You are on page 1of 13

The Dimensions of Colour

COLOURS IN SPACE

All painters, whether working in digital or traditional mediums, are in a real sense navigators in
space. Whether they are aware of it or not, each touch of colour they apply can be represented,
using various systems, as a point within a space defined by three dimensions. The conceptual
colour space most familiar to painters, which is applicable to colours perceived as belonging to
surfaces, has the dimensions of hue, chroma (strength of colour) and lightness (the latter usually
known to artists as tone, greyscale value, or simply value). Colours perceived as independent light,
on the other hand, should be described in terms of the fundamental dimensions of hue, saturation
(purity of colour) and brightness. A sixth parameter, "colorfulness", used in current literature on
colour appearance models (Fairchild, 2004, 2005), is a function of both the saturation and the
brightness of a light stimulus.

All painters familiar with the concepts of the colour wheel and the tonal scale must presumably
have some awareness of this spatial aspect of their activities, but many do not seem to take much
advantage of this awareness. Beginners often seem to mix colours entirely by trial and error, with
little thought of what effect a colour is likely to have before they add it.

Many painters observe colours as being in a vague sense "warmer" or "cooler" than others, but do
not trouble to apply the more precise concepts of hue and chroma. A great many think of colour
mixing as applying pigment "recipes", obtained either secondhand from art instruction books on,
say, how to paint flesh colours, or by relying on their own mixing charts to remember how they
mixed a particular colour previously.

Typically, artists of this sort have little knowledge of the physical principles involved in creating
effects of light and shade, and often rely for such effects on crude and inaccurate rules of thumb,
such as "get the shadow colour by adding the complimentary" or "keep your color most intense on
the edges of the lighted areas".

Other artists however find it invaluable to think consciously of their colouring activities as
maneuvering through a three-dimensional space. They use this spatial conception in three main
ways, as a framework for (1) observing colour relationships, (2) selecting and mixing colour, and
(3) creating colour relationships from the imagination:

1. As a framework for observing colour relationships.

Most artists of this group do not try to copy each colour in their subject in isolation (the strategy of
every beginner). Instead, they use the concept of colour space as a frame of reference for grasping
the relationship of each colour to the totality of colours present. Tonal realist painters, for
example, typically observe relationships of hue, brightness and "colorfulness" in the light from
their subject, and then, by a process of either conscious or unconscious translation, identify each
individual colour in terms of the hue, lightness and chroma of the paint colour they will need to
use in order that the whole ensemble replicates the visual appearance of the subject. In practice,
this usually involves first selecting the most important (say) six to ten colours in the subject, and
finding the place of these in relation to each other (Figure 1.2). This begins the process of building
what I think of as a scaffolding for progressively finding the place of all remaining colours.

Most of these painters utilize some system of premixed pools of paint. Many make their colour
decisions in relation to a fixed system of colours, usually either a published colour classification
like the Munsell system, or some other methodical procedure for laying out a standard set of
predetermined colour mixtures. Others think instead in terms of hue, chroma and lightness
diffrences relative to their "scaffolding" colours. Many painters who stop short of going the full
Munsell nevertheless find it invaluable to at least think of lightness in relation to a scale of (usually
about nine) absolute greyscale values.

Figure 1.2. : Left: Lyndall by David Briggs, 2005, oil on canvas. Right: plan view (above)
and side view (below) of ten selected colours from the image plotted in YCbCr space using
the programme ColorSpace. (Note that, as in most illustrations on this site, the CbCr plane
is shown in reverse to its standard orientation, to place the spectral sequence of colours in
clockwise order, following Newton, Munsell, and the hue circle in HSB colour space, among
others).

2. As a framework for selecting and mixing colour.

Artists who think in terms of colour space do not need to remember recipes for mixing colours:
they understand that most colours can be mixed from any number of combinations of
components, as long as the target colour is within the three-dimensional gamut of those
components. They literally visualize colour mixing as shifting colour from place to place through
coour space. They decide on the changes in hue, chroma and lightness required, and predict in
advance what effect various additions are are likely to have. These crafty painters may, for
example, premix a pool of colour on the other side of a target colour, and add this in stages to
draw the colour methodically towards its target (Figure 1.3).

3. As a framework for creating colour relationships from the imagination.

The dimensions of colour form an essential conceptual framework for any kind of activity that
involves creating colour relationships from the imagination. In the nineteenth and early twentieth
centuries, much thought on colour spaces was directed towards discovering rules of colour
harmony, and there are still many echoes of this kind of investigation today. Here I am much more
concerned with the relevance of colour space to the creation of convincing effects of light from
the imagination. The concept of colour space provides an essential quantitative framework for
applying the simple physical laws that govern the behaviour of light and colour. If the artist gets
these right in a painting, the payoff can be a vivid glow of light. And, as with, for example,
perspective and anatomy, having the understanding that allows you to do something from the
imagination makes working from nature far more efficient.
Figure 1.5. Fundamental dimensions for colours perceived as (A) surface colour and (B)
independent light. Saturation refers to purity of colour of light, and is independent of brightness.
Chroma (strength of colour) depends on the saturation and the brightness of the light given off by
a surface (for a given level of illumination). Chroma therefore is not independent of lightness, but
is necessarily zero at maximum and minimum lightness (white and black respectively), and reaches
its greatest range at intermediate lightness levels.

Hue

The hue of a given colour may be defined as its closest match within the range of spectral (red,
orange, yellow, green, cyan, blue, violet, and their intermediates) and nonspectral (crimson to
purple) colours. Hues form a continuous circular scale, as in the traditional artists' colour wheel
(with yellow, red and blue opposite purple, green and orange respectively), and in the 360o range
of angles used to specify hue digitally.

Hue circles differ according to whether opposing colours represent additive complementary
relationships (relevant if our question concerns light stimulus), opponent relationships (relevant if
our question concerns our experience of colour) or pigmentary complements (relevant if our
question concerns paint mixing).

The traditional artist's colour wheel is here considered to be a compromise solution, influenced
partly by opponent and partly by pigmentary relationships, while the hue circle used in digital
painting programmes represents additive complementary relationships.

Lightness and Chroma

Our visual system judges colours of surfaces in relation to a real or inferred white surface, which
makes possible certain colour experiences, such as grey and brown, that are not experienced in
colours seen as independent light. Because of this special way of seeing surface colours, different
terms are needed for describing light and colour intensity for surfaces and for independent light
(Figure 1.5).

Lightness and chroma (Figure 1.5A) are the appropriate terms for describing colours perceived as
attributes of surfaces. They apply to colours of surfaces seen in nature, or depicted in an image, or
making up an image, as long as these colours are viewed in a normal way (called related viewing) ,
as opposed to being observed as independent light, as for example through an aperture. The
terms apply to colours making up an image surface irrespective of whether that surface reflects
light (e.g. a photograph, painting, or projector screen), transmits light (e.g. a stained glass window)
or emits light (e.g. a computer or TV screen), as long as these surfaces are viewed in relation to an
actual or inferred component seen as being white.

Lightness is defined as the perceived brightness of a surface compared to that of a perfect white
surface under the same illumination. It ranges from black to white and can be quantified in various
ways, such as a scale of 10 (as in the Munsell system) or 100 (as in lightness or "L" in Lab colour
space, used in Photoshop). For surfaces that reflect light, lightness is the perceptual correlative of
physical reflectance. These two parameters however have a nonlinear relationship - a surface that
looks visually halfway between black and white has a lightness of 50%, but reflects only about 18%
of the light energy reflected by a white surface.

Artistic readers should feel free to substitute for the technical term lightness whichever of the
alternative terms tone, greyscale value, or value they prefer. Lightness is usually shown on an axis
through the centre of the hue circle.

Chroma is the strength of a surface colour, the degree of visual difference from neutral grey. On
light-reflecting surfaces, chroma depends on the reflectance curve of the surface - to have high
chroma a surface must reflect light of high spectral purity and plenty of it. The possible range of
chroma is therefore strongly dependent on lightness and hue. At maximum lightness (white) and
minimum lightness (black), chroma can only be zero. As we move away from these extremes the
range of possible chroma increases, up to a maximum at a lightness level that depends on hue -
high for yellow, low for violet-blue. All these points are true both for theoretical optimal colours
and for actual surface colours.

The term chroma was introduced by Albert Munsell (1905), who first quantified the concept, and
in doing so established that the maximum chroma that can be produced in surface paints differs
considerably between different hues. Chroma can be represented by distances outward from the
centre of the hue circle, either in equal perceptual steps (and hence to varying distances from the
centre), as in the Munsell system, or normalized to produce a regular circular arrangement, as in
the traditional artist's colour wheel.

Brightness, Saturation and "Colorfulness"

Independent light is seen as being brighter or dimmer, rather than having lighter or darker
greyscale values, and this scale of perceived intensity is called brightness (Figure 1.5B). Brightness
can be measured either on an open-ended scale, or on a finite scale set by the range of possible
values on a given device (e.g. the values of R, G and B in RGB space). Brightness is the perceptual
correlative of the physical quantity of luminance, which is radiance (amount of light energy),
weighted according to the influence of each wavelength on the human visual system. Brightness
bears the same nonlinear relationship to luminance/radiance as that between lightness and
reflectance.

The dimension of perceived purity (or relative colour intensity) for independent light is called
saturation, the perceptual correlative of the physical quantity of spectral purity. Saturation can be
quantified in various ways, but all divide up the range from zero for white light to a maximum
either for monochromatic light, or the for most saturated light possible for a given device. In
contrast to the close dependence of maximum chroma on lightness and hue noted above,
saturation can vary throughout this range at any level of brightness, irrespective of hue.

Current literature on colour appearance models quantifies saturation as the "colorfulness"


(absolute colour intensity) of a light stimulus relative to its brightness (Fairchild, 2004). Thus if two
lights have the same saturation, but one is brighter, the latter will exhibit more "colorfulness".

In normal (related) viewing, a colour must be substantially brighter than a perceived white surface
under the same illumination if it is to appear as an independent light source. However, all surface
colours can be seen as independent light if viewed in an unrelated way, as through an aperture.
This is true irrespective of whether the surface reflects, transmits, or actually emits light. The
parameters of brightness, saturation, and "colorfulness" therefore apply not only to independent
light sources, but also to surface colours viewed in this unrelated manner. Consequently, colours
of surfaces (for example, on the screen you are looking at) can be described in terms of all six
parameters, depending on how those surfaces are observed. Only three of these parameters are
independent, however.

The parameters of brightness and saturation are less familiar to most artists than those for
surfaces, although digital artists will have encountered these terms in programmes such as
Photoshop, where they are used in the context of the colour space known as HSB. Brightness and
saturation in HSB are specific incarnations of those concepts, defined in ways that relate to the
possible range of RGB (screen) colours. Brightness (B) in particular is quite a different concept in
HSB to absolute brightness, and refers to brightness relative to the maximum possible for RGB
colours of a given hue and saturation. Saturation (S) in HSB measures saturation relative to the
maximum possible for that hue in RGB colours.

Confusion of the concepts of saturation and chroma is at present endemic in most writing on
colour, not only for artists, and the two terms are very frequently used interchangeably. The
distinction may sound academic, but it isn't. Two surfaces can reflect/ transmit/ emit light of
exactly the same saturation, but if one gives off more light under the same conditions, it will have
higher chroma. For example, in Figure 1.6, strip AB has higher chroma than CD, but both reflect
light of the same saturation. "Colorfulness", not saturation, is the nearest equivalent for light to
the chroma of surfaces. Surfaces with high chroma tend to reflect light of high "colorfulness", the
latter varying however in proportion to the the level of illumination.

Figure 1.6. Is A the same colour as B or D? It all depends on how you look at it, which is why we
need to be so careful about terminology. Viewed as an image, we see two strips, AB and CD, each
of uniform lightness and chroma, AB lighter and higher in chroma than CD. Seen as screen colours,
however, A and B are surfaces of different lightness and chroma, and emit light of different
brightness and "colorfulness". They have been painted this way in order to represent the brighter
and more "colorful" light that a surface of uniform chroma reflects where it is more strongly lit.
Seen as light, all four areas emit light of the same (100% pure) saturation - in each area, only the
red phosphors are glowing (right). Areas A and D happen to emit light of the same hue, saturation
and brightness - that is, they are identical colours when considered as components of the image,
even though they are perceived as very different colours in the subject. Image: David Briggs,
Photoshop CS2.
Strictly speaking, hue, chroma, lightness, brightness, saturation, and "colorfulness" are all
psychological dimensions of subjective perception. Generally when we talk about them
quantitatively, we are really talking in each case about a corresponding psychophysical dimension
based either on the measurable properties of light, or defined with reference a set of standard
physical surfaces. Even under ideal conditions we should expect a close but not necessarily perfect
agreement between the two types of scale.

The next section reviews some of the basic facts of light and colour.

BASICS OF COLOUR VISION

BASICS OF COLOUR VISION

All students know about the colour wheel, the spectrum, and the primary colours. They can tell
you that the primary colours can be mixed to make all other colours, and can not themselves be
mixed at full intensity from other colours. They may know just enough to get into endless debates
as to whether the "real" primaries are the red, yellow and blue of the conventional artists colour
wheel, or the yellow, magenta and cyan printer's primaries, or perhaps the three optimal
primaries for mixing coloured lights - red, green and blue. But few have reflected on the
inconsistencies that seem to lurk within this universally held knowledge. For example, few have
wondered why there are three primary colours. If the visible spectrum consists of a continuous
range of wavelengths, why should just three colours be special? Why not five? And why for that
matter do colours form a circle, when the spectrum does not?

These questions do not arise in most treatments of "colour theory" for artists. Though
fundamental to the subject, they do not occur to us, perhaps because we internalize our
knowledge about colour at an early and uncritical age. To answer them we need to understand
some basics of colour vision. In broad outline, the processing of colour information by our visual
system seems to operate in three successive stages:

1. Trichromatic image capture: colour information is recorded by the responses of the L,M and S
cone cells in the retina.

2. Opponent processing: responses from the L,M and S cones are converted into three signals, for
brightness, yellowness vs blueness, and redness vs. greenness respectively, the latter two
providing our perception of hue and saturation.

3. Processing for colour constancy: information on hue, brightness and saturation from throughout
the visual field is analysed and resolved into an interpretation of the hue, brightness and
saturation of the illumination, and the hue, value and chroma of the visible surfaces.

Why three primaries? Well, the number three comes ultimately from the fact that there are three
kinds of colour-discriminating receptor cells, called cone cells, in the human retina. The three cone
types have broadly overlapping ranges of sensitivity, and are designated L, M and S according to
the location of their peak sensitivities in the long, medium and short wavelength parts of the
spectrum respectively (Figure 3.1). This trichromatic model of colour vision was in fact first
prompted by the evidence of the three painter's primaries.We will return to how these three cone
types result in our three primary colours at the end of this page.

The circular arrangement of hues, on the other hand, is believed to derive from the way the inputs
from these three cone types are subsequently processed. According to the widely held opponent
model of colour vision, proposed in the late nineteenth century by Ewald Hering and subsequently
quantified in particular by Hurvich and Jameson (1957), inputs from the three cone types are
added and subtracted together to create three signals: brightness, redness vs greenness (r/g), and
yellowness vs blueness (y/b). The 360o range of possible combinations of positive and negative r/g
and y/b values creates the circular range of hue of the familiar colour wheel (Figure 3.2, 3.3).

Figure 3.2. Brightness ("achromatic"), y/b and r/g opponent signals for light throughout the
spectrum. Hurvich and Jameson showed that by using a blue or yellow AND a red or green light
they could match by hue cancellation all wavelengths of the visible spectrum, thus determining
quantitatively the r/g and y/b components of the colour sensation induced by each wavelength.
Image source: http://webvision.med.utah.edu/imageswv/KallColor15.jpg

While details of the neurological mechanisms enabling the process are still disputed, the broad
outline of the opponent model seems firmly founded on the evidence of our own experience.
Firstly, the four opponent hues are the only ones that we can experience as unique, that is,
unmixed with other colours - all others seem to be mixtures of two opponent hues. For example,
we can experience or imagine a red that is neither orangeish or purplish, whereas all orange
colours are both yellowish and reddish. Secondly, these four colours are arrayed in these pairs,
such that no colour can be red and green, or yellow and blue, at the same time. Different sources
vary somewhat on the way the visual system calculates opponent signals, but Kuehni (2005) gives
the following account as representative (other sources give brightness as L+M+S).

(Brightness | |L + 0.5
|r/g (redness vs greenness) signal | |L-M
|y/b (yellowness vs blueness) signal | |L + M - S

Thus L>M creates the sensation of redness, M>L the sensation of greenness, L+M>S the sensation
of yellowness, and S>L+M the sensation of blueness. We experience each of these sensations in its
pure state, the unique hue, not when its signal is at its maximum, but when the other signal is at
zero. At the long-wavelength end, high L, low M, and zero S responses create positive r/g and y/b
signals, and we experience a broad zone of orange-red colours. Moving left towards shorter
wavelengths, the L response peaks before the M response (Figure 3.1), so that r/g begins
dropping.

At wavelengths around 577 nanometers the L and M cone influences effectively balance out, the
r/g signal passes through zero, and we experience unique yellow, that is, a yellow that is neither
reddish nor greenish. To the left of this point, negative r/g signals result in a series of greenish
colours. At around 513 nm the influence of S effectively balances (L+M), so that the y/b signal
drops to zero, and we experience unique green, neither yellowish nor bluish.

Throughout the remainder of the spectrum, S predominates over (L+M), so that the y/b signal is
negative, and all of the colours are bluish in character. Near the short wavelength end of the
spectrum the M response drops below the L response, and we get positive r/g signals and reddish
colours again, to which we apply the name violet. The second point of zero r/g, which we
experience as unique blue, occurs at wavelengths around 475 nm. All of the wavelengths quoted
here are subject to considerable individual variation.

Single wavelengths can not create the full 360 degree range of hues. Unique red, for example,
does not occur in the visible spectrum, because all long wavelengths create a positive y/b
component. Colours from unique red through magenta to red-violet can only be induced by a mix
of wavelengths from the red and violet ends of the spectrum.

The L, M and S cones have sometimes been described as red-, green-, and blue- sensitive
respectively. Many authors have been quick to point out that these descriptions are incorrect,
since all three cone types have broadly overlapping ranges of sensitivity, and the L and M cones
are sensitive throughout the visible spectrum (Figure 3.1). These descriptions are not entirely
incorrect, however, if we take into account the way the inputs from these cones are processed in
the opponent model. The three cone types effectively divide the visible spectrum into three
bands, in each of which the response of one cone type predominates over the other two.

At the short wavelength end is a band of blue to violet colours, where S > (L + M) gives negative
y/b values. The remainder of the spectrum is divided into a middle band of greenish colours,
where M > L gives negative r/g values, and a long-wavelength band of reddish colours, where R >
M gives positive r/g values. The first two bands overlap in the cyan part of the spectrum, while the
second and third meet at yellow. This division of the spectrum into a reddish, a greenish and a
blue-violet band is quite evident on visual inspection (Figure 3.4).
Figure 3.4. Solar spectrum. The orange-red, green and violet-blue bands, in which the responses of
the L, M and S cones respectively predominate over the other two, are clearly evident on visual
inspection of spectra such as this complete solar spectrum. Image source:
http://www.adlerplanetarium.org/cyberspace/sun/learning.html (Credit: Nigel Sharp,
NOAO/NSO/Kitt Peak FTS/ AURA/NSF).

The great importance of these three spectral bands is that if we have three light sources, one from
each band, we can make light mixtures with any possible combination of strongly positive and
negative r/g and y/b signals, and thus make strongly coloured mixtures throughout the full 360o
range of hues. These three lights thus make optimal primary colours for additive colour mixing. We
will see in a later section how the threefold nature both of the ideal subtractive primaries and of
our closest pigmentary equivalents results in turn from these three additive primaries.

Interestingly, both the division into a reddish, a greenish and a blue-violet band, and the derivation
of other colours (yellow) by the interaction of these colours, were recorded in the rainbow in
antiquity in Aristotle's Meteorologica.
ADAPTATION

We all know that the muscles of the iris can increase or decrease the size of the pupil in response
to dimmer or brighter light respectively, but this mechanism is only a minor component of our
ability to adjust to different light intensities. Alongside the shift between rod (low light) and cone
vision, the main process involved is called adaptation, which refers to our ability to adjust the
sensitivity of the receptor cells in the retina in response to the general level of illumination. On
going into a darkened room we may at first see little or nothing, but as the sensitivity of our
receptor cells slowly adjusts upwards, we begin to see more and more detail. On returning to a
more brightly illuminated environment we much more rapidly (and sometimes painfully) adjust
the sensitivity of those cells to the prevailing illumination.

The importance of adaptation to colour vision comes from the fact that the L, M and S cone
systems can to a certain extent adapt independently to the prevailing illumination if one set of
cones is more or less strongly stimulated than the others. For example under incandescent
lighting, where the S cones are less strongly stimulated than the M or L cones, the former increase
their relative sensitivity, causing the light to seem less strongly coloured than it otherwise would
appear.

This adjustment of the input from the cones should not be (but often is) confused with colour
constancy, which relates to the processing of that input into an interpretation as to the local
colour of surfaces and the quality of the illumination.

Figure 3.5. Afterimages and successive contrast. Stare fixedly at the centre of circle A for at least
20 seconds, then immediately look at the centre of circle B for about ten seconds, noting the
changing afterimage. Each sector will display an afterimage the colour of the additive complement
of the stimulus colour. Repeat the procedure, this time looking immediately at the centre of circle
C. Now the colours of the afterimage in each sector will influence the appearance of the red colour
in an example of successive contrast.

The coloured afterimages seen after exposing the eye to a coloured light for a period of time can
be explained in terms of changes in the relative adaptation of the three cone types. After a period
of exposure to coloured light, the cone type or types that are relatively weakly stimulated by that
light, due to a paucity of certain wavelengths, become proportionately dark-adapted. When
neutral light is restored, a temporary illusion of a light composed of the "missing" wavelengths is
seen.

Ewald Hering discussed afterimages as evidence for his opponent model of colour vision, and very
numerous subsequent authors down to the present have asserted that coloured afterimages take
the colour of the opponent colour in Hering's system. They do not - they in fact take the colour of
the additive complement - that colour of light that mixes with the stimulus to make white light.
For example, the afterimage seen after viewing red is cyan, not green. Pridmore (2008) examines
the history of this confusion, and points out that Hering himself never actually claimed that
chromatic induction in coloured afterimages was an opponent process, only lightness induction.
The afterimage takes the colour of the opposite stimulus, not the opposite colour experience.

The colour of an afterimage influences the apparent colour of objects viewed subsequently to the
stimulus, a phenomenon known as successive contrast. For example in Figure 3.5C, the
afterimages of red, yellow and magenta (the three colours that contain red light) dull the
appearance of the red, while the afterimages of the other three colours intensifies it. Successive
contrast thus also goes towards the additive complementary.

Successive contrast resulting from adaptation is the actual explanation of the phenomenon
sometimes mislabelled "fatigue" of the eye - for example, the apparent dulling of a high-chroma
red surface after being examined fixedly for a few seconds. In looking at a red object the L cones
are not being "fatigued" any more than they are in looking at a white object - the effect is caused
by the increased sensitivity of the other cone types.

COLOUR CONSTANCY

Our visual system does more than provide us with an image of the world. An image may contain
an array of light and dark areas, but which of these areas represent shadows, and which represent
dark-coloured objects? Does a red area in the image represent a red object in white light, or a
white object in red light? Is a particularly bright point in the image a light-coloured surface, or a
light source? To separate the effects of lighting and surface colour, the image must be interpreted.
Since we normally experience objects as having definite colours without having to make a
conscious effort of interpretation, it follows that this interpretation must be occurring
subconsciously as part of the processing of the image by our visual system. This feature of our
visual system allows us to recognize without conscious effort the local colour of objects under
lights of differing hue, brightness, and saturation, a capability referred to as colour constancy.

Colour constancy is never more than partial. Under weakly to moderately strongly coloured
illuminants, an array of surface colours may approximately maintain their perceived hues, but
changes in apparent lightness in relation to each other are inevitable (see Figure 10.8, slider near
middle of scale). Colours close to the hue of the illuminant tend to appear relatively lighter;
dissimilar hues tend to appear darker. The effect however depends on the exact spectral power
distributions both of the illuminants and of the surfaces. Two surfaces that match under a white
light may not match under a coloured light, or even under another white light with a different
spectral power distribution, a phenomenon known as metamerism. These factors naturally play
havoc with the tonal scheme of a painting, which is why paintings should if possible always be
executed under similar lighting conditions to those under which they will be viewed

Adaptation to the colour of a light source, which modifies the data provided by the eye by
adjusting the relative sensitivities of the three cone types, should not be confused with this
process of interpreting that data into illuminant and surface colour. Adaptation to the colour of
weakly to moderately strongly coloured light sources assists the process of interpreting surface
colour by boosting the available information on colour differences, while causing us to
underestimate the strength of colour of the illuminant. Surfaces under a monochromatic light,
however, can convey no information on differences in hue and chroma to the eye, and so differ
visibly only in lightness.

Given its ability to automatically interpret surface colour for us, we can infer that our visual system
must in some way be making a judgement as to what a white surface would look like in any part of
the visual field, and judging all colours by comparison with this white. A surfaces giving off light
whose brightness is consistent with being the result of diffuse reflection is seen thus as coloured
surface having a greater or lesser degree of "greyness". A colour that is too bright, compared to
this inferred white surface, to be the result of diffuse reflection is seen either as a fluorescent
surface, an independent light, or a specular reflection of an independent light. Evans (1974)
suggested the term brilliance for this scale of appearance, and used the term zero grey point for
the point where colours exhibit neither greyness nor fluorescence/luminescence. I argue here that
in a digital image, all colours having the same relative brightness ("B" in HSB) as a white surface at
the same location will be seen as having zero greyness.

Several dramatic optical illusions demonstrate colour constancy in action. In the checkerboard
illusion by Edward Adelson's (Figure 3.6A), the two squares marked A and B are actually identical
in lightness on the image (i.e. they are the same grey), but our visual system calculates that in a
shadow area this grey must belong to a white surface, while in the lit area the same grey must
belong to a dark surface, and that is how we see them. In the same way, in the cube illusion by R.
Beau Lotto (Figure 3.6B), our visual system sees the same image colour as being dark brown in the
context of strong lighting, and light orange where the same image colour appears in a deeply
shaded context. In the cross-piece illusion , also by Lotto (Figure 3.6C), the colour at the
intersection of the two rods is actually an identical colour (grey) in both cases, but in the context
of apparently yellow illumination on the left and blue illumination on the right, this is judged, and
seen, to be the reflectance of a blue-grey object and a yellow object respectively.

Figure 3.6. Three optical illusions demonstrating colour constancy in action (follow links for larger
images). A. The checkerboard illusion of Edward Adelson. B. The cube illusion of R. Beau Lotto. C.
The cross-piece illusion of R . Beau Lotto

In each case these comparisons are made unconsciously, and what we see in our normal way of
looking is the inferred local colour. Tonal painters have to learn above all to look at their subjects
with a different attitude to normal viewing, in order to judge objectively the hue, "colorfulness",
and brightness of the light coming to from each point in the subject. That is, we have to switch off
one kind of processing - one that is built into our visual system, wherein each colour is compared
with an inferred white - and learn a completely different kind of processing, where each colour is
compared with the full range of colours in subject as a whole. With practice we can learn to switch
at will between this painter's way of seeing and our normal mode of vision. But we always need to
be on guard against the tendency to slip into judging colours in constancy mode, that is, to paint
their perceived local colour, instead of the colour that we need to create the illusion of that
colour. The problem is very similar to the difficulties encountered in foreshortening in drawing,
where we also need to learn to see and draw what is actually in front of our eyes, and not what
our brain works out for us.

At this point the beginning painter might ask: "well, if that's the way it looks to my eyes, shouldn't
I paint it that way?" The answer to this is a definite no - if we can recreate the stimulus that
created the appearance, we will create the effect the we see in our subject; if we instead chase
the appearance, we will create something different.

Certain mechanical tricks or devices that are sometimes recommended to students for observing
colour objectively can be workable, but most have serious difficulties and limitations. For example,
the idea that you can hold up paint on a brush, palette knife or other device and match it with
your subject is in general workable only if you have some way of turning up the illumination on
your brush until you can match the brightest highlight on your subject with the tone of your paint,
and can keep the illumination at the same level while you compare the other colours. (One
teacher currently advertising such a method on the internet seems to get around this problem by
having his students paint only dimly lit subjects).

These methods of course also eliminate any option of translating the tonal range of the subject
into a your own choice of tonal level and range in your painting. Devices involving an aperture in a
card that bears a greyscale or colour chips for comparison suffer from the same difficulties and
limitations, and in addition run the risk of giving an excessive impression of the brightness and
"colorfulness" of colours seen in isolation, which can be avoided only if the colours are continually
compared with the brightest colours in the subject. The latter comparison can be made very
effectively however by using a blank card with two apertures, which can be moved towards or
away from the observer in order to compare more and less separated points.

SIMULTANEOUS CONTRAST AND ASSIMILATION

Simultaneous contrast and assimilation refer to the tendencies of the visual appearance of a
surface colour to be influenced by adjacent and interspersed colours respectively. The causes of
both of these slight failures of colour constancy are still under active scientific investigation. Both
may be broadly regarded as reflecting the fact that our visual system does not work like an
objective measuring device, sending raw light and colour data to the brain. We see always by
comparisons.

In simultaneous contrast the appearance of colours moves away from that of the surrounding
colour in all three dimensions of surface colour - hue, chroma and lightness (Figure 3.7, 3.8, 3.9).

Figure 3.7. Simultaneous contrast of hue and chroma. The purplish-red colour (A) looks more
purple against red (B), more red against purple (C), lower in chroma against bright purple-red (D),
and progressively higher in chroma against dull red-purple (E), grey (F), and green (G).

Figure 3.8. Simultaneous contrast of lightness and chroma. The same purple-red looks darker
against a lighter (B) and lighter against a darker (C) colour of similar hue and chroma. The
intensified change in appearance when combined with a contrast in chroma (D,E) was used in the
well-known optical illusion of this form devised by Josef Albers.

Figure 3.9. Simultaneous contrast of hue. A medium grey (A) is seen against a background of
similar lightness and green (B), magenta (C), red (D), cyan (E), blue (F) and yellow (G) hue. In each
case the appearance of the colour moves towards the hue of the additive complement.

An important special case of simultaneous contrast of hue is the phenomenon of complementary


colours in shadows. When we have a strongly coloured main light and a neutral secondary light,
the colours in the shadows of the main light shift in appearance towards the complementary of
the colour of the light. (Do not confuse this phenomenon with the hobby painter's recipe of adding
the pigmentary complementary of the colour of the object to get its colour in shadow). The
induced complimentary effect is the basis of the artist's conventional rule of warm lights creating
cool shadows and vice versa.

Simultaneous contrast is responsible for the well-known artists problem that paint mixtures may
look quite different on the palette (especially a white palette) to their appearance on the painting.
Experience and a growing capacity to see colours objectively can help here, as can devices such as
mixing over a grey scale placed under a glass palette. The strategies of using a coloured ground
closer to the tonal level of the finished painting than stark white, and of covering the canvas at an
early stage with a mosaic flat colours can both minimize problems caused by simultaneous
contrast on the canvas. The concept of a colour space reminds us that despite the shifting
appearance of each of our colours in different surroundings, the colours in any painting
nevertheless have a fixed relationship to each other (shifts due to different light sources excepted)
- and it is the business of the painter to get these relationships right.

Assimilation is a phenomenon having the opposite effect to simultaneous contrast, and is seen
when small areas of colour are closely interspersed. Examples include the illusions known as the
Von Bezold spreading effect, neon colour spreading, and White's illusion. In each case, surfaces
change appearance by moving towards the colour of the interspersed areas.

You might also like