Вы находитесь на странице: 1из 173

Mathematics and

art

Mathematics in art: Albrecht Dürer's copper plate


engraving Melencolia I, 1514. Mathematical
references include a compass for geometry, a magic
square and a truncated rhombohedron, while
measurement is indicated by the scales and
hourglass [1]
hourglass.[1]

Wireframe drawing[2] of a vase as a solid of


revolution[2] by Paolo Uccello. 15th century

Mathematics and art are related in a


variety of ways. Mathematics has itself
been described as an art motivated by
beauty. Mathematics can be discerned in
arts such as music, dance, painting,
architecture, sculpture, and textiles. This
article focuses, however, on
mathematics in the visual arts.

Mathematics and art have a long


historical relationship. Artists have used
mathematics since the 4th century BC
when the Greek sculptor Polykleitos
wrote his Canon, prescribing proportions
based on the ratio 1:√2 for the ideal
male nude. Persistent popular claims
have been made for the use of the
golden ratio in ancient art and
architecture, without reliable evidence. In
the Italian Renaissance, Luca Pacioli
wrote the influential treatise De Divina
Proportione (1509), illustrated with
woodcuts by Leonardo da Vinci, on the
use of the golden ratio in art. Another
Italian painter, Piero della Francesca,
developed Euclid's ideas on perspective
in treatises such as De Prospectiva
Pingendi, and in his paintings. The
engraver Albrecht Dürer made many
references to mathematics in his work
Melencolia I. In modern times, the
graphic artist M. C. Escher made
intensive use of tessellation and
hyperbolic geometry, with the help of the
mathematician H. S. M. Coxeter, while
the De Stijl movement led by Theo van
Doesburg and Piet Mondrian explicitly
embraced geometrical forms.
Mathematics has inspired textile arts
such as quilting, knitting, cross-stitch,
crochet, embroidery, weaving, Turkish
and other carpet-making, as well as kilim.
In Islamic art, symmetries are evident in
forms as varied as Persian girih and
Moroccan zellige tilework, Mughal jali
pierced stone screens, and widespread
muqarnas vaulting.

Mathematics has directly influenced art


with conceptual tools such as linear
perspective, the analysis of symmetry,
and mathematical objects such as
polyhedra and the Möbius strip. Magnus
Wenninger creates colourful stellated
polyhedra, originally as models for
teaching. Mathematical concepts such
as recursion and logical paradox can be
seen in paintings by Rene Magritte and in
engravings by M. C. Escher. Computer art
often makes use of fractals including the
Mandelbrot set, and sometimes explores
other mathematical objects such as
cellular automata. Controversially, the
artist David Hockney has argued that
artists from the Renaissance onwards
made use of the camera lucida to draw
precise representations of scenes; the
architect Philip Steadman similarly
argued that Vermeer used the camera
obscura in his distinctively observed
paintings.
Other relationships include the
algorithmic analysis of artworks by X-ray
fluorescence spectroscopy, the finding
that traditional batiks from different
regions of Java have distinct fractal
dimensions, and stimuli to mathematics
research, especially Filippo Brunelleschi's
theory of perspective, which eventually
led to Girard Desargues's projective
geometry. A persistent view, based
ultimately on the Pythagorean notion of
harmony in music, holds that everything
was arranged by Number, that God is the
geometer of the world, and that therefore
the world's geometry is sacred, as seen
in artworks such as William Blake's The
Ancient of Days.
Origins: from ancient Greece
to the Renaissance

Polykleitos's Canon and


symmetria

Roman copy in marble of Doryphoros, originally a


bronze by Polykleitos

Polykleitos the elder (c. 450–420 BC)


was a Greek sculptor from the school of
Argos, and a contemporary of Phidias.
His works and statues consisted mainly
of bronze and were of athletes.
According to the philosopher and
mathematician Xenocrates, Polykleitos is
ranked as one of the most important
sculptors of classical antiquity for his
work on the Doryphorus and the statue of
Hera in the Heraion of Argos.[3] While his
sculptures may not be as famous as
those of Phidias, they are much admired.
In the Canon of Polykleitos, a treatise he
wrote designed to document the
"perfect" anatomical proportions of the
male nude, Polykleitos gives us a
mathematical approach towards
sculpturing the human body.[3]
Polykleitos uses the distal phalanx of the
little finger as the basic module for
determining the proportions of the
human body.[4] Polykleitos multiplies the
length of the distal phalanx by the square
root of two (√2) to get the distance of
the second phalanges and multiplies the
length again by √2 to get the length of
the third phalanges. Next, he takes the
finger length and multiplies that by √2 to
get the length of the palm from the base
of the finger to the ulna. This geometric
series of measurements progresses until
Polykleitos has formed the arm, chest,
body, and so on.[5]
The influence of the Canon of Polykleitos
is immense in Classical Greek, Roman,
and Renaissance sculpture, many
sculptors following Polykleitos's
prescription. While none of Polykleitos's
original works survive, Roman copies
demonstrate his ideal of physical
perfection and mathematical precision.
Some scholars argue that Pythagorean
thought influenced the Canon of
Polykleitos.[6] The Canon applies the
basic mathematical concepts of Greek
geometry, such as the ratio, proportion,
and symmetria (Greek for "harmonious
proportions") and turns it into a system
capable of describing the human form
through a series of continuous geometric
progressions.[4]

Perspective and proportion

Brunelleschi's experiment with linear perspective

In classical times, rather than making


distant figures smaller with linear
perspective, painters sized objects and
figures according to their thematic
importance. In the Middle Ages, some
artists used reverse perspective for
special emphasis. The Muslim
mathematician Alhazen (Ibn al-Haytham)
described a theory of optics in his Book
of Optics in 1021, but never applied it to
art.[7] The Renaissance saw a rebirth of
Classical Greek and Roman culture and
ideas, among them the study of
mathematics to understand nature and
the arts. Two major motives drove artists
in the late Middle Ages and the
Renaissance towards mathematics. First,
painters needed to figure out how to
depict three-dimensional scenes on a
two-dimensional canvas. Second,
philosophers and artists alike were
convinced that mathematics was the true
essence of the physical world and that
the entire universe, including the arts,
could be explained in geometric terms.[8]

The rudiments of perspective arrived


with Giotto (1266/7 – 1337), who
attempted to draw in perspective using
an algebraic method to determine the
placement of distant lines. In 1415, the
Italian architect Filippo Brunelleschi and
his friend Leon Battista Alberti
demonstrated the geometrical method of
applying perspective in Florence, using
similar triangles as formulated by Euclid,
to find the apparent height of distant
objects.[9][10] Brunelleschi's own
perspective paintings are lost, but
Masaccio's painting of the Holy Trinity
shows his principles at work.[7][11][12]

Paolo Uccello made innovative use of perspective in


The Battle of San Romano (c. 1435–1460).

The Italian painter Paolo Uccello (1397–


1475) was fascinated by perspective, as
shown in his paintings of The Battle of
San Romano (c. 1435–1460): broken
lances lie conveniently along perspective
lines.[13][14]
The painter Piero della Francesca (c.
1415–1492) exemplified this new shift in
Italian Renaissance thinking. He was an
expert mathematician and geometer,
writing books on solid geometry and
perspective, including De Prospectiva
Pingendi (On Perspective for Painting),
Trattato d'Abaco (Abacus Treatise), and
De corporibus regularibus (On Regular
Solids).[15][16][17] The historian Vasari in
his Lives of the Painters calls Piero the
"greatest geometer of his time, or
perhaps of any time."[18] Piero's interest in
perspective can be seen in his paintings
including the Polyptych of Perugia,[19] the
San Agostino altarpiece and The
Flagellation of Christ. His work on
geometry influenced later
mathematicians and artists including
Luca Pacioli in his De Divina Proportione
and Leonardo da Vinci. Piero studied
classical mathematics and the works of
Archimedes.[20] He was taught
commercial arithmetic in "abacus
schools"; his writings are formatted like
abacus school textbooks,[21] perhaps
including Leonardo Pisano (Fibonacci)'s
1202 Liber Abaci. Linear perspective was
just being introduced into the artistic
world. Alberti explained in his 1435 De
pictura: "light rays travel in straight lines
from points in the observed scene to the
eye, forming a kind of pyramid with the
eye as vertex." A painting constructed
with linear perspective is a cross-section
of that pyramid.[22]

In De Prospectiva Pingendi, Piero


transforms his empirical observations of
the way aspects of a figure change with
point of view into mathematical proofs.
His treatise starts in the vein of Euclid: he
defines the point as "the tiniest thing that
is possible for the eye to
comprehend".[a][8] He uses deductive
logic to lead the reader to the perspective
representation of a three-dimensional
body.[23]

The artist David Hockney argued in his


book Secret Knowledge: Rediscovering
the Lost Techniques of the Old Masters
that artists started using a camera lucida
from the 1420s, resulting in a sudden
change in precision and realism, and that
this practice was continued by major
artists including Ingres, Van Eyck, and
Caravaggio.[24] Critics disagree on
whether Hockney was correct.[25][26]
Similarly, the architect Philip Steadman
argued controversially[27] that Vermeer
had used a different device, the camera
obscura, to help him create his
distinctively observed paintings.[28]

In 1509, Luca Pacioli (c. 1447–1517)


published De divina proportione on
mathematical and artistic proportion,
including in the human face. Leonardo da
Vinci (1452–1519) illustrated the text
with woodcuts of regular solids while he
studied under Pacioli in the 1490s.
Leonardo's drawings are probably the
first illustrations of skeletonic solids.[29]
These, such as the
rhombicuboctahedron, were among the
first to be drawn to demonstrate
perspective by being overlaid on top of
each other. The work discusses
perspective in the works of Piero della
Francesca, Melozzo da Forlì, and Marco
Palmezzano.[30] Da Vinci studied Pacioli's
Summa, from which he copied tables of
proportions.[31] In Mona Lisa and The
Last Supper, Da Vinci's work incorporated
linear perspective with a vanishing point
to provide apparent depth.[32] The Last
Supper is constructed in a tight ratio of
12:6:4:3, as is Raphael's The School of
Athens, which includes Pythagoras with a
tablet of ideal ratios, sacred to the
Pythagoreans.[33][34] In Vitruvian Man,
Leonardo expressed the ideas of the
Roman architect Vitruvius, innovatively
showing the male figure twice, and
centring him in both a circle and a
square.[35]

As early as the 15th century, curvilinear


perspective found its way into paintings
by artists interested in image distortions.
Jan van Eyck's 1434 Arnolfini Portrait
contains a convex mirror with reflections
of the people in the scene,[36] while
Parmigianino's Self-portrait in a Convex
Mirror, c. 1523–1524, shows the artist's
largely undistorted face at the centre,
with a strongly curved background and
artist's hand around the edge.[37]

Three-dimensional space can be


represented convincingly in art, as in
technical drawing, by means other than
perspective. Oblique projections,
including cavalier perspective (used by
French military artists to depict
fortifications in the 18th century), were
used continuously and ubiquitously by
Chinese artists from the first or second
centuries until the 18th century. The
Chinese acquired the technique from
India, which acquired it from Ancient
Rome. Oblique projection is seen in
Japanese art, such as in the Ukiyo-e
paintings of Torii Kiyonaga (1752–
1815).[38]
Woodcut from Luca Pacioli's 1509 De
divina proportione with an equilateral
triangle on a human face

Camera lucida in use. Scientific


American, 1879
Illustration of an artist using a camera
obscura. 17th century

Proportion: Leonardo's Vitruvian Man, c.


1490
Brunelleschi's theory of perspective:
Masaccio's Trinità, c. 1426–1428, in the
Basilica of Santa Maria Novella

Diagram from Leon Battista Alberti's


1435 Della Pittura, with pillars in
perspective on a grid
Linear perspective in Piero della
Francesca's Flagellation of Christ, c.
1455–1460

Curvilinear perspective: convex mirror in


Jan van Eyck's Arnolfini Portrait, 1434
Parmigianino, Self-portrait in a Convex
Mirror, c. 1523–1524

Pythagoras with tablet of ratios, in


Raphael's The School of Athens, 1509
Oblique projection: Entrance and yard of
a yamen. Detail of scroll about Suzhou by
Xu Yang, ordered by the Qianlong
Emperor. 18th century
Oblique projection: women playing Shogi,
Go and Ban-sugoroku board games.
Painting by Torii Kiyonaga, Japan, c.
1780

Golden ratio

The golden ratio (roughly equal to 1.618)


was known to Euclid.[39] The golden ratio
has persistently been claimed[40][41][42][43]
in modern times to have been used in art
and architecture by the ancients in Egypt,
Greece and elsewhere, without reliable
evidence.[44] The claim may derive from
confusion with "golden mean", which to
the Ancient Greeks meant "avoidance of
excess in either direction", not a ratio.[44]
Pyramidologists since the nineteenth
century have argued on dubious
mathematical grounds for the golden
ratio in pyramid design.[b] The Parthenon,
a 5th-century BC temple in Athens, has
been claimed to use the golden ratio in
its façade and floor plan,[47][48][49] but
these claims too are disproved by
measurement.[44] The Great Mosque of
Kairouan in Tunisia has similarly been
claimed to use the golden ratio in its
design,[50] but the ratio does not appear
in the original parts of the mosque.[51]
The historian of architecture Frederik
Macody Lund argued in 1919 that the
Cathedral of Chartres (12th century),
Notre-Dame of Laon (1157–1205) and
Notre Dame de Paris (1160) are designed
according to the golden ratio,[52] drawing
regulator lines to make his case. Other
scholars argue that until Pacioli's work in
1509, the golden ratio was unknown to
artists and architects.[53] For example,
the height and width of the front of Notre-
Dame of Laon have the ratio 8/5 or 1.6,
not 1.618. Such Fibonacci ratios quickly
become hard to distinguish from the
golden ratio.[54] After Pacioli, the golden
ratio is more definitely discernible in
artworks including Leonardo's Mona
Lisa.[55]

Another ratio, the only other morphic


number,[56] was named the plastic
number[c] in 1928 by the Dutch architect
Hans van der Laan (originally named le
nombre radiant in French).[57] Its value is
the solution of the cubic equation

an irrational number which is


approximately 1.325. According to the
architect Richard Padovan, this has
characteristic ratios 34 and 17 , which
govern the limits of human perception in
relating one physical size to another. Van
der Laan used these ratios when
designing the 1967 St. Benedictusberg
Abbey church in the Netherlands.[57]
Base:hypotenuse(b:a) ratios for the
Pyramid of Khufu could be: 1:φ (Kepler
triangle), 3:5 (3-4-5 Triangle), or 1:4/π

Supposed ratios: Notre-Dame of Laon


Golden rectangles superimposed on the
Mona Lisa

The 1967 St. Benedictusberg Abbey


church by Hans van der Laan has plastic
number proportions.

Planar symmetries
Powerful presence:[58] carpet with double medallion.
Central Anatolia (Konya – Karapınar), turn of the
16th/17th centuries. Alâeddin Mosque

Planar symmetries have for millennia


been exploited in artworks such as
carpets, lattices, textiles and
tilings.[59][60][61][62]

Many traditional rugs, whether pile


carpets or flatweave kilims, are divided
into a central field and a framing border;
both can have symmetries, though in
handwoven carpets these are often
slightly broken by small details,
variations of pattern and shifts in colour
introduced by the weaver.[59] In kilims
from Anatolia, the motifs used are
themselves usually symmetrical. The
general layout, too, is usually present,
with arrangements such as stripes,
stripes alternating with rows of motifs,
and packed arrays of roughly hexagonal
motifs. The field is commonly laid out as
a wallpaper with a wallpaper group such
as pmm, while the border may be laid out
as a frieze of frieze group pm11, pmm2
or pma2. Turkish and Central Asian
kilims often have three or more borders
in different frieze groups. Weavers
certainly had the intention of symmetry,
without explicit knowledge of its
mathematics.[59] The mathematician and
architectural theorist Nikos Salingaros
suggests that the "powerful presence"[58]
(aesthetic effect) of a "great carpet"[58]
such as the best Konya two-medallion
carpets of the 17th century is created by
mathematical techniques related to the
theories of the architect Christopher
Alexander. These techniques include
making opposites couple; opposing
colour values; differentiating areas
geometrically, whether by using
complementary shapes or balancing the
directionality of sharp angles; providing
small-scale complexity (from the knot
level upwards) and both small- and large-
scale symmetry; repeating elements at a
hierarchy of different scales (with a ratio
of about 2.7 from each level to the next).
Salingaros argues that "all successful
carpets satisfy at least nine of the above
ten rules", and suggests that it might be
possible to create a metric from these
rules.[58]

Elaborate lattices are found in Indian Jali


work, carved in marble to adorn tombs
and palaces.[60] Chinese lattices, always
with some symmetry, exist in 14 of the
17 wallpaper groups; they often have
mirror, double mirror, or rotational
symmetry. Some have a central
medallion, and some have a border in a
frieze group.[63] Many Chinese lattices
have been analysed mathematically by
Daniel S. Dye; he identifies Sichuan as
the centre of the craft.[64]

Girih tiles

Symmetries are prominent in textile arts


including quilting,[61] knitting,[65] cross-
stitch, crochet,[66] embroidery[67][68] and
weaving,[69] where they may be purely
decorative or may be marks of status.[70]
Rotational symmetry is found in circular
structures such as domes; these are
sometimes elaborately decorated with
symmetric patterns inside and out, as at
the 1619 Sheikh Lotfollah Mosque in
Isfahan.[71] Items of embroidery and lace
work such as tablecloths and table mats,
made using bobbins or by tatting, can
have a wide variety of reflectional and
rotational symmetries which are being
explored mathematically.[72]

Islamic art exploits symmetries in many


of its artforms, notably in girih tilings.
These are formed using a set of five tile
shapes, namely a regular decagon, an
elongated hexagon, a bow tie, a rhombus,
and a regular pentagon. All the sides of
these tiles have the same length; and all
their angles are multiples of 36° (π/5
radians), offering fivefold and tenfold
symmetries. The tiles are decorated with
strapwork lines (girih), generally more
visible than the tile boundaries. In 2007,
the physicists Peter Lu and Paul
Steinhardt argued that girih resembled
quasicrystalline Penrose tilings.[73]
Elaborate geometric zellige tilework is a
distinctive element in Moroccan
architecture.[62] Muqarnas vaults are
three-dimensional but were designed in
two dimensions with drawings of
geometrical cells.[74]
Hotamis kilim (detail), central Anatolia,
early 19th century

Detail of a Ming Dynasty brocade, using a


chamfered hexagonal lattice pattern
Jaali marble lattice at tomb of Salim
Chishti, Fatehpur Sikri, India

Symmetries: Florentine Bargello pattern


tapestry work
Ceiling of the Sheikh Lotfollah Mosque,
Isfahan, 1619

Rotational symmetry in lace: tatting work


Girih tiles: patterns at large and small
scales on a spandrel from the Darb-i
Imam shrine, Isfahan, 1453

Tessellations: zellige mosaic tiles at Bou


Inania Madrasa, Fes, Morocco
The complex geometry and tilings of the
muqarnas vaulting in the Sheikh
Lotfollah Mosque, Isfahan

Architect's plan of a muqarnas quarter


vault. Topkapı Scroll
Tupa Inca tunic from Peru, 1450 –1540,
an Andean textile denoting high rank[70]

Polyhedra

The first printed illustration of a


rhombicuboctahedron, by Leonardo da Vinci,
published in De Divina Proportione, 1509
The Platonic solids and other polyhedra
are a recurring theme in Western art.
They are found, for instance, in a marble
mosaic featuring the small stellated
dodecahedron, attributed to Paolo
Uccello, in the floor of the San Marco
Basilica in Venice;[13] in Leonardo da
Vinci's diagrams of regular polyhedra
drawn as illustrations for Luca Pacioli's
1509 book The Divine Proportion;[13] as a
glass rhombicuboctahedron in Jacopo
de Barbari's portrait of Pacioli, painted in
1495;[13] in the truncated polyhedron (and
various other mathematical objects) in
Albrecht Dürer's engraving Melencolia
I;[13] and in Salvador Dalí's painting The
Last Supper in which Christ and his
disciples are pictured inside a giant
dodecahedron.

Albrecht Dürer (1471–1528) was a


German Renaissance printmaker who
made important contributions to
polyhedral literature in his 1525 book,
Underweysung der Messung (Education
on Measurement), meant to teach the
subjects of linear perspective, geometry
in architecture, Platonic solids, and
regular polygons. Dürer was likely
influenced by the works of Luca Pacioli
and Piero della Francesca during his trips
to Italy.[75] While the examples of
perspective in Underweysung der
Messung are underdeveloped and
contain inaccuracies, there is a detailed
discussion of polyhedra. Dürer is also the
first to introduce in text the idea of
polyhedral nets, polyhedra unfolded to lie
flat for printing.[76] Dürer published
another influential book on human
proportions called Vier Bücher von
Menschlicher Proportion (Four Books on
Human Proportion) in 1528.[77]

Salvador Dalí, Crucifixion (Corpus Hypercubus),


1954, (with the net of an unfolded hypercube,
representing the divine perspective with four
dimensions), oil on canvas, 194.3 × 123.8 cm,
Metropolitan Museum of Art, New York[78][79]

Dürer's well-known engraving Melencolia


I depicts a frustrated thinker sitting by a
truncated triangular trapezohedron and a
magic square.[1] These two objects, and
the engraving as a whole, have been the
subject of more modern interpretation
than the contents of almost any other
print,[1][80][81] including a two-volume
book by Peter-Klaus Schuster,[82] and an
influential discussion in Erwin Panofsky's
monograph of Dürer.[1][83]Salvador Dalí's
Corpus Hypercubus depicts an unfolded
three-dimensional net for a hypercube, a
four-dimensional regular
polyhedron.[79][78][79]

Fractal dimensions

Batiks from Surakarta, Java, like this parang klithik


sword pattern, have a fractal dimension between 1.2
and 1.5.

Traditional Indonesian wax-resist batik


designs on cloth combine
representational motifs (such as floral
and vegetal elements) with abstract and
somewhat chaotic elements, including
imprecision in applying the wax resist,
and random variation introduced by
cracking of the wax. Batik designs have a
fractal dimension between 1 and 2,
varying in different regional styles. For
example, the batik of Cirebon has a
fractal dimension of 1.1; the batiks of
Yogyakarta and Surakarta (Solo) in
Central Java have a fractal dimension of
1.2 to 1.5; and the batiks of Lasem on the
north coast of Java and of Tasikmalaya
in West Java have a fractal dimension
between 1.5 and 1.7.[84]

The drip painting works of the modern


artist Jackson Pollock are similarly
distinctive in their fractal dimension. His
1948 Number 14 has a coastline-like
dimension of 1.45, while his later
paintings had successively higher fractal
dimensions and accordingly more
elaborate patterns. One of his last works,
Blue Poles, took six months to create,
and has the fractal dimension of 1.72.[85]

A complex relationship
The astronomer Galileo Galilei in his Il
Saggiatore wrote that "[The universe] is
written in the language of mathematics,
and its characters are triangles, circles,
and other geometric figures."[86] Artists
who strive and seek to study nature must
first, in Galileo's view, fully understand
mathematics. Mathematicians,
conversely, have sought to interpret and
analyse art through the lens of geometry
and rationality. The mathematician Felipe
Cucker suggests that mathematics, and
especially geometry, is a source of rules
for "rule-driven artistic creation", though
not the only one.[87] Some of the many
strands of the resulting complex
relationship[88] are described below.

The mathematician G. H. Hardy defined a set of


criteria for mathematical beauty.
Mathematics as an art

The mathematician Jerry P. King


describes mathematics as an art, stating
that "the keys to mathematics are beauty
and elegance and not dullness and
technicality", and that beauty is the
motivating force for mathematical
research.[89] King cites the
mathematician G. H. Hardy's 1940 essay
A Mathematician's Apology. In it, Hardy
discusses why he finds two theorems of
classical times as first rate, namely
Euclid's proof there are infinitely many
prime numbers, and the proof that the
square root of 2 is irrational. King
evaluates this last against Hardy's
criteria for mathematical elegance:
"seriousness, depth, generality,
unexpectedness, inevitability, and
economy" (King's italics), and describes
the proof as "aesthetically pleasing".[90]
The Hungarian mathematician Paul
Erdős agreed that mathematics
possessed beauty but considered the
reasons beyond explanation: "Why are
numbers beautiful? It's like asking why is
Beethoven's Ninth Symphony beautiful. If
you don't see why, someone can't tell you.
I know numbers are beautiful."[91]

Mathematical tools for art


Mathematics can be discerned in many
of the arts, such as music, dance,[92]
painting, architecture, and sculpture.
Each of these is richly associated with
mathematics.[93] Among the connections
to the visual arts, mathematics can
provide tools for artists, such as the rules
of linear perspective as described by
Brook Taylor and Johann Lambert, or the
methods of descriptive geometry, now
applied in software modelling of solids,
dating back to Albrecht Dürer and
Gaspard Monge.[94] Artists from Luca
Pacioli in the Middle Ages and Leonardo
da Vinci and Albrecht Dürer in the
Renaissance have made use of and
developed mathematical ideas in the
pursuit of their artistic work.[93][95] The
use of perspective began, despite some
embryonic usages in the architecture of
Ancient Greece, with Italian painters such
as Giotto in the 13th century; rules such
as the vanishing point were first
formulated by Brunelleschi in about
1413,[7] his theory influencing Leonardo
and Dürer. Isaac Newton's work on the
optical spectrum influenced Goethe's
Theory of Colours and in turn artists such
as Philipp Otto Runge, J. M. W. Turner,[96]
the Pre-Raphaelites and Wassily
Kandinsky.[97][98] Artists may also choose
to analyse the symmetry of a scene.[99]
Tools may be applied by mathematicians
who are exploring art, or artists inspired
by mathematics, such as M. C. Escher
(inspired by H. S. M. Coxeter) and the
architect Frank Gehry, who more
tenuously argued that computer aided
design enabled him to express himself in
a wholly new way.[100]

Octopod by Mikael Hvidtfeldt Christensen.


Algorithmic art produced with the software Structure
Synth

The artist Richard Wright argues that


mathematical objects that can be
constructed can be seen either "as
processes to simulate phenomena" or as
works of "computer art". He considers
the nature of mathematical thought,
observing that fractals were known to
mathematicians for a century before they
were recognised as such. Wright
concludes by stating that it is appropriate
to subject mathematical objects to any
methods used to "come to terms with
cultural artifacts like art, the tension
between objectivity and subjectivity, their
metaphorical meanings and the
character of representational systems."
He gives as instances an image from the
Mandelbrot set, an image generated by a
cellular automaton algorithm, and a
computer-rendered image, and
discusses, with reference to the Turing
test, whether algorithmic products can
be art.[101] Sasho Kalajdzievski's Math
and Art: An Introduction to Visual
Mathematics takes a similar approach,
looking at suitably visual mathematics
topics such as tilings, fractals and
hyperbolic geometry.[102]

Some of the first works of computer art


were created by Desmond Paul Henry's
"Drawing Machine 1", an analogue
machine based on a bombsight
computer and exhibited in 1962.[103][104]
The machine was capable of creating
complex, abstract, asymmetrical,
curvilinear, but repetitive line
drawings.[103][105] More recently, Hamid
Naderi Yeganeh has created shapes
suggestive of real world objects such as
fish and birds, using formulae that are
successively varied to draw families of
curves or angled lines.[106][107][108] Artists
such as Mikael Hvidtfeldt Christensen
create works of generative or algorithmic
art by writing scripts for a software
system such as Structure Synth: the artist
effectively directs the system to apply a
desired combination of mathematical
operations to a chosen set of
data.[109][110]
Mathematical sculpture by Bathsheba
Grossman, 2007

Fractal sculpture: 3D Fraktal 03/H/dd by


Hartmut Skerbisch, 2003
Fibonacci word: detail of artwork by
Samuel Monnier, 2009

Computer art image produced by


Desmond Paul Henry's "Drawing Machine
1", exhibited 1962
A Bird in Flight, by Hamid Naderi
Yeganeh, 2016, constructed with a family
of mathematical curves.

From mathematics to art

Proto-Cubism: Pablo Picasso's 1907 painting Les


Demoiselles d'Avignon uses a fourth dimension
projection to show a figure both full face and in
profile.[111]
The mathematician and theoretical
physicist Henri Poincaré's Science and
Hypothesis was widely read by the
Cubists, including Pablo Picasso and
Jean Metzinger.[112][113] Poincaré viewed
Euclidean geometry as just one of many
possible geometric configurations, rather
than as an absolute objective truth. The
possible existence of a fourth dimension
inspired artists to question classical
Renaissance perspective: non-Euclidean
geometry became a valid
alternative.[114][115][116] The concept that
painting could be expressed
mathematically, in colour and form,
contributed to Cubism, the art movement
that led to abstract art.[117] Metzinger, in
1910, wrote that: "[Picasso] lays out a
free, mobile perspective, from which that
ingenious mathematician Maurice
Princet has deduced a whole
geometry".[118] Later, Metzinger wrote in
his memoirs:

Maurice Princet joined us


often ... it was as an artist that
he conceptualized
mathematics, as an
aesthetician that he invoked n-
dimensional continuums. He
loved to get the artists
interested in the new views on
space that had been opened up
by Schlegel and some others.
He succeeded at that.[119]

The impulse to make teaching or


research models of mathematical forms
naturally creates objects that have
symmetries and surprising or pleasing
shapes. Some of these have inspired
artists such as the Dadaists Man Ray,[120]
Marcel Duchamp[121] and Max
Ernst,[122][123] and following Man Ray,
Hiroshi Sugimoto.[124]
Enneper surfaces as Dadaism: Man Ray's 1934 Objet

mathematique

Man Ray photographed some of the


mathematical models in the Institut Henri
Poincaré in Paris, including Objet
mathematique (Mathematical object). He
noted that this represented Enneper
surfaces with constant negative
curvature, derived from the pseudo-
sphere. This mathematical foundation
was important to him, as it allowed him
to deny that the object was "abstract",
instead claiming that it was as real as the
urinal that Duchamp made into a work of
art. Man Ray admitted that the object's
[Enneper surface] formula "meant
nothing to me, but the forms themselves
were as varied and authentic as any in
nature." He used his photographs of the
mathematical models as figures in his
series he did on Shakespeare's plays,
such as his 1934 painting Antony and
Cleopatra.[125] The art reporter Jonathan
Keats, writing in ForbesLife, argues that
Man Ray photographed "the elliptic
paraboloids and conic points in the same
sensual light as his pictures of Kiki de
Montparnasse", and "ingeniously
repurposes the cool calculations of
mathematics to reveal the topology of
desire".[126] Twentieth century sculptors
such as Henry Moore, Barbara Hepworth
and Naum Gabo took inspiration from
mathematical models.[127] Moore wrote
of his 1938 Stringed Mother and Child:
"Undoubtedly the source of my stringed
figures was the Science Museum ... I was
fascinated by the mathematical models I
saw there ... It wasn't the scientific study
of these models but the ability to look
through the strings as with a bird cage
and to see one form within another which
excited me."[128]
Theo van Doesburg's Six Moments in the
Development of Plane to Space, 1926 or 1929

The artists Theo van Doesburg and Piet


Mondrian founded the De Stijl movement,
which they wanted to "establish a visual
vocabulary comprised of elementary
geometrical forms comprehensible by all
and adaptable to any discipline".[129][130]
Many of their artworks visibly consist of
ruled squares and triangles, sometimes
also with circles. De Stijl artists worked in
painting, furniture, interior design and
architecture.[129] After the breakup of De
Stijl, Van Doesburg founded the Avant-
garde Art Concret movement, describing
his 1929–1930 Arithmetic Composition ,
a series of four black squares on the
diagonal of a squared background, as "a
structure that can be controlled, a definite
surface without chance elements or
individual caprice", yet "not lacking in
spirit, not lacking the universal and not ...
empty as there is everything which fits
the internal rhythm". The art critic Gladys
Fabre observes that two progressions
are at work in the painting, namely the
growing black squares and the
alternating backgrounds.[131]

The mathematics of tessellation,


polyhedra, shaping of space, and self-
reference provided the graphic artist M.
C. Escher (1898—1972) with a lifetime's
worth of materials for his
woodcuts.[132][133] In the Alhambra
Sketch, Escher showed that art can be
created with polygons or regular shapes
such as triangles, squares, and
hexagons. Escher used irregular
polygons when tiling the plane and often
used reflections, glide reflections, and
translations to obtain further patterns.
Many of his works contain impossible
constructions, made using geometrical
objects which set up a contradiction
between perspective projection and three
dimensions, but are pleasant to the
human sight. Escher's Ascending and
Descending is based on the "impossible
staircase" created by the medical
scientist Lionel Penrose and his son the
mathematician Roger
Penrose.[134][135][136]

Some of Escher's many tessellation


drawings were inspired by conversations
with the mathematician H. S. M. Coxeter
on hyperbolic geometry.[137] Escher was
especially interested in five specific
polyhedra, which appear many times in
his work. The Platonic solids—
tetrahedrons, cubes, octahedrons,
dodecahedrons, and icosahedrons—are
especially prominent in Order and Chaos
and Four Regular Solids.[138] These
stellated figures often reside within
another figure which further distorts the
viewing angle and conformation of the
polyhedrons and provides a multifaceted
perspective artwork.[139]

The visual intricacy of mathematical


structures such as tessellations and
polyhedra have inspired a variety of
mathematical artworks. Stewart Coffin
makes polyhedral puzzles in rare and
beautiful woods; George W. Hart works
on the theory of polyhedra and sculpts
objects inspired by them; Magnus
Wenninger makes "especially beautiful"
models of complex stellated
polyhedra.[140]

The distorted perspectives of


anamorphosis have been explored in art
since the sixteenth century, when Hans
Holbein the Younger incorporated a
severely distorted skull in his 1533
painting The Ambassadors. Many artists
since then, including Escher, have make
use of anamorphic tricks.[141]
The mathematics of topology has
inspired several artists in modern times.
The sculptor John Robinson (1935–
2007) created works such as Gordian
Knot and Bands of Friendship, displaying
knot theory in polished bronze.[8] Other
works by Robinson explore the topology
of toruses. Genesis is based on
Borromean rings – a set of three circles,
no two of which link but in which the
whole structure cannot be taken apart
without breaking.[142] The sculptor
Helaman Ferguson creates complex
surfaces and other topological
objects.[143] His works are visual
representations of mathematical objects;
The Eightfold Way is based on the
projective special linear group PSL(2,7), a
finite group of 168 elements.[144][145] The
sculptor Bathsheba Grossman similarly
bases her work on mathematical
structures.[146][147]

A liberal arts inquiry project examines


connections between mathematics and
art through the Möbius strip, flexagons,
origami and panorama photography.[148]

Mathematical objects including the


Lorenz manifold and the hyperbolic plane
have been crafted using fiber arts
including crochet.[d][150][151] The
American weaver Ada Dietz wrote a 1949
monograph Algebraic Expressions in
Handwoven Textiles, defining weaving
patterns based on the expansion of
multivariate polynomials.[152] The
mathematician J. C. P. Miller used the
Rule 90 cellular automaton to design
tapestries depicting both trees and
abstract patterns of triangles.[153] The
"mathekniticians"[154] Pat Ashforth and
Steve Plummer use knitted versions of
mathematical objects such as
hexaflexagons in their teaching, though
their Menger sponge proved too
troublesome to knit and was made of
plastic canvas instead.[155][156] Their
"mathghans" (Afghans for Schools)
project introduced knitting into the British
mathematics and technology
curriculum.[157][158]
Four-dimensional space to Cubism:
Esprit Jouffret's 1903 Traité élémentaire
de géométrie à quatre dimensions.[159][e]

De Stijl: Theo van Doesburg's geometric


Composition I (Still Life), 1916
Pedagogy to art: Magnus Wenninger with
some of his stellated polyhedra, 2009

A Möbius strip scarf in crochet, 2007


Anamorphism: The Ambassadors by
Hans Holbein the Younger, 1533, with
severely distorted skull in foreground

Semiotic joke: René Magritte's La condition humaine


1933

Illustrating mathematics
Front face of Giotto's Stefaneschi Triptych, 1320
illustrates recursion.

Detail of Cardinal Stefaneschi holding the triptych

Modelling is far from the only possible


way to illustrate mathematical concepts.
Giotto's Stefaneschi Triptych, 1320,
illustrates recursion in the form of mise
en abyme; the central panel of the
triptych contains, lower left, the kneeling
figure of Cardinal Stefaneschi, holding up
the triptych as an offering.[161] Giorgio
Chirico's metaphysical paintings such as
his 1917 Great Metaphysical Interior
explore the question of levels of
representation in art by depicting
paintings within his paintings.[162]

Art can exemplify logical paradoxes, as in


some paintings by the surrealist René
Magritte, which can be read as semiotic
jokes about confusion between levels. In
La condition humaine (1933), Magritte
depicts an easel (on the real canvas),
seamlessly supporting a view through a
window which is framed by "real"
curtains in the painting. Similarly,
Escher's Print Gallery (1956) is a print
which depicts a distorted city which
contains a gallery which recursively
contains the picture, and so ad
infinitum.[163] Magritte made use of
spheres and cuboids to distort reality in a
different way, painting them alongside an
assortment of houses in his 1931 Mental
Arithmetic as if they were children's
building blocks, but house-sized.[164] The
Guardian observed that the "eerie
toytown image" prophesied Modernism's
usurpation of "cosy traditional forms", but
also plays with the human tendency to
seek patterns in nature.[165]

Diagram of the apparent paradox embodied in M. C.


Escher's 1956 lithograph Print Gallery, as discussed
by Douglas Hofstadter in his 1980 book Gödel,
Escher, Bach

Salvador Dalí's last painting, The


Swallow's Tail (1983), was part of a
series inspired by René Thom's
catastrophe theory.[166] The Spanish
painter and sculptor Pablo Palazuelo
(1916–2007) focused on the
investigation of form. He developed a
style that he described as the geometry
of life and the geometry of all nature.
Consisting of simple geometric shapes
with detailed patterning and coloring, in
works such as Angular I and Automnes,
Palazuelo expressed himself in
geometric transformations.[8]

The artist Adrian Gray practises stone


balancing, exploiting friction and the
centre of gravity to create striking and
seemingly impossible compositions.[167]

Lithograph Print Gallery by M. C. Escher, 1956


Artists, however, do not necessarily take
geometry literally. As Douglas Hofstadter
writes in his 1980 reflection on human
thought, Gödel, Escher, Bach, by way of
(among other things) the mathematics of
art: "The difference between an Escher
drawing and non-Euclidean geometry is
that in the latter, comprehensible
interpretations can be found for the
undefined terms, resulting in a
comprehensible total system, whereas
for the former, the end result is not
reconcilable with one's conception of the
world, no matter how long one stares at
the pictures." Hofstadter discusses the
seemingly paradoxical lithograph Print
Gallery by M. C. Escher; it depicts a
seaside town containing an art gallery
which seems to contain a painting of the
seaside town, there being a "strange loop,
or tangled hierarchy" to the levels of
reality in the image. The artist himself,
Hofstadter observes, is not seen; his
reality and his relation to the lithograph
are not paradoxical.[168] The image's
central void has also attracted the
interest of mathematicians Bart de Smit
and Hendrik Lenstra, who propose that it
could contain a Droste effect copy of
itself, rotated and shrunk; this would be a
further illustration of recursion beyond
that noted by Hofstadter.[169][170]
Analysis of art history

Algorithmic analysis of images of


artworks, for example using X-ray
fluorescence spectroscopy, can reveal
information about art. Such techniques
can uncover images in layers of paint
later covered over by an artist; help art
historians to visualize an artwork before
it cracked or faded; help to tell a copy
from an original, or distinguish the
brushstroke style of a master from those
of his apprentices.[171][172]
Max Ernst making Lissajous figures, New York, 1942

Jackson Pollock's drip painting style[173]


has a definite fractal dimension;[174]
among the artists who may have
influenced Pollock's controlled chaos,[175]
Max Ernst painted Lissajous figures
directly by swinging a punctured bucket
of paint over a canvas.[176]

The computer scientist Neil Dodgson


investigated whether Bridget Riley's
stripe paintings could be characterised
mathematically, concluding that while
separation distance could "provide some
characterisation" and global entropy
worked on some paintings,
autocorrelation failed as Riley's patterns
were irregular. Local entropy worked
best, and correlated well with the
description given by the art critic Robert
Kudielka.[177]

The American mathematician George


Birkhoff's 1933 Aesthetic Measure
proposes a quantitative metric of the
aesthetic quality of an artwork. It does
not attempt to measure the connotations
of a work, such as what a painting might
mean, but is limited to the "elements of
order" of a polygonal figure. Birkhoff first
combines (as a sum) five such elements:
whether there is a vertical axis of
symmetry; whether there is optical
equilibrium; how many rotational
symmetries it has; how wallpaper-like the
figure is; and whether there are
unsatisfactory features such as having
two vertices too close together. This
metric, O, takes a value between −3 and
7. The second metric, C, counts elements
of the figure, which for a polygon is the
number of different straight lines
containing at least one of its sides.
Birkhoff then defines his aesthetic
measure of an object's beauty as O/C.
This can be interpreted as a balance
between the pleasure looking at the
object gives, and the amount of effort
needed to take it in. Birkhoff's proposal
has been criticized in various ways, not
least for trying to put beauty in a formula,
but he never claimed to have done
that.[178]

Stimuli to mathematical
research

Art has sometimes stimulated the


development of mathematics, as when
Brunelleschi's theory of perspective in
architecture and painting started a cycle
of research that led to the work of Brook
Taylor and Johann Heinrich Lambert on
the mathematical foundations of
perspective drawing,[179] and ultimately
to the mathematics of projective
geometry of Girard Desargues and Jean-
Victor Poncelet.[180]

The Japanese paper-folding art of


origami has been reworked
mathematically by Tomoko Fusé using
modules, congruent pieces of paper such
as squares, and making them into
polyhedra or tilings.[181] Paper-folding
was used in 1893 by T. Sundara Rao in
his Geometric Exercises in Paper Folding
to demonstrate geometrical proofs.[182]
The mathematics of paper folding has
been explored in Maekawa's theorem,[183]
Kawasaki's theorem,[184] and the Huzita–
Hatori axioms.[185]
Stimulus to projective geometry: Alberti's
diagram showing a circle seen in
perspective as an ellipse. Della Pittura,
1435–6

Mathematical origami: Spring Into Action,


by Jeff Beynon, made from a single
paper rectangle.[186]
Illusion to Op art

The Fraser spiral illusion, named for Sir James


Fraser who discovered it in 1908.

Optical illusions such as the Fraser spiral


strikingly demonstrate limitations in
human visual perception, creating what
the art historian Ernst Gombrich called a
"baffling trick." The black and white ropes
that appear to form spirals are in fact
concentric circles. The mid-twentieth
century Op art or optical art style of
painting and graphics exploited such
effects to create the impression of
movement and flashing or vibrating
patterns seen in the work of artists such
as Bridget Riley, Spyros Horemis,[187] and
Victor Vasarely.[188]

Sacred geometry

A strand of art from Ancient Greece


onwards sees God as the geometer of
the world, and the world's geometry
therefore as sacred. The belief that God
created the universe according to a
geometric plan has ancient origins.
Plutarch attributed the belief to Plato,
writing that "Plato said God geometrizes
continually" (Convivialium disputationum,
liber 8,2). This image has influenced
Western thought ever since. The Platonic
concept derived in its turn from a
Pythagorean notion of harmony in music,
where the notes were spaced in perfect
proportions, corresponding to the lengths
of the lyre's strings; indeed, the
Pythagoreans held that everything was
arranged by Number. In the same way, in
Platonic thought, the regular or Platonic
solids dictate the proportions found in
nature, and in art.[189][190] A Mediaeval
manuscript illustration may refer to a
verse in the Old Testament: "When he
established the heavens I was there:
when he set a compass upon the face of
the deep" (Proverbs 8:27), showing God
drawing out the universe with a pair of
compasses.[191] In 1596, the
mathematical astronomer Johannes
Kepler modelled the universe as a set of
nested Platonic solids, determining the
relative sizes of the orbits of the
planets.[191] William Blake's Ancient of
Days and his painting of the physicist
Isaac Newton, naked and drawing with a
compass, attempt to depict the contrast
between the mathematically perfect
spiritual world and the imperfect physical
world,[192] as in a different way does
Salvador Dalí's 1954 Crucifixion (Corpus
Hypercubus), which depicts the cross as
a hypercube, representing the divine
perspective with four dimensions rather
than the usual three.[79] In Dali's The
Sacrament of the Last Supper (1955)
Christ and his disciples are pictured
inside a giant dodecahedron.[193]
God the geometer. Codex Vindobonensis,
c. 1220

Johannes Kepler's Platonic solid model


of planetary spacing in the solar system
from Mysterium Cosmographicum, 1596
William Blake's The Ancient of Days, 1794

William Blake's Newton, taking God's


place as geometer, c. 1800

See also
Mathematics and architecture
Music and mathematics

Notes
a. In Piero's Italian: "una cosa tanto
picholina quanto e possible ad ochio
comprendere".
b. The ratio of the slant height to half
the base length is 1.619, less than
1% from the golden ratio, implying
use of Kepler's triangle (face angle
51°49').[44][45] It is more likely that
pyramids were made with the 3-4-5
triangle (face angle 53°8'), known
from the Rhind Mathematical
Papyrus; or with the triangle with
base to hypotenuse ratio 1:4/π (face
angle 51°50').[46]
c. 'Plastic' named the ability to take on
a chosen three-dimensional shape.
d. Images and videos of Hinke Osinga's
crocheted Lorenz manifold reached
international television news, as can
be seen in the linked website.[149]
e. Maurice Princet gave a copy to
Pablo Picasso, whose sketchbooks
for Les Demoiselles d'Avignon
illustrate Jouffret's influence.[112][160]

References
1. Ziegler, Günter M. (3 December
2014). "Dürer's polyhedron: 5
theories that explain Melencolia's
crazy cube" . The Guardian.
Retrieved 27 October 2015.
2. Colombo, C.; Del Bimbo, A.; Pernici,
F. (2005). "Metric 3D reconstruction
and texture acquisition of surfaces
of revolution from a single
uncalibrated view". IEEE
Transactions on Pattern Analysis
and Machine Intelligence. 27 (1):
99–114. CiteSeerX 10.1.1.58.8477 .
doi:10.1109/TPAMI.2005.14 .
PMID 15628272 .
3. Stewart, Andrew (November 1978).
"Polykleitos of Argos," One Hundred
Greek Sculptors: Their Careers and
Extant Works". Journal of Hellenic
Studies. 98: 122–131.
doi:10.2307/630196 .
JSTOR 630196 .
4. Tobin, Richard (October 1975). "The
Canon of Polykleitos". American
Journal of Archaeology. 79 (4): 307–
321. doi:10.2307/503064 .
JSTOR 503064 .
5. Lawton, Arthur J. (2013). "Pattern,
Tradition and Innovation in
Vernacular Architecture" . Past. 36.
Retrieved 25 June 2015. "The base
figure is a square the length and
width of the distal phalange of the
little finger. Its diagonals rotated to
one side transform the square to a
1 : √2 root rectangle. In Figure 5 this
rectangular figure marks the width
and length of the adjacent medial
phalange. Rotating the medial
diagonal proportions the proximal
phalange and similarly from there to
the wrist, from wrist to elbow and
from elbow to shoulder top. Each
new step advances the diagonal's
pivot point."
6. Raven, J. E. (1951). "Polyclitus and
Pythagoreanism". Classical
Quarterly. 1 (3–4): 147–.
doi:10.1017/s0009838800004122 .
7. O'Connor, J. J.; Robertson, E. F.
(January 2003). "Mathematics and
art – perspective" . University of St
Andrews. Retrieved 1 September
2015.
8. Emmer, Michelle, ed. (2005). The
Visual Mind II. MIT Press. ISBN 978-
0-262-05048-7.
9. Vasari, Giorgio (1550). Lives of the
Artists. Torrentino. p. Chapter on
Brunelleschi.
10. Alberti, Leon Battista; Spencer, John
R. (1956) [1435]. On Painting . Yale
University Press.
11. Field, J. V. (1997). The Invention of
Infinity: Mathematics and Art in the
Renaissance. Oxford University
Press. ISBN 978-0-19-852394-9.
12. Witcombe, Christopher L. C. E. "Art
History Resources" . Retrieved
5 September 2015.
13. Hart, George W. "Polyhedra in Art" .
Retrieved 24 June 2015.
14. Cunningham, Lawrence; Reich, John;
Fichner-Rathus, Lois (1 January
2014). Culture and Values: A Survey
of the Western Humanities .
Cengage Learning. p. 375. ISBN 978-
1-285-44932-6. "which illustrate
Uccello's fascination with
perspective. The jousting
combatants engage on a battlefield
littered with broken lances that have
fallen in a near-grid pattern and are
aimed toward a vanishing point
somewhere in the distance."
15. della Francesca, Piero (1942) [c.
1474]. G. Nicco Fasola (ed.). De
Prospectiva Pingendi. Florence.
16. della Francesca, Piero (1970)
[Fifteenth century]. G. Arrighi (ed.).
Trattato d'Abaco. Pisa.
17. della Francesca, Piero (1916). G.
Mancini (ed.). L'opera "De corporibus
regularibus" di Pietro Franceschi
detto della Francesca usurpata da
Fra Luca Pacioli.
18. Vasari, G. (1878). G. Milanesi (ed.).
Le Opere, volume 2. p. 490.
19. Zuffi, Stefano (1991). Piero della
Francesca. L'Unità – Mondadori Arte.
p. 53.
20. Heath, T.L. (1908). The Thirteen
Books of Euclid's Elements.
Cambridge University Press. p. 97.
21. Grendler, P. (1995). M.A. Lavin (ed.).
What Piero Learned in School:
Fifteenth-Century Vernacular
Education. Piero della Francesca and
His Legacy. University Press of New
England. pp. 161–176.
22. Alberti, Leon Battista; Grayson, Cecil
(trans.) (1991). Kemp, Martin (ed.).
On Painting. Penguin Classics.
23. Peterson, Mark. "The Geometry of
Piero della Francesca" . "In Book I,
after some elementary constructions
to introduce the idea of the apparent
size of an object being actually its
angle subtended at the eye, and
referring to Euclid's Elements Books I
and VI, and Euclid's Optics, he turns,
in Proposition 13, to the
representation of a square lying flat
on the ground in front of the viewer.
What should the artist actually draw?
After this, objects are constructed in
the square (tilings, for example, to
represent a tiled floor), and
corresponding objects are
constructed in perspective; in Book II
prisms are erected over these planar
objects, to represent houses,
columns, etc.; but the basis of the
method is the original square, from
which everything else follows."
24. Hockney, David (2006). Secret
Knowledge: Rediscovering the Lost
Techniques of the Old Masters.
Thames and Hudson. ISBN 978-0-
500-28638-8.
25. Van Riper, Frank. "Hockney's 'Lucid'
Bomb At the Art Establishment" .
The Washington Post. Retrieved
4 September 2015.
26. Marr, Andrew (7 October 2001).
"What the eye didn't see" . The
Guardian. Retrieved 4 September
2015.
27. Janson, Jonathan (25 April 2003).
"An Interview with Philip Steadman" .
Essential Vermeer. Retrieved
5 September 2015.
28. Steadman, Philip (2002). Vermeer's
Camera: Uncovering the Truth
Behind the Masterpieces. Oxford.
ISBN 978-0-19-280302-3.
29. Hart, George. "Luca Pacioli's
Polyhedra" . Retrieved 13 August
2009.
30. Morris, Roderick Conway (27
January 2006). "Palmezzano's
Renaissance:From shadows, painter
emerges" . New York Times.
Retrieved 22 July 2015.
31. Calter, Paul. "Geometry and Art Unit
1" . Dartmouth College. Retrieved
13 August 2009.
32. Brizio, Anna Maria (1980). Leonardo
the Artist. McGraw-Hill.
33. Ladwein, Michael (2006). Leonardo
Da Vinci, the Last Supper: A Cosmic
Drama and an Act of Redemption .
Temple Lodge Publishing. pp. 61–
62. ISBN 978-1-902636-75-7.
34. Turner, Richard A. (1992). Inventing
Leonardo. Alfred A. Knopf.
35. Wolchover, Natalie (31 January
2012). "Did Leonardo da Vinci copy
his famous 'Vitruvian Man'?" . NBC
News. Retrieved 27 October 2015.
36. Criminisi, A.; Kempz, M.; Kang, S. B.
(2004). "Reflections of Reality in Jan
van Eyck and Robert Campin" (PDF).
Historical Methods. 37 (3): 109–121.
doi:10.3200/hmts.37.3.109-122 .
37. Cucker, Felix (2013). Manifold
Mirrors: The Crossing Paths of the
Arts and Mathematics. Cambridge
University Press. pp. 299–300, 306–
307. ISBN 978-0-521-72876-8.
38. Cucker, Felix (2013). Manifold
Mirrors: The Crossing Paths of the
Arts and Mathematics. Cambridge
University Press. pp. 269–278.
ISBN 978-0-521-72876-8.
39. Joyce, David E. (1996). "Euclid's
Elements, Book II, Proposition 11" .
Clark University. Retrieved
24 September 2015.
40. Seghers, M. J.; Longacre, J. J.;
Destefano, G. A. (1964). "The Golden
Proportion and Beauty". Plastic and
Reconstructive Surgery. 34 (4): 382–
386. doi:10.1097/00006534-
196410000-00007 .
41. Mainzer, Klaus (1996). Symmetries
of Nature: A Handbook for
Philosophy of Nature and Science.
Walter de Gruyter. p. 118.
42. "Mathematical properties in ancient
theatres and amphitheatres" .
Retrieved 29 January 2014.
43. "Architecture: Ellipse?" . The-
Colosseum.net. Retrieved
29 January 2014.
44. Markowsky, George (January 1992).
"Misconceptions about the Golden
Ratio" (PDF). The College
Mathematics Journal. 23 (1): 2–19.
doi:10.2307/2686193 .
JSTOR 2686193 . Archived from the
original (PDF) on 2008-04-08.
Retrieved 2015-06-26.
45. Taseos, Socrates G. (1990). Back in
Time 3104 B.C. to the Great Pyramid.
SOC Publishers.
46. Gazale, Midhat (1999). Gnomon:
From Pharaohs to Fractals. European
Journal of Physics. 20. Princeton
University Press. p. 523.
Bibcode:1999EJPh...20..523G .
ISBN 978-0-691-00514-0.
47. Huntley, H.E. (1970). The Divine
Proportion. Dover.
48. Hemenway, Priya (2005). Divine
Proportion: Phi In Art, Nature, and
Science. Sterling. p. 96.
49. Usvat, Liliana. "Mathematics of the
Parthenon" . Mathematics
Magazine. Retrieved 24 June 2015.
50. Boussora, Kenza; Mazouz, Said
(Spring 2004). "The Use of the
Golden Section in the Great Mosque
of Kairouan" . Nexus Network
Journal. 6 (1): 7–16.
doi:10.1007/s00004-004-0002-y .
Archived from the original on 2008-
10-04. "The geometric technique of
construction of the golden section
seems to have determined the major
decisions of the spatial organisation.
The golden section appears
repeatedly in some part of the
building measurements. It is found
in the overall proportion of the plan
and in the dimensioning of the prayer
space, the court and the minaret.
The existence of the golden section
in some parts of Kairouan mosque
indicates that the elements designed
and generated with this principle
may have been realised at the same
period."
51. Brinkworth, Peter; Scott, Paul (2001).
"The Place of Mathematics".
Australian Mathematics Teacher. 57
(3): 2.
52. Chanfón Olmos, Carlos (1991).
Curso sobre Proporción.
Procedimientos reguladors en
construcción. Convenio de
intercambio Unam–Uady. México –
Mérica.
53. Livio, Mario (2002). The Golden
Ratio: The Story of Phi, The World's
Most Astonishing Number. The
Golden Ratio : The Story of Phi.
Bibcode:2002grsp.book.....L .
54. Smith, Norman A. F. (2001).
"Cathedral Studies: Engineering or
History" (PDF). Transactions of the
Newcomen Society. 73: 95–137.
doi:10.1179/tns.2001.005 . Archived
from the original (PDF) on 2015-12-
11.
55. McVeigh, Karen (28 December
2009). "Why golden ratio pleases the
eye: US academic says he knows art
secret" . The Guardian. Retrieved
27 October 2015.
56. Aarts, J.; Fokkink, R.; Kruijtzer, G.
(2001). "Morphic numbers" (PDF).
Nieuw Arch. Wiskd. 5. 2 (1): 56–58.
57. Padovan, Richard (2002). Williams,
Kim; Francisco Rodrigues, Jose
(eds.). "Dom Hans Van Der Laan And
The Plastic Number" . Nexus IV:
Architecture and Mathematics: 181–
193.
58. Salingaros, Nikos (November 1996).
"The 'life' of a carpet: an application
of the Alexander rules" . 8th
International Conference on Oriental
Carpets. Reprinted in Eiland, M.;
Pinner, M., eds. (1998). Oriental
Carpet and Textile Studies V.
Danville, CA: Conference on Oriental
Carpets.
59. Cucker, Felix (2013). Manifold
Mirrors: The Crossing Paths of the
Arts and Mathematics. Cambridge
University Press. pp. 89–102.
ISBN 978-0-521-72876-8.
60. Lerner, Martin (1984). The flame and
the lotus : Indian and Southeast
Asian art from the Kronos
collections (Exhibition Catalogue
ed.). Metropolitan Museum of Art.
61. Ellison, Elaine; Venters, Diana (1999).
Mathematical Quilts: No Sewing
Required. Key Curriculum.
62. Castera, Jean Marc; Peuriot,
Francoise (1999). Arabesques.
Decorative Art in Morocco. Art
Creation Realisation. ISBN 978-2-
86770-124-5.
63. Cucker, Felix (2013). Manifold
Mirrors: The Crossing Paths of the
Arts and Mathematics. Cambridge
University Press. pp. 103–106.
ISBN 978-0-521-72876-8.
64. Dye, Daniel S. (1974). Chinese
Lattice Designs. Dover. pp. 30–39.
65. belcastro, sarah-marie (2013).
"Adventures in Mathematical
Knitting" . American Scientist. 101
(2): 124. doi:10.1511/2013.101.124 .
66. Taimina, Daina (2009). Crocheting
Adventures with Hyperbolic Planes.
A K Peters. ISBN 978-1-56881-452-0.
67. Snook, Barbara. Florentine
Embroidery. Scribner, Second edition
1967.
68. Williams, Elsa S. Bargello: Florentine
Canvas Work. Van Nostrand
Reinhold, 1967.
69. Grünbaum, Branko; Shephard,
Geoffrey C. (May 1980). "Satins and
Twills: An Introduction to the
Geometry of Fabrics". Mathematics
Magazine. 53 (3): 139–161.
Bibcode:1975MathM..48...12G .
doi:10.2307/2690105 .
JSTOR 2690105 .
70. Gamwell, Lynn (2015). Mathematics
and Art: A Cultural History. Princeton
University Press. p. 423. ISBN 978-0-
691-16528-8.
71. Baker, Patricia L.; Smith, Hilary
(2009). Iran (3 ed.). Bradt Travel
Guides. p. 107. ISBN 978-1-84162-
289-7.
72. Irvine, Veronika; Ruskey, Frank
(2014). "Developing a Mathematical
Model for Bobbin Lace". Journal of
Mathematics and the Arts. 8 (3–4):
95–110. arXiv:1406.1532 .
doi:10.1080/17513472.2014.98293
8.
73. Lu, Peter J.; Steinhardt, Paul J.
(2007). "Decagonal and Quasi-
crystalline Tilings in Medieval
Islamic Architecture". Science. 315
(5815): 1106–1110.
Bibcode:2007Sci...315.1106L .
doi:10.1126/science.1135491 .
PMID 17322056 .
74. van den Hoeven, Saskia; van der
Veen, Maartje. "Muqarnas-
Mathematics in Islamic Arts" (PDF).
Archived from the original (PDF) on
27 September 2013. Retrieved
15 January 2016.
75. Panofsky, E. (1955). The Life and Art
of Albrecht Durer. Princeton.
76. Hart, George W. "Dürer's Polyhedra" .
Retrieved 13 August 2009.
77. Dürer, Albrecht (1528). Hierinn sind
begriffen vier Bucher von
menschlicher Proportion .
Nurenberg. Retrieved 24 June 2015.
78. Rudy Rucker, The Fourth Dimension:
Toward a Geometry of Higher
Reality , Courier Corporation, 2014,
ISBN 0486798194
79. "Crucifixion (Corpus Hypercubus)" .
Metropolitan Museum of Art.
Retrieved 5 September 2015.
80. Schreiber, P. (1999). "A New
Hypothesis on Durer's Enigmatic
Polyhedron in His Copper Engraving
'Melencolia I' ". Historia
Mathematica. 26 (4): 369–377.
doi:10.1006/hmat.1999.2245 .
81. Dodgson, Campbell (1926). Albrecht
Dürer. London: Medici Society. p. 94.
82. Schuster, Peter-Klaus (1991).
Melencolia I: Dürers Denkbild. Berlin:
Gebr. Mann Verlag. pp. 17–83.
83. Panofsky, Erwin; Klibansky,
Raymond; Saxl, Fritz (1964). Saturn
and melancholy . Basic Books.
84. Lukman, Muhamad; Hariadi, Yun;
Destiarmand, Achmad Haldani
(2007). "Batik Fractal : Traditional Art
to Modern Complexity" (PDF).
Proceeding Generative Art X, Milan,
Italy. Retrieved 26 September 2016.
85. Ouellette, Jennifer (November 2001).
"Pollock's Fractals" . Discover
Magazine. Retrieved 26 September
2016.
86. Galilei, Galileo (1623). The Assayer.,
as translated in Drake, Stillman
(1957). Discoveries and Opinions of
Galileo. Doubleday. pp. 237–238.
ISBN 978-0-385-09239-5.
87. Cucker, Felix (2013). Manifold
Mirrors: The Crossing Paths of the
Arts and Mathematics. Cambridge
University Press. p. 381. ISBN 978-0-
521-72876-8.
88. Cucker, Felix (2013). Manifold
Mirrors: The Crossing Paths of the
Arts and Mathematics. Cambridge
University Press. p. 10. ISBN 978-0-
521-72876-8.
89. King, Jerry P. (1992). The Art of
Mathematics. Fawcett Columbine.
pp. 8–9. ISBN 978-0-449-90835-8.
90. King, Jerry P. (1992). The Art of
Mathematics. Fawcett Columbine.
pp. 135–139. ISBN 978-0-449-
90835-8.
91. Devlin, Keith (2000). "Do
Mathematicians Have Different
Brains?" . The Math Gene: How
Mathematical Thinking Evolved And
Why Numbers Are Like Gossip. Basic
Books. p. 140. ISBN 978-0-465-
01619-8.
92. Wasilewska, Katarzyna (2012).
"Mathematics in the World of
Dance" (PDF). Bridges. Retrieved
1 September 2015.
93. Malkevitch, Joseph. "Mathematics
and Art" . American Mathematical
Society. Retrieved 1 September
2015.
94. Malkevitch, Joseph. "Mathematics
and Art. 2. Mathematical tools for
artists" . American Mathematical
Society. Retrieved 1 September
2015.
95. "Math and Art: The Good, the Bad,
and the Pretty" . Mathematical
Association of America. Retrieved
2 September 2015.
96. Cohen, Louise (1 July 2014). "How to
spin the colour wheel, by Turner,
Malevich and more" . Tate Gallery.
Retrieved 4 September 2015.
97. Kemp, Martin (1992). The Science of
Art: Optical Themes in Western Art
from Brunelleschi to Seurat. Yale
University Press. ISBN 978-968-867-
185-6.
98. Gage, John (1999). Color and
Culture: Practice and Meaning from
Antiquity to Abstraction . University
of California Press. p. 207. ISBN 978-
0-520-22225-0.
99. Malkevitch, Joseph. "Mathematics
and Art. 3. Symmetry" . American
Mathematical Society. Retrieved
1 September 2015.
100. Malkevitch, Joseph. "Mathematics
and Art. 4. Mathematical artists and
artist mathematicians" . American
Mathematical Society. Retrieved
1 September 2015.
101. Wright, Richard (1988). "Some
Issues in the Development of
Computer Art as a Mathematical Art
Form". Leonardo. 1 (Electronic Art,
supplemental issue): 103–110.
doi:10.2307/1557919 .
JSTOR 1557919 .
102. Kalajdzievski, Sasho (2008). Math
and Art: An Introduction to Visual
Mathematics. Chapman and Hall.
ISBN 978-1-58488-913-7.
103. Beddard, Honor (2011-05-26).
"Computer art at the V&A" . Victoria
and Albert Museum. Retrieved
22 September 2015.
104. "Computer Does Drawings:
Thousands of lines in each". The
Guardian. 17 September 1962. in
Beddard, 2015.
105. O'Hanrahan, Elaine (2005). Drawing
Machines: The machine produced
drawings of Dr. D. P. Henry in relation
to conceptual and technological
developments in machine-generated
art (UK 1960–1968). Unpublished
MPhil. Thesis. John Moores
University, Liverpool. in Beddard,
2015.
106. Bellos, Alex (24 February 2015).
"Catch of the day: mathematician
nets weird, complex fish" . The
Guardian. Retrieved 25 September
2015.
107. " "A Bird in Flight (2016)," by Hamid
Naderi Yeganeh" . American
Mathematical Society. March 23,
2016. Retrieved April 6, 2017.
108. Chung, Stephy (September 18,
2015). "Next da Vinci? Math genius
using formulas to create fantastical
works of art" . CNN.
109. Levin, Golan (2013). "Generative
Artists" . CMUEMS. Retrieved
27 October 2015. This includes a link
to Hvidtfeldts Syntopia .
110. Verostko, Roman. "The Algorists" .
Retrieved 27 October 2015.
111. Cucker, Felix (2013). Manifold
Mirrors: The Crossing Paths of the
Arts and Mathematics. Cambridge
University Press. pp. 315–317.
ISBN 978-0-521-72876-8.
112. Miller, Arthur I. (2001). Einstein,
Picasso: Space, Time, and the
Beauty That Causes Havoc. New
York: Basic Books. p. 171. ISBN 978-
0-465-01860-4.
113. Miller, Arthur I. (2012). Insights of
Genius: Imagery and Creativity in
Science and Art. Springer. ISBN 978-
1-4612-2388-7.
114. Henderson, Linda D. (1983). The
Fourth Dimension and Non-
Euclidean geometry in Modern Art.
Princeton University Press.
115. Antliff, Mark; Leighten, Patricia Dee
(2001). Cubism and Culture (PDF).
Thames & Hudson.
116. Everdell, William R. (1997). The First
Moderns: Profiles in the Origins of
Twentieth-Century Thought.
University of Chicago Press. p. 312.
ISBN 978-0-226-22480-0.
117. Green, Christopher (1987). Cubism
and its Enemies, Modern Movements
and Reaction in French Art, 1916–
1928. Yale University Press. pp. 13–
47.
118. Metzinger, Jean (October–November
1910). "Note sur la peinture". Pan:
60. in Miller. Einstein, Picasso. Basic
Books. p. 167.
119. Metzinger, Jean (1972). Le cubisme
était né. Éditions Présence. pp. 43–
44. in Ferry, Luc (1993). Homo
Aestheticus: The Invention of Taste
in the Democratic Age. Robert De
Loaiza, trans. University of Chicago
Press. p. 215. ISBN 978-0-226-
24459-4.
120. "Man Ray–Human Equations A
Journey from Mathematics to
Shakespeare. February 7 – May 10,
2015" . Phillips Collection. Retrieved
5 September 2015.
121. Adcock, Craig (1987). "Duchamp's
Eroticism: A Mathematical
Analysis" . Iowa Research Online. 16
(1): 149–167.
122. Elder, R. Bruce (2013). DADA,
Surrealism, and the Cinematic
Effect . Wilfrid Laurier University
Press. p. 602. ISBN 978-1-55458-
641-7.
123. Tubbs, Robert (2014). Mathematics
in Twentieth-Century Literature and
Art: Content, Form, Meaning . JHU
Press. p. 118. ISBN 978-1-4214-
1402-7.
124. "Hiroshi Sugimoto Conceptual Forms
and Mathematical Models February
7 – May 10, 2015" . Phillips
Collection. Retrieved 5 September
2015.
125. Tubbs, Robert (2014). Mathematics
in 20th-Century Literature and Art.
Johns Hopkins. pp. 8–10. ISBN 978-
1-4214-1380-8.
126. Keats, Jonathon (13 February 2015).
"See How Man Ray Made Elliptic
Paraboloids Erotic At This Phillips
Collection Photography Exhibit" .
Forbes. Retrieved 10 September
2015.
127. Gamwell, Lynn (2015). Mathematics
and Art: A Cultural History. Princeton
University Press. pp. 311–312.
ISBN 978-0-691-16528-8.
128. Hedgecoe, John, ed. (1968). Henry
Moore: Text on His Sculpture. Henry
Spencer Moore. Simon and Schuster.
p. 105.
129. "De Stijl" . Tate Glossary. The Tate.
Retrieved 11 September 2015.
130. Curl, James Stevens (2006). A
Dictionary of Architecture and
Landscape Architecture (Second
ed.). Oxford University Press.
ISBN 978-0-19-860678-9.
131. Tubbs, Robert (2014). Mathematics
in Twentieth-Century Literature and
Art: Content, Form, Meaning. JHU
Press. pp. 44–47. ISBN 978-1-4214-
1402-7.
132. "Tour: M.C. Escher – Life and Work" .
NGA. Archived from the original on
3 August 2009. Retrieved 13 August
2009.
133. "MC Escher" . Mathacademy.com. 1
November 2007. Retrieved
13 August 2009.
134. Penrose, L.S.; Penrose, R. (1958).
"Impossible objects: A special type
of visual illusion". British Journal of
Psychology. 49 (1): 31–33.
doi:10.1111/j.2044-
8295.1958.tb00634.x .
PMID 13536303 .
135. Kirousis, Lefteris M.; Papadimitriou,
Christos H. (1985). The complexity
of recognizing polyhedral scenes.
26th Annual Symposium on
Foundations of Computer Science
(FOCS 1985). pp. 175–185.
CiteSeerX 10.1.1.100.4844 .
doi:10.1109/sfcs.1985.59 .
ISBN 978-0-8186-0644-1.
136. Cooper, Martin (2008). "Tractability
of Drawing Interpretation". Line
Drawing Interpretation. Springer-
Verlag. pp. 217–230.
doi:10.1007/978-1-84800-229-6_9 .
ISBN 978-1-84800-229-6.
137. Roberts, Siobhan (2006). 'Coxetering'
with M.C. Escher. King of Infinite
Space: Donald Coxeter, the Man Who
Saved Geometry. Walker. p. Chapter
11.
138. Escher, M.C. (1988). The World of
MC Escher. Random House.
139. Escher, M.C.; Vermeulen, M.W.; Ford,
K. (1989). Escher on Escher:
Exploring the Infinite. HN Abrams.
140. Malkevitch, Joseph. "Mathematics
and Art. 5. Polyhedra, tilings, and
dissections" . American
Mathematical Society. Retrieved
1 September 2015.
141. Marcolli, Matilde (July 2016). The
notion of Space in Mathematics
through the lens of Modern Art
(PDF). Century Books. pp. 23–26.

142. "John Robinson" . Bradshaw


Foundation. 2007. Retrieved
13 August 2009.
143. "Helaman Ferguson web site" .
Helasculpt.com. Archived from the
original on 11 April 2009. Retrieved
13 August 2009.
144. Thurston, William P. (1999). Levy,
Silvio (ed.). The Eightfold Way: A
Mathematical Sculpture by Helaman
Ferguson (PDF). Volume 35: The
Eightfold Way: The Beauty of Klein's
Quartic Curve. MSRI Publications.
pp. 1–7.
145. "MAA book review of The Eightfold
Way: The Beauty of Klein's Quartic
Curve" . Maa.org. 14 November
1993. Retrieved 13 August 2009.
146. "The Math Geek Holiday Gift Guide" .
Scientific American. 23 November
2014. Retrieved 7 June 2015.
147. Hanna, Raven. "Gallery: Bathsheba
Grossman" . Symmetry Magazine.
Retrieved 7 June 2015.
148. Fleron, Julian F.; Ecke, Volker; von
Renesse, Christine; Hotchkiss, Philip
K. (January 2015). Art and Sculpture:
Mathematical Inquiry in the Liberal
Arts (2nd ed.). Discovering the Art
of Mathematics project.
149. Osinga, Hinke (2005). "Crocheting
the Lorenz manifold" . University of
Auckland. Archived from the
original on 10 April 2015. Retrieved
12 October 2015.
150. Henderson, David; Taimina, Daina
(2001). "Crocheting the hyperbolic
plane" (PDF). Mathematical
Intelligencer. 23 (2): 17–28.
doi:10.1007/BF03026623 ..
151. Osinga, Hinke M; Krauskopf, Bernd
(2004). "Crocheting the Lorenz
manifold" . The Mathematical
Intelligencer. 26 (4): 25–37.
CiteSeerX 10.1.1.108.4594 .
doi:10.1007/BF02985416 .
152. Dietz, Ada K. (1949), Algebraic
Expressions in Handwoven Textiles
(PDF), Louisville, Kentucky: The Little
Loomhouse
153. Miller, J. C. P. (1970). "Periodic
forests of stunted trees".
Philosophical Transactions of the
Royal Society of London. Series A,
Mathematical and Physical
Sciences. 266 (1172): 63–111.
Bibcode:1970RSPTA.266...63M .
doi:10.1098/rsta.1970.0003 .
JSTOR 73779 .
154. "Pat Ashforth & Steve Plummer –
Mathekniticians" . Woolly Thoughts.
Retrieved 4 October 2015.
155. Ward, Mark (20 August 2012).
"Knitting reinvented: Mathematics,
feminism and metal" . BBC News.
BBC. Retrieved 23 September 2015.
156. Ashforth, Pat; Plummer, Steve.
"Menger Sponge" . Woolly Thoughts:
In Pursuit of Crafty Mathematics.
Retrieved 23 September 2015.
157. Ashforth, Pat; Plummer, Steve.
"Afghans for Schools" . Woolly
Thoughts: Mathghans. Retrieved
23 September 2015.
158. "Mathghans with a Difference" .
Simply Knitting Magazine. 1 July
2008. Retrieved 23 September 2015.
159. Jouffret, Esprit (1903). Traité
élémentaire de géométrie à quatre
dimensions et introduction à la
géométrie à n dimensions (in
French). Paris: Gauthier-Villars.
OCLC 1445172 . Retrieved
26 September 2015.
160. Seckel, Hélène (1994). "Anthology of
Early Commentary on Les
Demoiselles d'Avignon". In William
Rubin; Hélène Seckel; Judith Cousins
(eds.). Les Demoiselles d'Avignon.
New York: Museum of Modern Art.
p. 264. ISBN 978-0-87070-162-7.
161. "Giotto di Bondone and assistants:
Stefaneschi triptych" . The Vatican.
Retrieved 16 September 2015.
162. Gamwell, Lynn (2015). Mathematics
and Art: A Cultural History. Princeton
University Press. pp. 337–338.
ISBN 978-0-691-16528-8.
163. Cooper, Jonathan (5 September
2007). "Art and Mathematics" .
Retrieved 5 September 2015.
164. Hofstadter, Douglas R. (1980). Gödel,
Escher, Bach: An Eternal Golden
Braid. Penguin. p. 627. ISBN 978-0-
14-028920-6.
165. Hall, James (10 June 2011). "René
Magritte: The Pleasure Principle –
exhibition" . The Guardian. Retrieved
5 September 2015.
166. King, Elliott (2004). Ades, Dawn (ed.).
Dali. Milan: Bompiani Arte. pp. 418–
421.
167. "Stone balancing" (PDF). Monthly
Maths (29). July 2013. Retrieved
10 June 2017.
168. Hofstadter, Douglas R. (1980). Gödel,
Escher, Bach: An Eternal Golden
Braid. Penguin. pp. 98–99, 690–717.
ISBN 978-0-394-74502-2.
169. de Smit, B. (2003). "The
Mathematical Structure of Escher's
Print Gallery". Notices of the
American Mathematical Society. 50
(4): 446–451.
170. Lenstra, Hendrik; De Smit, Bart.
"Applying mathematics to Escher's
Print Gallery" . Leiden University.
Retrieved 10 November 2015.
171. Stanek, Becca (16 June 2014). "Van
Gogh and the Algorithm: How Math
Can Save Art" . Time Magazine.
Retrieved 4 September 2015.
172. Sipics, Michelle (18 May 2009). "The
Van Gogh Project: Art Meets
Mathematics in Ongoing
International Study" . Society for
Industrial and Applied Mathematics.
Archived from the original on 7
September 2015. Retrieved
4 September 2015.
173. Emmerling, Leonhard (2003).
Jackson Pollock, 1912–1956 . p. 63.
ISBN 978-3-8228-2132-9.
174. Taylor, Richard P.; Micolich, Adam P.;
Jonas, David (June 1999). "Fractal
analysis of Pollock's drip paintings"
(PDF). Nature. 399 (6735): 422.
Bibcode:1999Natur.399..422T .
doi:10.1038/20833 . Archived from
the original (PDF) on 2015-08-16.
175. Taylor, Richard; Micolich, Adam P.;
Jonas, David (October 1999).
"Fractal Expressionism: Can Science
Be Used To Further Our
Understanding Of Art?" . Physics
World. 12 (10): 25–28.
doi:10.1088/2058-7058/12/10/21 .
Archived from the original on 2012-
08-05. "Pollock died in 1956, before
chaos and fractals were discovered.
It is highly unlikely, therefore, that
Pollock consciously understood the
fractals he was painting.
Nevertheless, his introduction of
fractals was deliberate. For example,
the colour of the anchor layer was
chosen to produce the sharpest
contrast against the canvas
background and this layer also
occupies more canvas space than
the other layers, suggesting that
Pollock wanted this highly fractal
anchor layer to visually dominate the
painting. Furthermore, after the
paintings were completed, he would
dock the canvas to remove regions
near the canvas edge where the
pattern density was less uniform."
176. King, M. (2002). "From Max Ernst to
Ernst Mach: epistemology in art and
science" (PDF). Retrieved
17 September 2015.
177. Dodgson, N. A. (2012).
"Mathematical characterisation of
Bridget Riley's stripe paintings"
(PDF). Journal of Mathematics and
the Arts. 5 (2–3): 89–106.
doi:10.1080/17513472.2012.67946
8 . "over the course [of] the early
1980s, Riley's patterns moved from
more regular to more random (as
characterised by global entropy),
without losing their rhythmic
structure (as characterised by local
entropy). This reflects Kudielka's
description of her artistic
development."
178. Cucker, Felix (2013). Manifold
Mirrors: The Crossing Paths of the
Arts and Mathematics. Cambridge
University Press. pp. 116–120.
ISBN 978-0-521-72876-8.
179. Treibergs, Andrejs (24 July 2001).
"The Geometry of Perspective
Drawing on the Computer" .
University of Utah. Retrieved
5 September 2015.
180. Gamwell, Lynn (2015). Mathematics
and Art: A Cultural History. Princeton
University Press. p. xviii. ISBN 978-0-
691-16528-8.
181. Malkevitch, Joseph. "Mathematics
and Art. 6. Origami" . American
Mathematical Society. Retrieved
1 September 2015.
182. T. Sundara Rao (1893). Geometric
Exercises in Paper Folding .
Addison.
183. Justin, J. (June 1986). "Mathematics
of Origami, part 9". British Origami:
28–30..
184. Alsina, Claudi; Nelsen, Roger (2010).
Charming Proofs: A Journey Into
Elegant Mathematics . Dolciani
Mathematical Expositions. 42.
Mathematical Association of
America. p. 57. ISBN 978-0-88385-
348-1.
185. Alperin, Roger C.; Lang, Robert J.
(2009). "One-, Two-, and Multi-Fold
Origami Axioms" (PDF). 4OSME.
186. The World of Geometric Toys ,
Origami Spring , August, 2007.
187. Cucker, Felix (2013). Manifold
Mirrors: The Crossing Paths of the
Arts and Mathematics. Cambridge
University Press. pp. 163–166.
ISBN 978-0-521-72876-8.
188. Gamwell, Lynn (2015). Mathematics
and Art: A Cultural History. Princeton
University Press. pp. 406–410.
ISBN 978-0-691-16528-8.
189. Ghyka, Matila (2003). The Geometry
of Art and Life. Dover. pp. ix–xi.
ISBN 978-0-486-23542-4.
190. Lawlor, Robert (1982). Sacred
Geometry: Philosophy and Practice.
Thames & Hudson. ISBN 978-0-500-
81030-9.
191. Calter, Paul (1998). "Celestial
Themes in Art & Architecture" .
Dartmouth College. Retrieved
5 September 2015.
192. "The Thought of a Thought – Edgar
Allan Poe" . MathPages. Retrieved
5 September 2015.
193. Livio, Mario (November 2002). "The
golden ratio and aesthetics" .
Retrieved 26 June 2015.

External links

Wikimedia Commons has media


related to Mathematics in art.
Bridges Organization conference on
connections between art and
mathematics
Bridging the Gap Between Math and
Art – Slide Show from Scientific
American
Connections in Space – topology in
art
Discovering the Art of Mathematics
Mathematics and Art – AMS
Mathematics and Art – Cut-the-Knot
Mathematical Imagery – American
Mathematical Society
Mathematics in Art and Architecture –
National University of Singapore
Mathematical Art – Virtual Math
Museum
When art and math collide – Science
News
Why the history of maths is also the
history of art : Lynn Gamwell in The
Guardian

Retrieved from
"https://en.wikipedia.org/w/index.php?
title=Mathematics_and_art&oldid=897396385"

Last edited 23 days ago by Reify-t…

Content is available under CC BY-SA 3.0 unless


otherwise noted.

Вам также может понравиться