision--V.

Perception lag and

reaction time in relation to flicker and flash thresholds. Vision Research, 1974, 14, 853-869. Rushton, D. Use of the Pulfrich pendulum for detecting abnormal delay in Brain, 1975, 98, 283-296.

the visual pathway in multiple sclerosis. Sakitt, B. & Long, G. M.

Cones determine subjective offset of a stimulus Vision Research, 1979, 19, 1439-

but rods determine total persistence. 1441. Sanford, E. C. Personal equation.

American Journal of Psychology, 1889,

2, 3-38, 271-298, 403-430. Schiller, P. H. The primate superior colliculus and its sensory inputs.

In S. J. Cool & E. L. Smith III (Eds.), Frontiers in visual science. New York: Scott, A. C. Springer-Verlag, 1978, 437-448. The electrophysics of a nerve fiber. Reviews of Modern

Physics, 1975, 47, 487-533. Shlaer, S. The relation between visual acuity and illumination. Journal

of General Physiology, 1938, 21, 165-188.

J. M. Williams Singer, W.

Luminance and the Hess Effect

252

Control of thalamic transmission by corticofugal and ascending Physiological Reviews, 1977,

reticular pathways in the visual system. 57, 386-420. Singer, W., Tretter, F., & Cynader, M. cortex:

Organization of cat striate

A correlation of receptive-field properties with afferent and Journal of Neurophysiology, 1975, 38, 1080-1098. Model for visual luminance discrimination

efferent connections.

Sperling, G., & Sondhi, M. M. and flicker detection. 58, 1133-1145. Stevens, J. C., & Marks, L. E.

Journal of the Optical Society of America, 1968,

Cross-modality matching functions Perception and Psychophysics, 1980,

generated by magnitude estimation. 27, 379-389.

Stone, J., Leicester, J., & Sherman, S. M. the monkey retina. 348. Stroobant, P.

The naso-temporal division of

Journal of Comparative Neurology, 1973, 150, 333-

Recherches exprimentales sur l'equation personelle dans Comptes rendus des sances de l'acadmie

les observations de passage.

des sciences, 1891, 113, 457-460. Sullivan, G. D., Oatley, K., & Sutherland, N. S. affected by target length and separation. 1972, 12, 438-444. Vernier acuity as

Perception and Psychophysics,

J. M. Williams

Luminance and the Hess Effect

253

Sumitomo, I., Ide, K., & Iwama, K. fibers.

Conduction velocity of rat optic nerve

Brain Research, 1969, 12, 261-264. On the glial-cell origin of (Announcement of the

Szamier, R. B., Ripps, H., & Chappell, R. L. the ERG b-wave.

ARVO, 1980, p. 39, Abstract #1.

annual meeting of the Association for Research in Vision and Ophthalmology at Orlando, Florida, in May, 1980). Szentagothai, J. The elementary vestibulo-ocular reflex arc. Journal of

Neurophysiology, 1950, 13, 395-407. Szili, A. Flatternden Herzen. Zeitschrift fur Psychologie und

Physiologie der Sinnesorgane, 1892, 3, 359-387. Teichner, W. H., & Krebs, M. J. Laws of the simple visual reaction time.

Psychological Review, 1972, 79, 344-358. Tepas, D. I., & Armington, J. C. Properties of evoked visual potentials.

Vision Research, 1962, 2, 449-461. Tranchina, D., Gordon, J., Shapley, R. M., & Toyoda, J. outer plexiform layer in the turtle retina. Abstract #9. Linearity of the

ARVO, 1980, p. 103,

(Announcement of the annual meeting of the Association for

Research in Vision and Ophthalmology). Tretter, F., Cynader, M., & Singer, W. or secondary visual area? 1113. Troelstra, A. Intraocular noise: Origin and characteristics. Vision Cat parastriate cortex: A primary

Journal of Neurophysiology, 1975, 38, 1099-

Research, 1972, 12, 1313-1326.

J. M. Williams

Luminance and the Hess Effect A analysis of the b-wave in the

254

Troelstra, A., & Schweitzer, N. M. J. human ERG.

Vision Research, 1963, 3, 213-226. Invariance of the Vision

van der Tweel, L. H., Estevez, O., & Cavonius. C. R.

contrast evoked potential with changes in retinal illuminance. Research, 1979, 19, Vaughan, H. G. Jr. 1283-1287.

The analysis of scalp-recorded brain potentials.

In

R. F. Thompson & M. M. Patterson (Eds.), Bioelectric Recording Techniques, Part B. New York: Academic Press, 1974, ch. 4. The functional relation of Vision

Vaughan, H. G. Jr., Costa, L. D., & Gilden, L.

visual evoked response and reaction time to stimulus intensity. Research, 1966, 6, 645-656. Veringa, F. On the mechanisms underlying de Lange's results.

Koninklijke Nederlandse Akademie van Wetenschappen, 1961, 64, 413-416. Veringa, F. J. 998-999. Veringa, F. Diffusion model of linear flicker response. Journal of the Phase shifts in the human retina. Nature, 1963, 197,

Optical Society of America, 1970, 60, 285-286. Vicars, W. M., & Lit, A. Reaction time to incremental and decremental Vision

target luminance changes at various photopic background levels. Research, 1975, 15, 261-265.

J. M. Williams de Voe, R. D.

Luminance and the Hess Effect A nonlinear model for transient responses from lightJournal of General Physiology, 1967, 50,

255

adapted Wolf spider eyes. 1993-2030. Vogel, D. A., & Green, D. G.

Potassium release and b-wave generation: ARVO, 1980, p. 39, Abstract #2.

test of the Mller-cell hypothesis.

(Announcement of the annual meeting of the Association for Research in Vision and Ophthalmology at Orlando, Florida, in May, 1980). Vogelsang, K. Empfindungen. Die Empfindungszeit und der zeitliche Verlauf der Ergebnisse der Physiologie, 1928, 26, 122-184. Gray's Anatomy. (35th British ed.)

Warwick, R., & Williams, P. L. Philadelphia: Neale, R. A.

W. B. Saunders, 1973. Ophthalmologica, 1954,

Theory of the Pulfrich effect.

Separate #l28(6), 380-388. Weaver, K. S. A provisional standard observer for low level photometry.

Journal of the Optical Society of America, 1949, 39, 278-291. Webster, A. G. (Znd ed.) Partial differential equations of mathematical physics. New York: Dover, 1955. Science, 1972,

Weingarten, F. S. 176, 692 -694. Weisstein, N. metacontrast.

Wavelength effect on visual latency.

A Rashevsky-Landahl neural net:

Simulation of

Psychological Review, 1968, 75, 494-521.

J. M. Williams Westheimer, G. vision.

Luminance and the Hess Effect Spatial interaction in the human retina during scotopic

256

Journal of Physiology, 1965, 181, 881-894. Spatial interaction in human cone vision. Journal of

Westheimer, G.

Physiology, 1967, 190, 139-154. Westheimer, G. The spatial sense of the eye. Investigative

Ophthalmology and Visual Science, 1979, 18, 893-912. Williams. J. M. Distortions of vision and pain: Two functional facets of

d-lysergic acid diethylamide. 499-528. Wilson, J. A., & Anstis. S. M.

Perceptual and Motor Skills, 1979, 49,

Visual delay as a function of luminance.

American Journal of Psychology, 1969, 82, 350-358. Wist, E. R., Brandt, T., Diener, H. C., & Dichgans, J. effect on the Pulfrich stereophenomenon. 391-397. Wolf, C. Recherches sur 1'quation personnelle dans les observations de sa dtermination absolue, ses lois et son origine. Comptes Spatial frequency

Vision Research, 1977, 17,

passages,

rendus de l'acadmie des sciences, 1865, 60, 1268-1272. SIU Pulfrich web site] Wolken, J. J. York: Invertebrate photoreceptors:

[Posted at the

A comparative analysis.

New

Academic Press, 1971. Experimental psychology. (rev ed.)

Woodworth, R. S., & Scblosberg, H. Chicago:

Holt, Rinehart & Winston, 1954.

J. M. Williams

Luminance and the Hess Effect Visual-motor function of the primate

257

Wurtz, R. H., & Albano, J. E. superior colliculus.

Annual Review of Neuroscience, 1980, 3, 189-226. Co1or science. New York: John Wiley &

Wyszecki, G., & Stiles, W. S. Sons, 1967. Young, R. H., & Lit, A.

Stereoscopic acuity for photometrically matched Perception and

background wavelengths at scotopic and photopic levels. Psychophysics, 1972, 11, 213-216. Zierler, K. L.

Mechanism of muscle contraction and its energetics. Medical physiology. (13th ed.)

In C.

V. B. Mountcastle (Ed.), V. Mosby, 1974, ch. 3. Zuckerman, R. darkness.

St. Louis:

Mechanisms of photoreceptor current generation in light and Nature New Biology, 1971, 234, 29-31.

J. M. Williams

Luminance and the Hess Effect

258

APPENDIX I STUDIES OF VISUAL LATENCY AS A FUNCTION OF STIMULUS INTENSITY (PART A:


STUDY # 1a 1b 2a AUTHOR(S) & YEAR Bartley (1934) " Adrian & Matthews (1927a) " " " " Bartley & Bishop (1940) Hecht (1928) Roufs (1963) Adrian & Matthews (1927a) "

IDENTIFYING INFORMATION)
SOURCE FOR PART B DATA BELOW Fig. 1 Fig. 2 Fig. 9 (latent periods) Fig. 9 (peak freqs.) Fig. 10 (curve a) Table I ("On") Table I ("Off") Fig. 2 Fig. 2 450 nm Fig. 8 Fig. 19D Fig. 19B

ORGANISM

PARADIGM/PREPARATION Flashes of various durations; light anaesthetic " Prolonged flashes in vivo

Rabbit " Eel

2b 2c 2d 2e 3 4 5 2f 2g

" " " " Rabbit Pholas Human Eel Frog

" " " " Decorticated (frontal); Recorded ERG & N. II RT Simple RT; 2-flash order; eye-and-ear Flashes in vivo; compare retinal & N. II latency "

(Continued)

J. M. Williams

Luminance and the Hess Effect

259

APPENDIX IA (Continued) STUDIES OF VISUAL LATENCY AS A FUNCTION OF STIMULUS INTENSITY


STUDY # 6a 7a 8 9 6b 10a 10b 11a 12a 12b 13a 14 15a 15b AUTHOR(S) & YEAR Crescitelli (1961) Rogers & Anstis (1972) Bartlett (1963) Bixler, et al (1967) Crescitelli (1961) Bernhard (1940) " Ashmore & Falk (1979) Prestrude & Baker (1971) " Lit (1949) Troelstra & Schweitzer (1963) Cattell (1885) " ORGANISM Ground Squirrel Human Human Human Ground Squirrel Human " Shark Human " " " " " PARADIGM/PREPARATION ("all-cone" retina) Pulfrich effect Simple RT; switch close, muscle strain gauge, EMG Eyeblink: mechanical & EMG SOURCE FOR PART B DATA BELOW Fig. 3, #1 to #8 Fig. 4 Table 1 Fig. 1A, S1 EMG Fig. 2B Fig. 20 & Table 2 Table 3

("all-cone" retina) Alpha-block correlated with ERG RT or bicep EMG correlated with alpha block

Fig. 1 ("all-rod" retina) eyecup; flashes near (hyperpol. absolute threshold horizntl) Binocular nonstereo retinal disparity " Pulfrich effect ERG Discriminative RT Simple RT Fig. 3, 2.8 Fig. 3, 1.4 Fig. 5, CGM 1.98 Fig. 3A ranges of means "

(Continued)

J. M. Williams

Luminance and the Hess Effect

260

APPENDIX IA (Continued) STUDIES OF VISUAL LATENCY AS A FUNCTION OF STIMULUS INTENSITY


STUDY # 15c 16 17a 18 19 13b 20a 21 15d 22a 7b 23 24 25 AUTHOR(S) & YEAR Cattell (1885) Kleinschmidt (1973) Teichner & Krebs (1972) Baylor et al (1979) Penn & Hagins (1972) Lit (1949) Chapman (1961) Kuffler (1953) Cattell (1885) Cone (1964a) Rogers & Anstis (1972) Alpern (1968) Cone (1964b) Johnson & Bartlett (1956) ORGANISM Human Gekko Human Toad Rat Human Bullfrog Cat Human Albino rat Human Human Rat Human PARADIGM/PREPARATION 2-alternative FC RT ("all-rod" retina) Dark-adapted eyecup; single photoreceptor peak response Grossberg (1968) data on RT (low photopic/scotopic) Time-to-peak of single rod OS in pipette Dark-adapted retinal fragments; recorded across receptor layer Pulfrich effect Single-units in vivo; light & dark adapted On-cell; light-adapted Simple RT; all data Large-field flash; dark-adapted 12 hours; nembutol Pulfrich effect All transformed Lit (1949) data Diffuse 3-ms flash Dark-adapted; 7.5o flash of various exposure duration SOURCE FOR PART B DATA BELOW ranges of means Fig. 1A Fig. 1 Fig. 3A Fig. 10 Fig. 5, CMG, .98 Figs. 2 & 3 Fig. 9 & p. 56 ranges of means Fig. 3 Fig. 4 Fig. 2 Fig. 2 Fig. 8 implied times

(Continued)

J. M. Williams

Luminance and the Hess Effect

261

APPENDIX IA (Continued) STUDIES OF VISUAL LATENCY AS A FUNCTION OF STIMULUS INTENSITY


STUDY # 6c 26 17b 27 28 29 30 11b 31 32a 33 34 35 32b AUTHOR(S) & YEAR Crescitelli (1961) Creed & Granit (1933) Teichner & Krebs (1972) Arden & Weale (1954b) Durup & Fessard (1935) Durup & Fessard (1936) Prestrude (1971) Ashmore & Falk (1979) May (1964) Brauner & Lit (1976) Pickering & Varju (1969) Baylor & Hodgkin (1974) Roufs (1974) Brauner & Lit (1976) ORGANISM Ground Squirrel Cat Human " " " " Shark Human " Frog Turtle Human " PARADIGM/PREPARATION ("all-cone" retina) Decerebrate; dark-adapted & pupil dilated Raab, et al (1961) data; RT in photopic range 2-alternative FC temporal order for 2.7' diameter flashes VEP for 37o field 10 ms exposure duration; n = 20 trials Binocular nonstereo disparity (cascaded data) ("all-rod" retina) eyecup; flashes near absolute threshold Flash between 2 clicks Pulfrich effect (transformed) Retinal single-unit on-off; in vivo "early" response Cone I/I for constant I and I varied Same as Roufs (1963) but over wider stimulus ranges Simple RT SOURCE FOR PART B DATA BELOW Fig. 2A Fig. 1, curve c. Fig. 1 Fig. 2 extrafov. Fig. 8 (S2) Table Fig. 6 Fig. 1 Depolzing Bipolar Fig. 1 (S1) Fig. 2 Fig. 7 Fig. 4, curve 2 Fig. 2 combined Fig. 2

(Continued)

J. M. Williams

Luminance and the Hess Effect

262

APPENDIX IA (Continued) STUDIES OF VISUAL LATENCY AS A FUNCTION OF STIMULUS INTENSITY


STUDY # AUTHOR(S) & YEAR ORGANISM Human " Frog Toad Human Albino Rat Human Toad Human " PARADIGM/PREPARATION VEP (cII) to checkerboard pattern; manual/simple RT Alpha-block to 1-second flash in dark Dark-adapted; white & color flash Eyecup ERG; 10 ms diffuse flash at 502 nm Black-on-white Hess effect; transformed data Large-field flash; dark-adapted for 12 hours; Nembutol Simple RT to 1o foveal target Eyecup ERG; 10 ms diffuse flash at 502 nm Mean (2 Ss) median RT & VER VEP (cII) to checkerboard pattern; manual simple RT SOURCE FOR PART B DATA BELOW Fig 4 & p. 1286 Fig. 20 Fig. 22 Fig. 7a Fig. 2B Fig. 3 Fig. 4, control Fig. 7a Table 1,4o field Fig. 4; p. 1286 Figs. 2 & 3 Fig. 2 ( inverted) Fig. 3

36a & van der Tweel, 36b et al (1979) 10c 10d 37a 38 22b & 22c 39 37b 40 36c 20b 41 42 Bernhard (1940) " Mansfield & Daugman (1978) Wilson & Anstis (1969) Cone (1964a) Vicars & Lit (1975) Mansfield & Daugman (1978) Vaughan, et al (1966) van der Tweel, et al (1979)

Chapman (1961) Bullfrog Light/dark adapted in vivo single units Gouras & Link (1966) Arden & Brown (1965) Rhesus Monkey Cat Dark-adapted large ganglion cells in vivo; perifoveal 10 ms colored flash of 83o visual angle Dark-adapted local ERG; barbiturate anaesthesia

(Continued)

J. M. Williams

Luminance and the Hess Effect

263

APPENDIX IA (Concluded) STUDIES OF VISUAL LATENCY AS A FUNCTION OF STIMULUS INTENSITY


STUDY # 43 44 45 AUTHOR(S) & YEAR Lit, Young, & Shaffer (1971) Hartline (1938) Tepas & Armington (1962) ORGANISM Human Frog Human PARADIGM/PREPARATION Simple RT; scotopic and photopic Single-unit, optic nerve Occipital scalp VEPs, n = 2 Ss SOURCE FOR PART B DATA BELOW Fig. 1 (mean) Fig. 3 ("On") Fig. 8 ranges of means

J. M. Williams

Luminance and the Hess Effect

264

APPENDIX I STUDIES OF VISUAL LATENCY AS A FUNCTION OF STIMULUS INTENSITY (PART B:


LOG-UNITS STIMULUS STUDY RANGE # USED (Decadic) 1a 1b 2a 2b 2c 2d 2e 3 4 5 2f 2g 6a 7a 0.8 1.1 1.2 1.2 1.2 1.47 1.47 1.6 1.6 1.75 1.8 1.83 1.86 1.9 1.35 75 64 57 (t) 120 100 83 67

DEPENDENCE OF LATENCIES ON INTENSITY)


ratio

RANGE (or 3.5 x sample s. d.) OF LATENCIES OBTAINED (ms) ERG (local or gross) a-wave b-wave NONRETINAL CNS (N. II except as noted otherwise) 5.7 (Area 17 VEP) 10.5 (Area 17 VEP) 125 140 260 133 100 50 2093 (UCR) 45 (t) BEHAVIORAL CNS

ERG2 CNS2 *note below

0.46 0.45

1.27 0.56

*Note: ERG2/CNS2 is an estimated variance ratio discussed in the text and equals (ERG range)2/(CNS range)2. (Continued)

J. M. Williams

Luminance and the Hess Effect

265

APPENDIX IB (Continued) STUDIES OF VISUAL LATENCY AS A FUNCTION OF STIMULUS INTENSITY


LOG-UNITS STIMULUS STUDY RANGE # USED (Decadic) 8 9 6b 10a 10b 11a 12a 12b 13a 14 15a 15b 15e 16 17a 18 19 13b 20a 1.9 1.9 2.0 2.0 2.0 2.02 2.1 2.1 2.4 2.5 2.5 2.5 2.5 2.5 2.7 2.7 2.75 2.8 3 73 663 (photoreceptor) 23 (tN) 43 (photoreceptor layer t) 540 (photoreceptor) 220 57 139.5 99 122.5 306 16 (t) 31 (t) 12 (tF) 20 105 63 80 (alpha block) about 80 (alpha block) about 80 0.62 RANGE (or 3.5 x sample s. d.) OF LATENCIES OBTAINED (ms) ERG (local or gross) a-wave b-wave NONRETINAL CNS (N. II except as noted otherwise) BEHAVIORAL CNS 10 43
ratio

ERG2 CNS2

(Continued)

J. M. Williams

Luminance and the Hess Effect

266

APPENDIX IB (Continued) STUDIES OF VISUAL LATENCY AS A FUNCTION OF STIMULUS INTENSITY


LOG-UNITS STIMULUS STUDY RANGE # USED (Decadic) 21 15d 22a 7b 23 24 25 6c 26 17b 27 28 29 30 11b 31 32a 33 34 35 3.0 3.0 3.4 3.5 3.5 3.53 3.75 3.9 4.0 4.0 4.0 4.0 4.0 4.3 4.44 4.5 4.5 4.53 4.8 4.8 115.6 62 (photoreceptor) 117 178 86 (t) 49 (t) 158 (alpha-block) 189 (alpha-block) 35.6 (t) 54.2 65.2 60 55 77 95 (t) 51 25.6 (t) 50 (t) RANGE (or 3.5 x sample s. d.) OF LATENCIES OBTAINED (ms) ERG (local or gross) a-wave b-wave NONRETINAL CNS (N. II except as noted otherwise) 78 110.5 BEHAVIORAL CNS
ratio

ERG2 CNS2

(Continued)

J. M. Williams

Luminance and the Hess Effect

267

APPENDIX IB (Continued) STUDIES OF VISUAL LATENCY AS A FUNCTION OF STIMULUS INTENSITY


LOG-UNITS STIMULUS STUDY RANGE # USED (Decadic) 32b 36a 36b 10c 10d 37a 38 22b 22c 39 37b 40 36c 20b 41 4.9 5.0 5.0 5.0 5.0 5.4 5.5 5.8 5.8 5.8 5.9 6.0 6.0 6.0 6.1 267 (peak) 413 (onset) 127 (VER) 181 (VER peak) 560 173 (ganglion cell) 151 (RT) 70 (b2) 84 (peak) 390 181 (onset) 146 (t) 139 (VEP cII onset) 170 (VEP cII onset) 380 (alpha block) 195 RANGE (or 3.5 x sample s. d.) OF LATENCIES OBTAINED (ms) ERG (local or gross) a-wave b-wave NONRETINAL CNS (N. II except as noted otherwise) BEHAVIORAL CNS 140
ratio

ERG2 CNS2

(Continued)

J. M. Williams

Luminance and the Hess Effect

268

APPENDIX IB (Concluded) STUDIES OF VISUAL LATENCY AS A FUNCTION OF STIMULUS INTENSITY


LOG-UNITS STIMULUS STUDY RANGE # USED (Decadic) 42 43 44 45 6.4 6.7 7.0 8.0 240 247 (VEP) RANGE (or 3.5 x sample s. d.) OF LATENCIES OBTAINED (ms) ERG (local or gross) a-wave 50 b-wave 63 455 NONRETINAL CNS (N. II except as noted otherwise) BEHAVIORAL CNS
ratio

ERG2 CNS2

J. M. Williams

Luminance and the Hess Effect

269

APPENDIX II NORMAL EQUATIONS FOR LEAST-SQUARES FIT OF VARIOUS LATENCY FGRMULAS

These equations are included here although the sums of squares of deviations from the data obtained actually were minimized empirically rather than analytically in the present work. In what follows, Charpentier's absolute latency formula will be treated first to illustrate the method, then the normal equations for the various relative latency formulas discussed in the text will be provided with no further comment. Source references for the

fomulas, many of which were adapted by the present author from the absolute latency formulas actually given by the proponents named, have been provided in the text above. Abbreviations used below are as follows: Data are indicated by

asterisks (*); parameters follow the semicolon in the functional notation; SS stands for sum of squares of deviations; the datumsample is of size r, each datum being indexed by a value of j, j being different for each different set of stimulus conditions under which the data were gathered.

J. M. Williams

Luminance and the Hess Effect

270

1. Charpentier's formula for absolute latency: tC = f(I; a, n, K) = a I-1/n + K . (II-1)

In general, any sum of squares of deviations SS of any set of r predicted values f, given some set of r data f*, is obtained by computing the value of
r

ss

fjf* j j=1

(11-2)

To find the minimum SS possible given the theoretical function in (II-l) and the data, it can be proven that it is sufficient to find the first partial derivatives of SS with respect to each theoretical parameter and then to set those derivatives simultaneously and independently equal to O. The result for each parameter is an equation called a normal equation because the parameters are considered independent and thus the partial derivatives are orthogonal (normal). Solving the normal

equations simultaneously (in principle) determines the parameter values minimizing SS, given the data. Unfortunately, the latency

formulas being considered are not at all linear, so analytic solution of the normal equations often is impossible and iterative algorithms must be used. In practise, the iterations for solution of the

normal equations often are not much easier computationally than those for direct minimization of SS empirically.

J. M. Williams

Luminance and the Hess Effect

271

In any case, proceeding from (II-2), the partial derivatives of (II-2) with respect to the parameters in (II-1) are obtained first by substituting (II-1) into (II-2) to get
r

SS

a I-1/n j j=1

K f * 2 . j

(II-3)

Now, consider only the partial derivative with respect to a:


r

SS/a

fj f* 2 j a j=1 aI-1/n K f*2 . j j a j=1


r

SS/a

(II-4)

Now, the partial derivative is evaluated in the same way as an ordinary derivative would be for all parameters other than a treated as constants. This is explained in introductory textbooks such as Therefore, using the

Johnson & Kiokemeister (1964, section l0.7).

chain rule of the differential calculus, we obtain from (II-4) the result
r

SS/a

2aI-1/n j j=1

K f * I-1/n . j j

(II-5)

Thus, the normal equation for a is

J. M. Williams

Luminance and the Hess Effect

272

aI-1/n j j=1

K f * I-1/n . j j

(II-6)

The second parameter of (II-1) is n so, proceeding in the same way as above, from (II-3),
r

SS/n

aI-1/n K f*2 j j n j=1 2 aI-1/n Kf* aI -1/n n2 ln Ij . j j j


j=1 r

(II-7) This is

Note the complication of (II-7) as compared with (II-6): because (II-1) is linear in a but not linear in n.

Setting (II-7)

equal to 0, which is possible only if n 0, we obtain the alternatives (roots), 0 0 0 = = = a n

I-1/n lnIj aI-1/nKf* j j j j=1

(II-8)

The third parameter f (II-1) is K; so, from (II-3),


r

SS/K

aI -1/n K f* 2 . j j K j=1

(II-9)

J. M. Williams

Luminance and the Hess Effect

273

Setting (II-9) equal to 0, we obtain


r

aI-1/n j j=1

K f* . j

(II-10)

This result is relatively simple because (II-1) is linear in K as well as in a. Thus, the three normal equations (assuming neither a nor n equal to 0) are found to be (II-6), (II-8), and (II-10) above. When these

are solved simultaneously, given the r experimental data, the resulting values of a, n, and K will be just those that minimize the SS end thereby provide the best possible fit in the least-squares sense.

2. Charpentier's formula for tC:

This is a relative latency formula

obtained by taking the difference between two functions as in (II-1) above but with two different values I1 and I2 of intensity I. have tC = f(I1, I2; a, n) = a (I1-1/n - I2-1/n) . We

So, as above,
r

SS

[ aI-1/n I-1/n 1j 2j j=1

f* ] j

(II-11)

J. M. Williams

Luminance and the Hess Effect

274

For parameter a,
r r

a I
j=1

-1/n 1j

-1/n 2 2j

f* I-1/nI-1/n j 1j 2j j=1

(II-12)

For parameter n, 0 0 0 = = = a n

[ aI-1/nI-1/n f*] [ lnI1jI-1/nlnI2j I-1/n ] 1j 2j j 1j 2j j=1

(II-13)

There is no equation for "K" because K cancelled out in the subtraction by which relative latency was obtained from the absolute latency formula (II-1) above.

3. Exner's formula for tE: tE = f(I1,I2; a, b)


r

(a logI1 + b)-1 - (a logI2 + b)-1 .

(II-14)

SS

[alog I1jb1 j=1

alog I2j b1 f * ] j

(II-15)

J. M. Williams

Luminance and the Hess Effect

275

For a,
r

0=

[alogI1jb1alog I2jb1 f* ] a logI j j=1

logI2j logI1j 2 alog I 1j b2 2j b

]
]

.(II-16)

For b,
r

0=

[ alog I1jb1alog I2jb1f* ] alog I j j=1

2j

1 2 alog I 1jb

.(II-17)

4. Pieron's formula for tP: tp = f(I1, 12; b, k) = k(I1b - I2b) . (II-18)

This is the same as tC above, with b -1/n.

5. Ives's formula for tI: tI = f(I1, 12; a, b) = (alogI1 + b)-1 + (alogI2 + b)-1 . This is the same as Exner's above (II-19)

6. Lythgoe's formula for tL: tL = f(Eu, El; c) = .02 log(Eu/El) - c . (II-20)

This formula has no free parameter, as Lythgoe (1938) assumes c near 0 in normal cases. See Rouf's, below.

J. M. Williams

Luminance and the Hess Effect

276

7. Roufs's formula for tR: tR = f(E0, E1; T)


r

-T ln(E1/E0) .

(II-21)

SS

[ TlnE1 /E0j j=1

f* ] j

(II-22)

For T,
r r

T [ lnE 1 /E 0 j ]
j=1

f* lnE1 /E0 j j j=1

(II-23)

8. A1pern's formula for tA: tA = f(Eu; n, K, T0) = T0 - K Eu-1/(n+1) . (II-24)

This is similar to one term of Pieron's formula above, with b = -1/(n+1).


r

SS

[ T0 j=1

KEu -1/(n+1) f* ] j j

(II-25)

For n, 0 0 0 = = = K n + 1

[ T0KEu -1/(n+1)f*] Eu -1/(n+1) lnEu j j j j j=1

(II-26)

J. M. Williams

Luminance and the Hess Effect

277

For K,
r

[ KEu-1/(n+1)T0f*] Eu -1/(n+1) j j j j=1

(II-27)

For T0,
r

KEu-1/(n+1) j j=1

T 0 f* . j

(II-28)

9. Hecht's formula for tHT: tHT = f[(I/I0)1, (I/I0)2; k] (II-29)

= [k log(I/I0)1]-1 - [k log(I/I0)2]-1 .
r

SS

{[ k log I/I01j ]1 j=1

[ klogI/I 0 2j ]

f* } . j

(II-30)

For k,
r

1 1 1 2 {[ log I /I0 1j ] [ log I /I0 2j ] } k j=1 f * {[ log I/I 0 1j ] [ log I/I0 2j ] j


1 j=1 r 1

. (II-31)

10. Enright's formula for tEN: tEN = f(Eu, El; Y, S) = log(Eu/El)[Y + (1/2)Slog(EuEl)] . (II-32)

J. M. Williams

Luminance and the Hess Effect

278

SS

j=1

S * [ log Eu /El j ] Y 2 log Eu El j fj

(II-33)

The normal equations are expressed explicitly in terms of the parameter solutions for Enright's formula, because it is more convenient than in the majority of the other cases above: For Y,

j=1

f* log Eu /El j j
r j=1

S 2 log Eu /El j ] log Eu E l j 2 j=1 [


2

(II-34)

[logEu /Elj ]

For S,

2 f log Eu El j Y log Eu /El jlog Eu El j S =


j=1 * j j=1

log Eu/ Elj[ logEu Elj]


j=1

.
2

(II-35)

11. Present proposed formula for t:

f(c, E; a, k)

K1 lnE + k a K2 lnE + k

K1 ln(cE) + k a K2 ln(cE) + k

(II-36)

in which Kl and K2 are functions of a given in the text.

J. M. Williams

Luminance and the Hess Effect

279

SS

j=1

K 1 ln Ej ka K 1 lncEjk a * fj K 2 ln Ej k K 2 lncEjk

(II-37)

For a, assuming a 0,

j=1

Aj Bj

B 'j C j C'j

in which Aj Bj B'j = = = K1 lnEj + k a K2 lnEj + k 2 K1 lnEj + k K2 lnEj + k 2 K1 ln(cE)j + k K2 ln(cE)j + k K2 ln(cE)j[K1 ln(cE)j + ka] [K2 ln(cE)j + k]2 K2 ln(E)j[K1 ln(E)j + ka] [K2 ln(E)j + k]2 . K1 ln(cE)j + k a K2 ln(cE)j + k - fj*

(II-38)

Cj

C'j

For k, assuming exposure duration t0 0,

Aj j=1

lncEj
2

[ K 2 lncEj k ]

[[

lnE j K 2 ln Ej k ]
2

]}

(II-39)

J. M. Williams

Luminance and the Hess Effect

280

APPENDIX III SOLUTION OF LINEAR FIRST-ORDER NONHOMOGENEOUS ORDINARY DIFFERENTIAL EQUATION

This refers to equation (28) of the text, which is linear because it is made up of a linear combination of the unknown function g and its derivatives: it is first-order because only a first derivative dg/dt is involved; it is nonhomogeneous because a function of time, 1n[E(t)] other than g(t) is involved; it is ordinary because no partial derivatives are involved (in the degenerate case being considered); it is a differential equation because the unknown is a function g(t) constrained by the equation, which makes a linear "statement" about g(t) and its derivatives. Without proof, it is assumed that equation (28) of the text may be rewritten as 1nE(t) dt = a dg + g dt , (III-1)

in which E(t) is the time-varying stimulus input. The solution depends upon restating (III-1) as the familiar formula of elementary calculus, d(u v) = u dv + v du . First, multiply (III-1) by exp(t/a): exp(t/a) lnE(t) dt = exp(t/a) a dg + exp(t/a) g dt . (III-3) (III-2)

J. M. Williams

Luminance and the Hess Effect

281

But now (III-3) may be rewritten immediately as

exp(t/a) 1nE(t)dt

[a exp(t/a)]dg + [exp(t/a)dt]g . u dv + du v

(III-4)

As can be seen, the left side of (III-4) formally is the same as d(uv), because the right side is. By definition, the integral of

d(uv) must be uv; therefore, the integral of the left side of (III-4) must be equal to uv, whatever uv might happen to be. But simply

reading on the right side of (III-4) reveals that u = a exp(t/a) and v = g. The function exp(t/a), called an integrating factor, was Therefore, we have

cleverly chosen just so this would happen.

exp(t/a)1nE(t) dt = d(u v)

(III-5)

which may be integrated formally as

expt/alnE tdt
or,

= uv

aexp t/ag ;

(III-6)

exp/alnE d
0

uv

aexpt/agt ;

(III-7)

or, solving algebraically for g(t),

J. M. Williams

Luminance and the Hess Effect

282

1 k exp [ t/a ]lnE d exp t/a a a 0

gt .

(III-8)

In the final answer (III-8), a by convention represents a time-constant of some process defined by (III-l), is a "dummy" variable introduced to prevent confusion with the variable t of the limits of integration in (III-7), and k is a constant of integration introduced because d(uv) on the right side of (III-5) would be the same whether derived from the function uv or from uv + k, and omitting the constant k would reduce the generality of the solution (III-8) by excluding legitimate solution-functions g(t).

J. M. Williams

Luminance and the Hess Effect

283

VITA Name: John Michael Williams

Address: 504 South Wall Street, #210, Carbondale, Illinois 62901. Prior Education: BA, Columbia University, Spring 1969 AM, University of Chicago, Winter 1976.

Honors: BA Magna cum laude and with Honors in Psychology Woodrow Wilson Fellow (elected 1969) Phi Beta Kappa (elected 1969) Psi Chi (elected 1969) Dissertation Research Award (1979-80) Sigma Xi (elected Associate, 1980).

Dissertation Title: Hess effect. Alfred Lit.

Luminance-dependent latency as measured by the

Written under the Major advisorship of Professor

Previous Publications: "A multiple-choice airstream design for olfactory discrimination training of small animals." Behavior Research Methods and

Instrumentation. 1970, 2, 195-197. (coauthored with Burton M. Slotnick). "Distortions of vision and pain: acid diethylamide." Two functional facets of d-lysergic

Perceptual and Motor Skills, 1979, 49, 499-528.

(Monograph Supplement).

You might also like