Download as pdf or txt
Download as pdf or txt
You are on page 1of 34

Mathematics and art

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 conjectured to have been 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 Mathematics in art: Albrecht Dürer's
Prospectiva Pingendi, and in his paintings. The engraver copper plate engraving Melencolia I,
Albrecht Dürer made many references to mathematics in his 1514. Mathematical references include a
work Melencolia I. In modern times, the graphic artist M. C. compass for geometry, a magic square
Escher made intensive use of tessellation and hyperbolic and a truncated rhombohedron, while
geometry, with the help of the mathematician H. S. M. measurement is indicated by the scales
Coxeter, while the De Stijl movement led by Theo van and hourglass.[1]
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 René 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.

Origins: from ancient Greece to the


Renaissance

Polykleitos's Canon and symmetria

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
Wireframe drawing[2] of a vase as a solid
and the statue of Hera in the Heraion of Argos.[3] While his
of revolution[2] by Paolo Uccello. 15th
sculptures may not be as famous as those of Phidias, they are
century
much admired. In his Canon, a treatise he wrote designed to
document the "perfect" body proportions of the male nude,
Polykleitos gives us a mathematical approach towards sculpturing the human
body.[3]

The Canon itself has been lost but it is conjectured that Polykleitos used a
sequence of proportions where each length is that of the diagonal of a square
drawn on its predecessor, 1:√ 2 (about 1:1.4142).[4]

The influence of the Canon of Polykleitos is immense in Classical Greek, Roman,


and Renaissance sculpture, with 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.[5]
The Canon applies the basic mathematical concepts of Greek geometry, such as
the ratio, proportion, and symmetria (Greek for "harmonious proportions") and Roman copy in
turns it into a system capable of describing the human form through a series of marble of
continuous geometric progressions.[4] Doryphoros,
originally a bronze by
Polykleitos
Perspective and proportion

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.[6] 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.[7]

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 Brunelleschi's experiment with linear
Florence, using similar triangles as formulated by Euclid, to find the perspective
apparent height of distant objects.[8][9] Brunelleschi's own
perspective paintings are lost, but Masaccio's painting of the Holy
Trinity shows his principles at work.[6][10][11]

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.[12][13]

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
Paolo Uccello made innovative use of
solid geometry and perspective, including De prospectiva
perspective in The Battle of San Romano
pingendi (On Perspective for Painting), Trattato d'Abaco
(c. 1435–1460).
(Abacus Treatise), and De quinque corporibus regularibus
(On the Five Regular Solids).[14][15][16] The historian Vasari
in his Lives of the Painters calls Piero the "greatest geometer
of his time, or perhaps of any time." [17] Piero's interest in perspective can be seen in his paintings including
the Polyptych of Perugia, [18] 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.[19] He was
taught commercial arithmetic in "abacus schools"; his writings are formatted like abacus school
textbooks,[20] 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.[21]

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][7] He uses deductive logic to lead
the reader to the perspective representation of a three-dimensional body.[22]

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.[23] Critics disagree on whether Hockney was correct.[24][25] Similarly, the architect Philip
Steadman argued controversially[26] that Vermeer had used a different device, the camera obscura, to help
him create his distinctively observed paintings.[27]
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.[28] 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.[29] Leonardo studied
Pacioli's Summa, from which he copied tables of proportions.[30] In Mona Lisa and The Last Supper,
Leonardo's work incorporated linear perspective with a vanishing point to provide apparent depth.[31] 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.[32][33] 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.[34]

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,[35] 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.[36]

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).[37]
Illustration of an artist
using a camera obscura.
17th century
Woodcut from Luca Camera lucida in use.
Pacioli's 1509 De divina Scientific American,
proportione with an 1879
equilateral triangle on a
human face

Diagram from Leon


Proportion: Leonardo's Brunelleschi's theory of Battista Alberti's 1435
Vitruvian Man, c. 1490 perspective: Masaccio's Della Pittura, with pillars
Trinità, c. 1426–1428, in in perspective on a grid
the Basilica of Santa
Maria Novella
Linear perspective in
Piero della Francesca's Curvilinear perspective:
Flagellation of Christ, c. convex mirror in Jan van
1455–1460 Parmigianino, Self-
Eyck's Arnolfini Portrait,
portrait in a Convex
1434
Mirror, c. 1523–1524

Oblique projection: Oblique projection:


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

Golden ratio

The golden ratio (roughly equal to 1.618) was known to Euclid.[38] The golden ratio has persistently been
claimed[39][40][41][42] in modern times to have been used in art and architecture by the ancients in Egypt,
Greece and elsewhere, without reliable evidence.[43] The claim may derive from confusion with "golden
mean", which to the Ancient Greeks meant "avoidance of excess in either direction", not a ratio.[43]
Pyramidologists since the 19th 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,[46][47][48] but these claims too are disproved by measurement.[43]
The Great Mosque of Kairouan in Tunisia has similarly been claimed to use the golden ratio in its
design,[49] but the ratio does not appear in the original parts of the mosque.[50] 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,[51] 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.[52] 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.[53] After Pacioli, the golden ratio is more definitely discernible in artworks including
Leonardo's Mona Lisa.[54]
Another ratio, the only other morphic number,[55] was named the plastic number[c] in 1928 by the Dutch
architect Hans van der Laan (originally named le nombre radiant in French).[56] 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.[56]

Base:hypotenuse(b:a)
ratios for the Pyramid of
Khufu could be: 1:φ
(Kepler triangle), 3:5 (3- Supposed ratios: Notre- Golden rectangles
4-5 Triangle), or 1:4/π Dame of Laon superimposed on the
Mona Lisa

The 1967 St.


Benedictusberg Abbey
church by Hans van der
Laan has plastic number
proportions.

Planar symmetries

Planar symmetries have for millennia been exploited in artworks such as carpets, lattices, textiles and
tilings.[58][59][60][61]

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.[58] 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.[58] The mathematician and architectural theorist Nikos
Salingaros suggests that the "powerful presence"[57] (aesthetic effect) of a
"great carpet"[57] 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
Powerful presence:[57] at least nine of the above ten rules", and suggests that it might be possible to
carpet with double create a metric from these rules.[57]
medallion. Central Anatolia
(Konya – Karapınar), turn of Elaborate lattices are found in Indian Jali work, carved in marble to adorn
the 16th/17th centuries. tombs and palaces.[59] Chinese lattices, always with some symmetry, exist
Alâeddin Mosque 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.[62] Many Chinese lattices have been analysed
mathematically by Daniel S. Dye; he identifies Sichuan as the centre of the craft.[63]

Symmetries are prominent in textile arts including quilting,[60] knitting,[64]


cross-stitch, crochet,[65] embroidery[66][67] and weaving,[68] where they
may be purely decorative or may be marks of status.[69] 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.[70] 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.[71] Girih tiles

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.[72] Elaborate geometric zellige tilework is a distinctive element in
Moroccan architecture.[61] Muqarnas vaults are three-dimensional but were designed in two dimensions
with drawings of geometrical cells.[73]
Jaali marble lattice at
Hotamis kilim (detail), Detail of a Ming Dynasty
tomb of Salim Chishti,
central Anatolia, early brocade, using a
Fatehpur Sikri, India
19th century chamfered hexagonal
lattice pattern

Ceiling of the Sheikh Rotational symmetry in


Symmetries: Florentine Lotfollah Mosque, lace: tatting work
Bargello pattern tapestry Isfahan, 1619
work
Girih tiles: patterns at The complex geometry
large and small scales and tilings of the
on a spandrel from the muqarnas vaulting in the
Darb-i Imam shrine, Tessellations: zellige Sheikh Lotfollah
Isfahan, 1453 mosaic tiles at Bou Mosque, Isfahan
Inania Madrasa, Fes,
Morocco

Architect's plan of a Tupa Inca tunic from


muqarnas quarter vault. Peru, 1450 –1540, an
Topkapı Scroll Andean textile denoting
high rank[69]

Polyhedra

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;[12] in Leonardo da Vinci's diagrams of regular polyhedra drawn as
illustrations for Luca Pacioli's 1509 book The Divine Proportion;[12] as a glass rhombicuboctahedron in
Jacopo de Barbari's portrait of Pacioli, painted in 1495;[12] in the truncated polyhedron (and various other
mathematical objects) in Albrecht Dürer's engraving Melencolia I;[12] and in Salvador Dalí's painting The
Last Supper in which Christ and his disciples are pictured inside a giant dodecahedron.[74]

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]

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][78][79] including a two-
volume book by Peter-Klaus Schuster,[80] and an influential discussion in Erwin Panofsky's monograph of
Dürer.[1][81]

Salvador Dalí's 1954 painting Corpus Hypercubus uniquely depicts the cross of Christ as an unfolded
three-dimensional net for a hypercube, also known as a tesseract: the unfolding of a tesseract into these
eight cubes is analogous to unfolding the sides of a cube into a cross shape of six squares, here representing
the divine perspective with a four-dimensional regular polyhedron.[82][83] The painting shows the figure of
Christ in front of the tessaract; he would normally be shown fixed with nails to the cross, but there are no
nails in the painting. Instead, there are four small cubes in front of his body, at the corners of the frontmost
of the eight tessaract cubes. The mathematician Thomas Banchoff states that Dalí was trying to go beyond
the three-dimensional world, while the poet and art critic Kelly Grovier says that "The painting seems to
have cracked the link between the spirituality of Christ's salvation and the materiality of geometric and
physical forces. It appears to bridge the divide that many feel separates science from religion."[84]

The first printed Icosahedron as a part of


illustration of a the monument to Baruch
rhombicuboctahedron, Spinoza, Amsterdam
by Leonardo da Vinci,
published in De Divina
Proportione, 1509

Fractal dimensions

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.[85]
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.[86]

A complex relationship
Batiks from Surakarta, Java, like
this parang klithik sword pattern,
The astronomer Galileo Galilei in his Il Saggiatore wrote that "[The
have a fractal dimension between
universe] is written in the language of mathematics, and its
1.2 and 1.5.
characters are triangles, circles, and other geometric figures."[87]
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.[88]
Some of the many strands of the resulting complex relationship[89] are described below.

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.[90] 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 The mathematician G. H.
economy" (King's italics), and describes the proof as "aesthetically Hardy defined a set of
pleasing".[91] The Hungarian mathematician Paul Erdős agreed that criteria for mathematical
mathematics possessed beauty but considered the reasons beyond beauty.
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."[92]

Mathematical tools for art

Mathematics can be discerned in many of the arts, such as music, dance,[93] painting, architecture, and
sculpture. Each of these is richly associated with mathematics.[94] 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.[95] 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.[94][96] 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,[6] 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,[97] the Pre-Raphaelites and
Wassily Kandinsky.[98][99] Artists may also choose to analyse the symmetry of a scene.[100] 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.[101]

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 Octopod by Mikael Hvidtfeldt
representational systems." He gives as instances an image from the Christensen. Algorithmic art
Mandelbrot set, an image generated by a cellular automaton produced with the software Structure
algorithm, and a computer-rendered image, and discusses, with Synth
reference to the Turing test, whether algorithmic products can be
art.[102] 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.[103]

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.[104][105] The machine was
capable of creating complex, abstract, asymmetrical, curvilinear, but repetitive line drawings.[104][106] 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.[107][108][109] 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.[110][111]
Fractal sculpture: 3D
Fraktal 03/H/dd by
Hartmut Skerbisch, 2003
Mathematical sculpture Fibonacci word: detail of
by Bathsheba artwork by Samuel
Grossman, 2007 Monnier, 2009

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

From mathematics to art

The mathematician and theoretical physicist Henri Poincaré's Science and


Hypothesis was widely read by the Cubists, including Pablo Picasso and
Jean Metzinger.[113][114] Being thoroughly familiar with Bernhard
Riemann's work on non-Euclidean geometry, Poincaré was more than
aware that Euclidean geometry is 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.[115][116][117] The concept that painting could be expressed
mathematically, in colour and form, contributed to Cubism, the art Proto-Cubism: Pablo
movement that led to abstract art.[118] Metzinger, in 1910, wrote that: " Picasso's 1907 painting Les
[Picasso] lays out a free, mobile perspective, from which that ingenious Demoiselles d'Avignon uses
a fourth dimension
mathematician Maurice Princet has deduced a whole geometry".[119] Later,
projection to show a figure
Metzinger wrote in his memoirs:
both full face and in
profile.[112]
Maurice Princet joined us often ... it was as an artist that he conceptua
aesthetician that he invoked n-dimensional continuums. He loved to g
the new views on space that had been opened up by Schlegel and som
at that.[120]

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,[121] Marcel Duchamp[122] and Max Ernst,[123][124] and following Man Ray, Hiroshi Sugimoto.[125]

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.[126] The art
reporter Jonathan Keats, writing in ForbesLife, argues that Man Ray
Enneper surfaces as photographed "the elliptic paraboloids and conic points in the same sensual
Dadaism: Man Ray's 1934 light as his pictures of Kiki de Montparnasse", and "ingeniously repurposes
Objet mathematique the cool calculations of mathematics to reveal the topology of desire".[127]
Twentieth century sculptors such as Henry Moore, Barbara Hepworth and
Naum Gabo took inspiration from mathematical models.[128] 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."[129]

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".[130][131] 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.[130] After the breakup of De Stijl, Van Doesburg founded the Avant-garde Art Concret
movement, describing his 1929–1930 Arithmetic Composition (https://commons.wikimedia.org/wiki/File:T
heo_van_Doesburg_218.jpg), 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.[132]

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.[133][134] 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.[135][136][137]

Some of Escher's many tessellation drawings were inspired by conversations with the
mathematician H. S. M. Coxeter on hyperbolic geometry.[138] 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.[139] 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.[140]

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.[141]

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.[142]

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.[7] 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 Theo van
without breaking.[143] The sculptor Helaman Ferguson creates complex surfaces and Doesburg's Six
Moments in
other topological objects.[144] His works are visual representations of mathematical
the
objects; The Eightfold Way is based on the projective special linear group PSL(2,7), a
Development
finite group of 168 elements.[145][146] The sculptor Bathsheba Grossman similarly
of Plane to
bases her work on mathematical structures.[147][148] The artist Nelson Saiers Space, 1926 or
incorporates mathematical concepts and theorems in his art from toposes and schemes to
1929
the four color theorem and the irrationality of π.[149]

A liberal arts inquiry project examines connections between mathematics and art through the Möbius strip,
flexagons, origami and panorama photography.[150]

Mathematical objects including the Lorenz manifold and the hyperbolic plane have been crafted using fiber
arts including crochet.[d][152] The American weaver Ada Dietz wrote a 1949 monograph Algebraic
Expressions in Handwoven Textiles, defining weaving patterns based on the expansion of multivariate
polynomials.[153] The mathematician Daina Taimiņa demonstrated features of the hyperbolic plane by
crocheting in 2001.[154] This led Margaret and Christine Wertheim to crochet a coral reef, consisting of
many marine animals such as nudibranchs whose shapes are based on hyperbolic planes.[155] The
mathematician J. C. P. Miller used the Rule 90 cellular automaton to design tapestries depicting both trees
and abstract patterns of triangles.[156] The "mathekniticians"[157] 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.[158][159] Their
"mathghans" (Afghans for Schools) project introduced knitting into the British mathematics and technology
curriculum.[160][161]

Pedagogy to art: Magnus


Wenninger with some of
his stellated polyhedra,
Four-dimensional space De Stijl: Theo van 2009
to Cubism: Esprit Doesburg's geometric
Jouffret's 1903 Traité Composition I (Still Life),
élémentaire de 1916
géométrie à quatre
dimensions.[162][e]

Crocheted coral reef:


many animals modelled
as hyperbolic planes
with varying parameters
A Möbius strip scarf in Anamorphism: The by Margaret and
crochet, 2007 Ambassadors by Hans Christine Wertheim. Föhr
Holbein the Younger, Reef, Tübingen, 2013
1533, with severely
distorted skull in
foreground

Illustrating mathematics

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.[164] Giorgio
de 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.[165]

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.[166] 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
Front face of Giotto's Stefaneschi children's building blocks, but
Triptych, 1320 illustrates recursion.
house-sized.[167] The Guardian Semiotic joke: René
observed that the "eerie toytown
Magritte's La condition
image" prophesied Modernism's humaine 1933
usurpation of "cosy traditional forms", but also
plays with the human tendency to seek patterns in
nature.[168]

Salvador Dalí's last painting, The


Swallow's Tail (1983), was part of a
series inspired by René Thom's
catastrophe theory.[170] 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
Detail of Cardinal Diagram of the apparent paradox embodied in
all nature. Consisting of simple
Stefaneschi M. C. Escher's 1956 lithograph Print Gallery,
geometric shapes with detailed
holding the as discussed by Douglas Hofstadter in his
patterning and coloring, in works
triptych 1980 book Gödel, Escher, Bach[169]
such as Angular I and Automnes,
Palazuelo expressed himself in
geometric transformations.[7]

The artist Adrian Gray practises stone balancing, exploiting friction and the centre of gravity to create
striking and seemingly impossible compositions.[171]

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
Lithograph Print Gallery by
of the seaside town, there being a "strange loop, or tangled hierarchy" to the M. C. Escher, 1956
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.[169]
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.[172][173]

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.[174][175]

Jackson Pollock's drip painting style[176] has a definite fractal


dimension;[177] among the artists who may have influenced Pollock's
controlled chaos,[178] Max Ernst painted Lissajous figures directly by
swinging a punctured bucket of paint over a canvas.[179]

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.[180]

The American mathematician George Birkhoff's 1933 Aesthetic Measure


proposes a quantitative metric of the aesthetic quality of an artwork. It does Max Ernst making
Lissajous figures, New
not attempt to measure the connotations of a work, such as what a painting
York, 1942
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.[181]

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,[182] and ultimately to
the mathematics of projective geometry of Girard Desargues and Jean-Victor Poncelet.[183]

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.[184] Paper-
folding was used in 1893 by T. Sundara Rao in his Geometric Exercises in Paper Folding to demonstrate
geometrical proofs.[185] The mathematics of paper folding has been explored in Maekawa's theorem,[186]
Kawasaki's theorem,[187] and the Huzita–Hatori axioms.[188]
Mathematical origami:
Spring Into Action, by
Jeff Beynon, made from
Stimulus to projective a single paper
geometry: Alberti's rectangle.[189]
diagram showing a circle
seen in perspective as
an ellipse. Della Pittura,
1435–1436

Illusion to Op art

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,[190] and Victor
Vasarely.[191]
The Fraser spiral illusion,
Sacred geometry named for Sir James Fraser
who discovered it in 1908.
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.[192][193] An illumination in the 13th-century Codex Vindobonensis
shows God drawing out the universe with a pair of compasses, which 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), .[194] 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.[194] William
Blake's Ancient of Days (depicting Urizen, Blake's embodiment of reason and law) and his painting of the
physicist Isaac Newton, naked, hunched and drawing with a compass, use the symbolism of compasses to
critique conventional reason and materialism as narrow-minded.[195][196] Salvador Dalí's 1954 Crucifixion
(Corpus Hypercubus) depicts the cross as a hypercube, representing the divine perspective with four
dimensions rather than the usual three.[83] In Dalí's The Sacrament of the Last Supper (1955) Christ and his
disciples are pictured inside a giant dodecahedron.[197]

Johannes Kepler's
God the geometer. The creation, with the
Platonic solid model of
Codex Vindobonensis, Pantocrator bearing.
planetary spacing in the
c. 1220 Bible of St Louis, c.
Solar System from
1220–1240
Mysterium
Cosmographicum, 1596

William Blake's Newton,


c. 1800
William Blake's The
Ancient of Days, 1794

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 the use of the Kepler triangle (face angle 51°49').[43][44] 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').[45]
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.[151]
e. Maurice Princet gave a copy to Pablo Picasso, whose sketchbooks for Les Demoiselles
d'Avignon illustrate Jouffret's influence.[113][163]

References
1. Ziegler, Günter M. (3 December 2014). "Dürer's polyhedron: 5 theories that explain
Melencolia's crazy cube" (https://www.theguardian.com/science/alexs-adventures-in-number
land/2014/dec/03/durers-polyhedron-5-theories-that-explain-melencolias-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 (http
s://citeseerx.ist.psu.edu/viewdoc/summary?doi=10.1.1.58.8477).
doi:10.1109/TPAMI.2005.14 (https://doi.org/10.1109%2FTPAMI.2005.14). PMID 15628272
(https://pubmed.ncbi.nlm.nih.gov/15628272). S2CID 13387519 (https://api.semanticscholar.o
rg/CorpusID:13387519).
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 (https://doi.org/10.2307%2F630196). JSTOR 630196 (https://www.jstor.
org/stable/630196). S2CID 162410725 (https://api.semanticscholar.org/CorpusID:16241072
5).
4. Tobin, Richard (October 1975). "The Canon of Polykleitos". American Journal of
Archaeology. 79 (4): 307–321. doi:10.2307/503064 (https://doi.org/10.2307%2F503064).
JSTOR 503064 (https://www.jstor.org/stable/503064). S2CID 191362470 (https://api.semanti
cscholar.org/CorpusID:191362470).
5. Raven, J. E. (1951). "Polyclitus and Pythagoreanism". Classical Quarterly. 1 (3–4): 147–.
doi:10.1017/s0009838800004122 (https://doi.org/10.1017%2Fs0009838800004122).
S2CID 170092094 (https://api.semanticscholar.org/CorpusID:170092094).
6. O'Connor, J. J.; Robertson, E. F. (January 2003). "Mathematics and art – perspective" (http://
www-history.mcs.st-and.ac.uk/HistTopics/Art.html). University of St Andrews. Retrieved
1 September 2015.
7. Emmer, Michelle, ed. (2005). The Visual Mind II (https://archive.org/details/visualmindiileon0
0mich). MIT Press. ISBN 978-0-262-05048-7.
8. Vasari, Giorgio (1550). Lives of the Artists. Torrentino. p. Chapter on Brunelleschi.
9. Alberti, Leon Battista; Spencer, John R. (1956) [1435]. On Painting (http://www.noteaccess.c
om/Texts/Alberti/). Yale University Press.
10. Field, J. V. (1997). The Invention of Infinity: Mathematics and Art in the Renaissance. Oxford
University Press. ISBN 978-0-19-852394-9.
11. Witcombe, Christopher L. C. E. "Art History Resources" (http://arthistoryresources.net/renais
sance-art-theory-2012/perspective.html). Retrieved 5 September 2015.
12. Hart, George W. "Polyhedra in Art" (http://www.georgehart.com/virtual-polyhedra/art.html).
Retrieved 24 June 2015.
13. Cunningham, Lawrence; Reich, John; Fichner-Rathus, Lois (1 January 2014). Culture and
Values: A Survey of the Western Humanities (https://books.google.com/books?id=0t0bCgAA
QBAJ&pg=PA375). 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."
14. della Francesca, Piero (1942) [c. 1474]. G. Nicco Fasola (ed.). De prospectiva pingendi.
Florence.
15. della Francesca, Piero (1970) [Fifteenth century]. G. Arrighi (ed.). Trattato d'Abaco. Pisa.
16. della Francesca, Piero (1916). G. Mancini (ed.). L'opera "De corporibus regularibus" di
Pietro Franceschi detto della Francesca usurpata da Fra Luca Pacioli.
17. Vasari, Giorgio (1878). G. Milanesi (ed.). Le Opere, volume 2. p. 490.
18. Zuffi, Stefano (1991). Piero della Francesca (https://archive.org/details/pierodellafrance00zuf
f). L'Unità – Mondadori Arte. p. 53 (https://archive.org/details/pierodellafrance00zuff/page/n5
2).
19. Heath, T. L. (1908). The Thirteen Books of Euclid's Elements (https://archive.org/details/bub_
gb_lxkPAAAAIAAJ). Cambridge University Press. p. 97 (https://archive.org/details/bub_gb_l
xkPAAAAIAAJ/page/n101).
20. Grendler, P. (1995). M.A. Lavin (ed.). What Piero Learned in School: Fifteenth-Century
Vernacular Education. pp. 161–176. {{cite book}}: |work= ignored (help)
21. Alberti, Leon Battista; Grayson, Cecil (trans.) (1991). Kemp, Martin (ed.). On Painting.
Penguin Classics.
22. Peterson, Mark. "The Geometry of Piero della Francesca" (http://www.mtholyoke.edu/course
s/rschwart/mac/Italian/geometry.shtml). "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."
23. Hockney, David (2006). Secret Knowledge: Rediscovering the Lost Techniques of the Old
Masters. Thames and Hudson. ISBN 978-0-500-28638-8.
24. Van Riper, Frank. "Hockney's 'Lucid' Bomb At the Art Establishment" (https://www.washingto
npost.com/wp-srv/photo/essays/vanRiper/030220.htm). The Washington Post. Retrieved
4 September 2015.
25. Marr, Andrew (7 October 2001). "What the eye didn't see" (https://www.theguardian.com/theo
bserver/2001/oct/07/featuresreview.review1). The Guardian. Retrieved 4 September 2015.
26. Janson, Jonathan (25 April 2003). "An Interview with Philip Steadman" (http://www.essential
vermeer.com/interviews_newsletter/steadman_interview.html#.VeqxKJdUWHg). Essential
Vermeer. Retrieved 5 September 2015.
27. Steadman, Philip (2002). Vermeer's Camera: Uncovering the Truth Behind the Masterpieces
(https://archive.org/details/vermeerscameraun0000stea). Oxford. ISBN 978-0-19-280302-3.
28. Hart, George. "Luca Pacioli's Polyhedra" (http://www.georgehart.com/virtual-polyhedra/pacio
li.html). Retrieved 13 August 2009.
29. Morris, Roderick Conway (27 January 2006). "Palmezzano's Renaissance:From shadows,
painter emerges" (https://www.nytimes.com/2006/01/27/arts/27iht-conway.html?pagewanted
=all&_r=0). New York Times. Retrieved 22 July 2015.
30. Calter, Paul. "Geometry and Art Unit 1" (http://www.dartmouth.edu/~matc/math5.geometry/uni
t14/unit14.html). Dartmouth College. Retrieved 13 August 2009.
31. Brizio, Anna Maria (1980). Leonardo the Artist (https://archive.org/details/leonardoartist00bri
z). McGraw-Hill. ISBN 9780070079311.
32. Ladwein, Michael (2006). Leonardo Da Vinci, the Last Supper: A Cosmic Drama and an Act
of Redemption (https://books.google.com/books?id=wMFem_x1M04C&pg=PA62). Temple
Lodge Publishing. pp. 61–62. ISBN 978-1-902636-75-7.
33. Turner, Richard A. (1992). Inventing Leonardo (https://archive.org/details/inventingleonard00
turn). Alfred A. Knopf. ISBN 9780679415510.
34. Wolchover, Natalie (31 January 2012). "Did Leonardo da Vinci copy his famous 'Vitruvian
Man'?" (http://www.nbcnews.com/id/46204318/ns/technology_and_science-science/#.Vi_b7
msvuHg). NBC News. Retrieved 27 October 2015.
35. Criminisi, A.; Kempz, M.; Kang, S. B. (2004). "Reflections of Reality in Jan van Eyck and
Robert Campin" (http://research.microsoft.com/pubs/72436/Criminisi_ReflectionsOfReality_
2003.pdf) (PDF). Historical Methods. 37 (3): 109–121. doi:10.3200/hmts.37.3.109-122 (http
s://doi.org/10.3200%2Fhmts.37.3.109-122). S2CID 14289312 (https://api.semanticscholar.or
g/CorpusID:14289312).
36. Cucker, Felipe (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.
37. Cucker, Felipe (2013). Manifold Mirrors: The Crossing Paths of the Arts and Mathematics.
Cambridge University Press. pp. 269–278. ISBN 978-0-521-72876-8.
38. Joyce, David E. (1996). "Euclid's Elements, Book II, Proposition 11" (http://aleph0.clarku.ed
u/~djoyce/elements/bookII/propII11.html). Clark University. Retrieved 24 September 2015.
39. 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 (https://doi.org/10.1097%2F00006534-196410000-00007).
S2CID 70643014 (https://api.semanticscholar.org/CorpusID:70643014).
40. Mainzer, Klaus (1996). Symmetries of Nature: A Handbook for Philosophy of Nature and
Science. Walter de Gruyter. p. 118.
41. "Mathematical properties in ancient theatres and amphitheatres" (https://web.archive.org/we
b/20170715212252/http://mathsforeurope.digibel.be/amphi.htm). Archived from the original
(http://mathsforeurope.digibel.be/amphi.htm) on 15 July 2017. Retrieved 29 January 2014.
42. "Architecture: Ellipse?" (http://www.the-colosseum.net/architecture/ellipsis.htm). The-
Colosseum.net. Retrieved 29 January 2014.
43. Markowsky, George (January 1992). "Misconceptions about the Golden Ratio" (https://web.a
rchive.org/web/20080408200850/http://www.math.nus.edu.sg/aslaksen/teaching/maa/marko
wsky.pdf) (PDF). The College Mathematics Journal. 23 (1): 2–19. doi:10.2307/2686193 (http
s://doi.org/10.2307%2F2686193). JSTOR 2686193 (https://www.jstor.org/stable/2686193).
Archived from the original (http://www.math.nus.edu.sg/aslaksen/teaching/maa/markowsky.p
df) (PDF) on 2008-04-08. Retrieved 2015-06-26.
44. Taseos, Socrates G. (1990). Back in Time 3104 B.C. to the Great Pyramid. SOC Publishers.
45. Gazale, Midhat (1999). Gnomon: From Pharaohs to Fractals. p. 523.
Bibcode:1999EJPh...20..523G (https://ui.adsabs.harvard.edu/abs/1999EJPh...20..523G).
doi:10.1088/0143-0807/20/6/501 (https://doi.org/10.1088%2F0143-0807%2F20%2F6%2F50
1). ISBN 978-0-691-00514-0. {{cite book}}: |journal= ignored (help)
46. Huntley, H.E. (1970). The Divine Proportion (https://archive.org/details/divineproportion0000
hunt). Dover.
47. Hemenway, Priya (2005). Divine Proportion: Phi In Art, Nature, and Science. Sterling. p. 96.
48. Usvat, Liliana. "Mathematics of the Parthenon" (http://www.mathematicsmagazine.com/Articl
es/Mathematics_ofTheParthenon.php#.VYqGl0Z0dIQ). Mathematics Magazine. Retrieved
24 June 2015.
49. Boussora, Kenza; Mazouz, Said (Spring 2004). "The Use of the Golden Section in the Great
Mosque of Kairouan" (https://doi.org/10.1007%2Fs00004-004-0002-y). Nexus Network
Journal. 6 (1): 7–16. doi:10.1007/s00004-004-0002-y (https://doi.org/10.1007%2Fs00004-00
4-0002-y). "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."
50. Brinkworth, Peter; Scott, Paul (2001). "The Place of Mathematics". Australian Mathematics
Teacher. 57 (3): 2.
51. Chanfón Olmos, Carlos (1991). Curso sobre Proporción. Procedimientos reguladors en
construcción. Convenio de intercambio Unam–Uady. México – Mérica.
52. Livio, Mario (2002). The Golden Ratio: The Story of Phi, The World's Most Astonishing
Number (https://archive.org/details/goldenratio00mari). Bibcode:2002grsp.book.....L (https://u
i.adsabs.harvard.edu/abs/2002grsp.book.....L). ISBN 9780767908160. {{cite book}}:
|journal= ignored (help)
53. Smith, Norman A. F. (2001). "Cathedral Studies: Engineering or History" (https://www.webcit
ation.org/6dh0wipJN?url=http://pubs-newcomen.com/tfiles/73ap095.pdf) (PDF).
Transactions of the Newcomen Society. 73: 95–137. doi:10.1179/tns.2001.005 (https://doi.or
g/10.1179%2Ftns.2001.005). S2CID 110300481 (https://api.semanticscholar.org/CorpusID:1
10300481). Archived from the original (http://pubs-newcomen.com/tfiles/73ap095.pdf) (PDF)
on 2015-12-11.
54. McVeigh, Karen (28 December 2009). "Why golden ratio pleases the eye: US academic
says he knows art secret" (https://www.theguardian.com/artanddesign/2009/dec/28/golden-r
atio-us-academic). The Guardian. Retrieved 27 October 2015.
55. Aarts, J.; Fokkink, R.; Kruijtzer, G. (2001). "Morphic numbers" (http://www.nieuwarchief.nl/ser
ie5/pdf/naw5-2001-02-1-056.pdf) (PDF). Nieuw Arch. Wiskd. 5. 2 (1): 56–58.
56. Padovan, Richard (2002). Williams, Kim; Francisco Rodrigues, Jose (eds.). "Dom Hans Van
Der Laan And The Plastic Number" (http://www.nexusjournal.com/conferences/N2002-Pado
van.html). Nexus IV: Architecture and Mathematics: 181–193.
57. Salingaros, Nikos (November 1996). "The 'life' of a carpet: an application of the Alexander
rules" (http://zeta.math.utsa.edu/~yxk833/life.carpet.html). 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.
58. Cucker, Felipe (2013). Manifold Mirrors: The Crossing Paths of the Arts and Mathematics.
Cambridge University Press. pp. 89–102. ISBN 978-0-521-72876-8.
59. Lerner, Martin (1984). The flame and the lotus: Indian and Southeast Asian art from the
Kronos collections (http://libmma.contentdm.oclc.org/cdm/compoundobject/collection/p1532
4coll10/id/105494) (Exhibition Catalogue ed.). Metropolitan Museum of Art.
60. Ellison, Elaine; Venters, Diana (1999). Mathematical Quilts: No Sewing Required. Key
Curriculum.
61. Castera, Jean Marc; Peuriot, Francoise (1999). Arabesques. Decorative Art in Morocco. Art
Creation Realisation. ISBN 978-2-86770-124-5.
62. Cucker, Felipe (2013). Manifold Mirrors: The Crossing Paths of the Arts and Mathematics.
Cambridge University Press. pp. 103–106. ISBN 978-0-521-72876-8.
63. Dye, Daniel S. (1974). Chinese Lattice Designs (https://archive.org/details/chineselatticede0
0dyed). Dover. pp. 30–39 (https://archive.org/details/chineselatticede00dyed/page/30).
ISBN 9780486230962.
64. belcastro, sarah-marie (2013). "Adventures in Mathematical Knitting" (http://www.americansc
ientist.org/issues/feature/adventures-in-mathematical-knitting/1). American Scientist. 101 (2):
124. doi:10.1511/2013.101.124 (https://doi.org/10.1511%2F2013.101.124).
65. Taimina, Daina (2009). Crocheting Adventures with Hyperbolic Planes. A K Peters.
ISBN 978-1-56881-452-0.
66. Snook, Barbara. Florentine Embroidery. Scribner, Second edition 1967.
67. Williams, Elsa S. Bargello: Florentine Canvas Work. Van Nostrand Reinhold, 1967.
68. Grünbaum, Branko; Shephard, Geoffrey C. (May 1980). "Satins and Twills: An Introduction to
the Geometry of Fabrics". Mathematics Magazine. 53 (3): 139–161. doi:10.2307/2690105 (ht
tps://doi.org/10.2307%2F2690105). hdl:10338.dmlcz/104026 (https://hdl.handle.net/10338.d
mlcz%2F104026). JSTOR 2690105 (https://www.jstor.org/stable/2690105).
69. Gamwell, Lynn (2015). Mathematics and Art: A Cultural History. Princeton University Press.
p. 423. ISBN 978-0-691-16528-8.
70. Baker, Patricia L.; Smith, Hilary (2009). Iran (https://books.google.com/books?id=a40CkMNq
U8AC&pg=PA108) (3 ed.). Bradt Travel Guides. p. 107. ISBN 978-1-84162-289-7.
71. 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 (https://arxiv.o
rg/abs/1406.1532). Bibcode:2014arXiv1406.1532I (https://ui.adsabs.harvard.edu/abs/2014ar
Xiv1406.1532I). doi:10.1080/17513472.2014.982938 (https://doi.org/10.1080%2F17513472.
2014.982938). S2CID 119168759 (https://api.semanticscholar.org/CorpusID:119168759).
72. 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 (http
s://ui.adsabs.harvard.edu/abs/2007Sci...315.1106L). doi:10.1126/science.1135491 (https://d
oi.org/10.1126%2Fscience.1135491). PMID 17322056 (https://pubmed.ncbi.nlm.nih.gov/173
22056). S2CID 10374218 (https://api.semanticscholar.org/CorpusID:10374218).
73. van den Hoeven, Saskia; van der Veen, Maartje. "Muqarnas-Mathematics in Islamic Arts" (htt
ps://web.archive.org/web/20130927070005/http://www.wiskuu.nl/muqarnas/Muqarnas_engli
sh8.pdf) (PDF). Archived from the original (http://www.wiskuu.nl/muqarnas/Muqarnas_englis
h8.pdf) (PDF) on 27 September 2013. Retrieved 15 January 2016.
74. Markowsky, George (March 2005). "Book review: The Golden Ratio" (https://www.ams.org/n
otices/200503/rev-markowsky.pdf) (PDF). Notices of the American Mathematical Society. 52
(3): 344–347.
75. Panofsky, E. (1955). The Life and Art of Albrecht Durer. Princeton.
76. Hart, George W. "Dürer's Polyhedra" (http://www.georgehart.com/virtual-polyhedra/durer.htm
l). Retrieved 13 August 2009.
77. Dürer, Albrecht (1528). Hierinn sind begriffen vier Bucher von menschlicher Proportion (http
s://archive.org/details/hierinnsindbegri00dure). Nuremberg. Retrieved 24 June 2015.
78. Schreiber, P. (1999). "A New Hypothesis on Durer's Enigmatic Polyhedron in His Copper
Engraving 'Melencolia I' " (https://doi.org/10.1006%2Fhmat.1999.2245). Historia
Mathematica. 26 (4): 369–377. doi:10.1006/hmat.1999.2245 (https://doi.org/10.1006%2Fhm
at.1999.2245).
79. Dodgson, Campbell (1926). Albrecht Dürer. London: Medici Society. p. 94.
80. Schuster, Peter-Klaus (1991). Melencolia I: Dürers Denkbild. Berlin: Gebr. Mann Verlag.
pp. 17–83.
81. Panofsky, Erwin; Klibansky, Raymond; Saxl, Fritz (1964). Saturn and Melancholy (http://quo
d.lib.umich.edu/cgi/t/text/text-idx?c=genpub;cc=genpub;q1=Klibansky;op2=and;op3=and;rgn
=works;rgn1=author;rgn2=author;rgn3=author;view=toc;idno=0431529.0001.001). Basic
Books.
82. Rucker, Rudy (2014). The Fourth Dimension: Toward a Geometry of Higher Reality (https://b
ooks.google.com/books?id=0xReBAAAQBAJ). Courier Corporation. ISBN 978-0-486-
79819-6.
83. "Crucifixion (Corpus Hypercubus)" (http://www.metmuseum.org/collection/the-collection-onli
ne/search/488880). Metropolitan Museum of Art. Retrieved 5 September 2015.
84. Macdonald, Fiona (11 May 2016). "The painter who entered the fourth dimension" (https://w
ww.bbc.com/culture/article/20160511-the-painter-who-entered-the-fourth-dimension). BBC.
Retrieved 8 February 2022.
85. Lukman, Muhamad; Hariadi, Yun; Destiarmand, Achmad Haldani (2007). "Batik Fractal:
Traditional Art to Modern Complexity". Proceeding Generative Art X, Milan, Italy.
86. Ouellette, Jennifer (November 2001). "Pollock's Fractals" (http://discovermagazine.com/200
1/nov/featpollock). Discover Magazine. Retrieved 26 September 2016.
87. Galilei, Galileo (1623). The Assayer., as translated in Drake, Stillman (1957). Discoveries
and Opinions of Galileo (https://archive.org/details/discoveriesopini00gali_0/page/237).
Doubleday. pp. 237–238 (https://archive.org/details/discoveriesopini00gali_0/page/237).
ISBN 978-0-385-09239-5.
88. Cucker, Felipe (2013). Manifold Mirrors: The Crossing Paths of the Arts and Mathematics.
Cambridge University Press. p. 381. ISBN 978-0-521-72876-8.
89. Cucker, Felipe (2013). Manifold Mirrors: The Crossing Paths of the Arts and Mathematics.
Cambridge University Press. p. 10. ISBN 978-0-521-72876-8.
90. King, Jerry P. (1992). The Art of Mathematics. Fawcett Columbine. pp. 8–9. ISBN 978-0-449-
90835-8.
91. King, Jerry P. (1992). The Art of Mathematics. Fawcett Columbine. pp. 135–139. ISBN 978-
0-449-90835-8.
92. Devlin, Keith (2000). "Do Mathematicians Have Different Brains?" (https://books.google.co
m/books?id=AJdmfYEaLG4C). The Math Gene: How Mathematical Thinking Evolved And
Why Numbers Are Like Gossip. Basic Books. p. 140. ISBN 978-0-465-01619-8.
93. Wasilewska, Katarzyna (2012). "Mathematics in the World of Dance" (http://archive.bridgesm
athart.org/2012/bridges2012-453.pdf) (PDF). Bridges. Retrieved 1 September 2015.
94. Malkevitch, Joseph. "Mathematics and Art" (https://www.ams.org/samplings/feature-column/f
carc-art1). American Mathematical Society. Retrieved 1 September 2015.
95. Malkevitch, Joseph. "Mathematics and Art. 2. Mathematical tools for artists" (https://www.am
s.org/samplings/feature-column/fcarc-art2). American Mathematical Society. Retrieved
1 September 2015.
96. "Math and Art: The Good, the Bad, and the Pretty" (http://www.maa.org/meetings/calendar-ev
ents/math-and-art-the-good-the-bad-and-the-pretty). Mathematical Association of America.
Retrieved 2 September 2015.
97. Cohen, Louise (1 July 2014). "How to spin the colour wheel, by Turner, Malevich and more"
(http://www.tate.org.uk/context-comment/articles/how-to-spin-the-colour-wheel). Tate Gallery.
Retrieved 4 September 2015.
98. 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.
99. Gage, John (1999). Color and Culture: Practice and Meaning from Antiquity to Abstraction (ht
tps://books.google.com/books?id=oq_GtjmoTNgC&pg=PA207). University of California
Press. p. 207. ISBN 978-0-520-22225-0.
100. Malkevitch, Joseph. "Mathematics and Art. 3. Symmetry" (https://www.ams.org/samplings/fea
ture-column/fcarc-art3). American Mathematical Society. Retrieved 1 September 2015.
101. Malkevitch, Joseph. "Mathematics and Art. 4. Mathematical artists and artist mathematicians"
(https://www.ams.org/samplings/feature-column/fcarc-art4). American Mathematical Society.
Retrieved 1 September 2015.
102. 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 (https://doi.org/10.2307%2F1557919). JSTOR 1557919 (https://www.j
stor.org/stable/1557919).
103. Kalajdzievski, Sasho (2008). Math and Art: An Introduction to Visual Mathematics. Chapman
and Hall. ISBN 978-1-58488-913-7.
104. Beddard, Honor (2011-05-26). "Computer art at the V&A" (http://www.vam.ac.uk/content/jour
nals/research-journal/issue-02/computer-art-at-the-v-and-a/). Victoria and Albert Museum.
Retrieved 22 September 2015.
105. "Computer Does Drawings: Thousands of lines in each". The Guardian. 17 September
1962. in Beddard, 2015.
106. 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.
107. Bellos, Alex (24 February 2015). "Catch of the day: mathematician nets weird, complex fish"
(https://www.theguardian.com/science/alexs-adventures-in-numberland/2015/feb/24/catch-of
-the-day-mathematician-nets-weird-complex-fish). The Guardian. Retrieved 25 September
2015.
108. " "A Bird in Flight (2016)," by Hamid Naderi Yeganeh" (https://www.ams.org/mathimagery/dis
playimage.php?album=40&pid=684#top_display_media). American Mathematical Society.
March 23, 2016. Retrieved April 6, 2017.
109. Chung, Stephy (September 18, 2015). "Next da Vinci? Math genius using formulas to create
fantastical works of art" (http://www.cnn.com/2015/09/17/arts/math-art/). CNN.
110. Levin, Golan (2013). "Generative Artists" (http://cmuems.com/2013/a/resources/artists-gener
ative/). CMUEMS. Retrieved 27 October 2015. This includes a link to Hvidtfeldts Syntopia (h
ttp://blog.hvidtfeldts.net/index.php/generative-art-links/).
111. Verostko, Roman. "The Algorists" (http://www.verostko.com/algorist.html). Retrieved
27 October 2015.
112. Cucker, Felipe (2013). Manifold Mirrors: The Crossing Paths of the Arts and Mathematics.
Cambridge University Press. pp. 315–317. ISBN 978-0-521-72876-8.
113. Miller, Arthur I. (2001). Einstein, Picasso: Space, Time, and the Beauty That Causes Havoc
(https://archive.org/details/einsteinpicassos00mill). New York: Basic Books. p. 171 (https://ar
chive.org/details/einsteinpicassos00mill/page/171). ISBN 978-0-465-01860-4.
114. Miller, Arthur I. (2012). Insights of Genius: Imagery and Creativity in Science and Art.
Springer. ISBN 978-1-4612-2388-7.
115. Henderson, Linda Dalrymple (1983). The Fourth Dimension and Non-Euclidean geometry in
Modern Art. Princeton University Press.
116. Antliff, Mark; Leighten, Patricia Dee (2001). Cubism and Culture (https://web.archive.org/we
b/20200726154229/http://eres.lndproxy.org/edoc/AH317Antliff-09.pdf) (PDF). Thames &
Hudson. Archived from the original (http://eres.lndproxy.org/edoc/AH317Antliff-09.pdf) (PDF)
on 26 July 2020.
117. Everdell, William R. (1997). The First Moderns: Profiles in the Origins of Twentieth-Century
Thought (https://archive.org/details/firstmodernsprof00ever/page/312). University of Chicago
Press. p. 312 (https://archive.org/details/firstmodernsprof00ever/page/312). ISBN 978-0-226-
22480-0.
118. Green, Christopher (1987). Cubism and its Enemies, Modern Movements and Reaction in
French Art, 1916–1928. Yale University Press. pp. 13–47.
119. Metzinger, Jean (October–November 1910). "Note sur la peinture". Pan: 60. in Miller (2001).
Einstein, Picasso (https://archive.org/details/einsteinpicassos00mill). Basic Books. p. 167 (ht
tps://archive.org/details/einsteinpicassos00mill/page/167). ISBN 9780465018598.
120. 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 (https://archive.org/
details/homoaestheticusi0000ferr/page/215). Robert De Loaiza, trans. University of Chicago
Press. p. 215 (https://archive.org/details/homoaestheticusi0000ferr/page/215). ISBN 978-0-
226-24459-4.
121. "Man Ray–Human Equations A Journey from Mathematics to Shakespeare. February 7 –
May 10, 2015" (http://www.phillipscollection.org/events/2015-02-07-exhibition-man-ray-hum
an-equations). Phillips Collection. 7 February 2015. Retrieved 5 September 2015.
122. Adcock, Craig (1987). "Duchamp's Eroticism: A Mathematical Analysis" (http://ir.uiowa.edu/c
gi/viewcontent.cgi?article=1208&context=dadasur). Iowa Research Online. 16 (1): 149–167.
123. Elder, R. Bruce (2013). DADA, Surrealism, and the Cinematic Effect (https://books.google.co
m/books?id=mhXaAgAAQBAJ&pg=PA602). Wilfrid Laurier University Press. p. 602.
ISBN 978-1-55458-641-7.
124. Tubbs, Robert (2014). Mathematics in Twentieth-Century Literature and Art: Content, Form,
Meaning (https://books.google.com/books?id=h1vBAwAAQBAJ&pg=PA118). JHU Press.
p. 118. ISBN 978-1-4214-1402-7.
125. "Hiroshi Sugimoto Conceptual Forms and Mathematical Models February 7 – May 10, 2015"
(http://www.phillipscollection.org/events/2015-02-07-exhibition-hiroshi-sugimoto). Phillips
Collection. 7 February 2015. Retrieved 5 September 2015.
126. Tubbs, Robert (2014). Mathematics in 20th-Century Literature and Art. Johns Hopkins.
pp. 8–10. ISBN 978-1-4214-1380-8.
127. Keats, Jonathon (13 February 2015). "See How Man Ray Made Elliptic Paraboloids Erotic
At This Phillips Collection Photography Exhibit" (https://www.forbes.com/sites/jonathonkeat
s/2015/02/13/see-how-man-ray-made-elliptic-paraboloids-sexy-at-this-phillips-collection-ph
otography-exhibit/). Forbes. Retrieved 10 September 2015.
128. Gamwell, Lynn (2015). Mathematics and Art: A Cultural History. Princeton University Press.
pp. 311–312. ISBN 978-0-691-16528-8.
129. Hedgecoe, John, ed. (1968). Henry Moore: Text on His Sculpture. p. 105. {{cite book}}:
|work= ignored (help)
130. "De Stijl" (http://www.tate.org.uk/learn/online-resources/glossary/d/de-stijl). Tate Glossary.
The Tate. Retrieved 11 September 2015.
131. Curl, James Stevens (2006). A Dictionary of Architecture and Landscape Architecture (http
s://archive.org/details/dictionaryofarch00curl_0) (Second ed.). Oxford University Press.
ISBN 978-0-19-860678-9.
132. Tubbs, Robert (2014). Mathematics in Twentieth-Century Literature and Art: Content, Form,
Meaning. JHU Press. pp. 44–47. ISBN 978-1-4214-1402-7.
133. "Tour: M.C. Escher – Life and Work" (https://web.archive.org/web/20090803124602/http://ww
w.nga.gov/collection/gallery/ggescher/ggescher-main1.html). NGA. Archived from the
original (http://www.nga.gov/collection/gallery/ggescher/ggescher-main1.html) on 3 August
2009. Retrieved 13 August 2009.
134. "MC Escher" (http://www.mathacademy.com/pr/minitext/escher/). Mathacademy.com. 1
November 2007. Retrieved 13 August 2009.
135. 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 (http
s://doi.org/10.1111%2Fj.2044-8295.1958.tb00634.x). PMID 13536303 (https://pubmed.ncbi.
nlm.nih.gov/13536303).
136. Kirousis, Lefteris M.; Papadimitriou, Christos H. (1985). "The complexity of recognizing
polyhedral scenes". 26th Annual Symposium on Foundations of Computer Science (SFCS
1985). pp. 175–185. CiteSeerX 10.1.1.100.4844 (https://citeseerx.ist.psu.edu/viewdoc/summ
ary?doi=10.1.1.100.4844). doi:10.1109/sfcs.1985.59 (https://doi.org/10.1109%2Fsfcs.1985.5
9). ISBN 978-0-8186-0644-1.
137. Cooper, Martin (2008). "Tractability of Drawing Interpretation". Inequality, Polarization and
Poverty (https://archive.org/details/linedrawinginter00coop_873). Springer-Verlag. pp. 217 (h
ttps://archive.org/details/linedrawinginter00coop_873/page/n218)–230. doi:10.1007/978-1-
84800-229-6_9 (https://doi.org/10.1007%2F978-1-84800-229-6_9). ISBN 978-1-84800-229-
6.
138. Roberts, Siobhan (2006). 'Coxetering' with M.C. Escher. p. Chapter 11. {{cite book}}:
|work= ignored (help)
139. Escher, M.C. (1988). The World of MC Escher. Random House.
140. Escher, M.C.; Vermeulen, M.W.; Ford, K. (1989). Escher on Escher: Exploring the Infinite. HN
Abrams.
141. Malkevitch, Joseph. "Mathematics and Art. 5. Polyhedra, tilings, and dissections" (https://ww
w.ams.org/samplings/feature-column/fcarc-art5). American Mathematical Society. Retrieved
1 September 2015.
142. Marcolli, Matilde (July 2016). The notion of Space in Mathematics through the lens of
Modern Art (http://www.its.caltech.edu/~matilde/SpaceMathArticle.pdf) (PDF). Century
Books. pp. 23–26.
143. "John Robinson" (http://www.bradshawfoundation.com/jr/). Bradshaw Foundation. 2007.
Retrieved 13 August 2009.
144. "Helaman Ferguson web site" (https://web.archive.org/web/20090411034819/http://www.hel
asculpt.com/index.html). Helasculpt.com. Archived from the original (http://www.helasculpt.c
om/index.html) on 11 April 2009. Retrieved 13 August 2009.
145. Thurston, William P. (1999). Levy, Silvio (ed.). The Eightfold Way: A Mathematical Sculpture
by Helaman Ferguson (http://www.msri.org/publications/books/Book35/files/thurston.pdf)
(PDF). pp. 1–7. {{cite book}}: |work= ignored (help)
146. "MAA book review of The Eightfold Way: The Beauty of Klein's Quartic Curve" (http://www.m
aa.org/reviews/eightfold.html). Maa.org. 14 November 1993. Retrieved 13 August 2009.
147. "The Math Geek Holiday Gift Guide" (http://blogs.scientificamerican.com/roots-of-unity/the-m
ath-geek-holiday-gift-guide/). Scientific American. 23 November 2014. Retrieved 7 June
2015.
148. Hanna, Raven. "Gallery: Bathsheba Grossman" (http://www.symmetrymagazine.org/article/s
eptember-2005/gallery-bathsheba-grossman). Symmetry Magazine. Retrieved 7 June 2015.
149. Mastroianni, Brian (26 May 2015). "The perfect equation: Artist combines math and art" (http
s://www.foxnews.com/science/the-perfect-equation-artist-combines-math-and-art). Fox
News. Retrieved 28 January 2021.
150. Fleron, Julian F.; Ecke, Volker; von Renesse, Christine; Hotchkiss, Philip K. (January 2015).
Art and Sculpture: Mathematical Inquiry in the Liberal Arts (https://www.artofmathematics.org/
books/art-and-sculpture) (2nd ed.). Discovering the Art of Mathematics project.
151. Osinga, Hinke (2005). "Crocheting the Lorenz manifold" (https://web.archive.org/web/20150
410054845/https://www.math.auckland.ac.nz/~hinke/crochet/). University of Auckland.
Archived from the original (https://www.math.auckland.ac.nz/~hinke/crochet/) on 10 April
2015. Retrieved 12 October 2015.
152. Osinga, Hinke M.; Krauskopf, Bernd (2004). "Crocheting the Lorenz manifold" (http://www.en
m.bris.ac.uk/anm/preprints/2004r03.html). The Mathematical Intelligencer. 26 (4): 25–37.
CiteSeerX 10.1.1.108.4594 (https://citeseerx.ist.psu.edu/viewdoc/summary?doi=10.1.1.108.
4594). doi:10.1007/BF02985416 (https://doi.org/10.1007%2FBF02985416).
S2CID 119728638 (https://api.semanticscholar.org/CorpusID:119728638).
153. Dietz, Ada K. (1949). Algebraic Expressions in Handwoven Textiles (https://web.archive.org/
web/20160222003421/http://www.cs.arizona.edu/patterns/weaving/monographs/dak_alge.p
df) (PDF). Louisville, Kentucky: The Little Loomhouse. Archived from the original (http://www.
cs.arizona.edu/patterns/weaving/monographs/dak_alge.pdf) (PDF) on 2016-02-22.
Retrieved 2015-06-26.
154. Henderson, David; Taimina, Daina (2001). "Crocheting the hyperbolic plane" (http://www.ma
th.cornell.edu/%7Edwh/papers/crochet/crochet.PDF) (PDF). Mathematical Intelligencer. 23
(2): 17–28. doi:10.1007/BF03026623 (https://doi.org/10.1007%2FBF03026623).
S2CID 120271314 (https://api.semanticscholar.org/CorpusID:120271314)..
155. Barnett, Rebekah (31 January 2017). "Gallery: What happens when you mix math, coral and
crochet? It's mind-blowing" (https://ideas.ted.com/gallery-what-happens-when-you-mix-math
-coral-and-crochet-its-mind-blowing/). Ideas.TED.com. Retrieved 28 October 2019.
156. 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 (https://ui.adsabs.harvard.edu/abs/1970RSPTA.266...6
3M). doi:10.1098/rsta.1970.0003 (https://doi.org/10.1098%2Frsta.1970.0003). JSTOR 73779
(https://www.jstor.org/stable/73779). S2CID 123330469 (https://api.semanticscholar.org/Corp
usID:123330469).
157. "Pat Ashforth & Steve Plummer – Mathekniticians" (http://www.woollythoughts.com/aboutus.
html). Woolly Thoughts. Retrieved 4 October 2015.
158. Ward, Mark (20 August 2012). "Knitting reinvented: Mathematics, feminism and metal" (http
s://www.bbc.co.uk/news/technology-19208292). BBC News. BBC. Retrieved 23 September
2015.
159. Ashforth, Pat; Plummer, Steve. "Menger Sponge" (http://www.woollythoughts.com/menger.ht
ml). Woolly Thoughts: In Pursuit of Crafty Mathematics. Retrieved 23 September 2015.
160. Ashforth, Pat; Plummer, Steve. "Afghans for Schools" (http://www.woollythoughts.com/schoo
ls/index.html). Woolly Thoughts: Mathghans. Retrieved 23 September 2015.
161. "Mathghans with a Difference" (https://web.archive.org/web/20150925133153/http://www.si
mplyknitting.co.uk/2008/07/01/mathghans-with-a-difference/). Simply Knitting Magazine. 1
July 2008. Archived from the original (http://www.simplyknitting.co.uk/2008/07/01/mathghans
-with-a-difference/) on 25 September 2015. Retrieved 23 September 2015.
162. Jouffret, Esprit (1903). Traité élémentaire de géométrie à quatre dimensions et introduction à
la géométrie à n dimensions (http://historical.library.cornell.edu/cgi-bin/cul.math/docviewer?
did=04810001) (in French). Paris: Gauthier-Villars. OCLC 1445172 (https://www.worldcat.or
g/oclc/1445172). Retrieved 26 September 2015.
163. 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.
164. "Giotto di Bondone and assistants: Stefaneschi triptych" (http://mv.vatican.va/3_EN/pages/PI
N/PIN_Sala02_03.html). The Vatican. Retrieved 16 September 2015.
165. Gamwell, Lynn (2015). Mathematics and Art: A Cultural History. Princeton University Press.
pp. 337–338. ISBN 978-0-691-16528-8.
166. Cooper, Jonathan (5 September 2007). "Art and Mathematics" (http://www.doctordada.com/a
rt/art-and-mathematics/). Retrieved 5 September 2015.
167. Hofstadter, Douglas R. (1980). Gödel, Escher, Bach: An Eternal Golden Braid. Penguin.
p. 627. ISBN 978-0-14-028920-6.
168. Hall, James (10 June 2011). "René Magritte: The Pleasure Principle – exhibition" (https://ww
w.theguardian.com/artanddesign/2011/jun/10/rene-magritte-pleasure-principle-exhibition).
The Guardian. Retrieved 5 September 2015.
169. Hofstadter, Douglas R. (1980). Gödel, Escher, Bach: An Eternal Golden Braid. Penguin.
pp. 98–99, 690–717. ISBN 978-0-394-74502-2.
170. King, Elliott (2004). Ades, Dawn (ed.). Dali. Milan: Bompiani Arte. pp. 418–421.
171. "Stone balancing" (http://mei.org.uk/files/pdf/MM_July_2013.pdf) (PDF). Monthly Maths (29).
July 2013. Retrieved 10 June 2017.
172. de Smit, B. (2003). "The Mathematical Structure of Escher's Print Gallery". Notices of the
American Mathematical Society. 50 (4): 446–451.
173. Lenstra, Hendrik; De Smit, Bart. "Applying mathematics to Escher's Print Gallery" (http://esch
erdroste.math.leidenuniv.nl/index.php?menu=intro). Leiden University. Retrieved
10 November 2015.
174. Stanek, Becca (16 June 2014). "Van Gogh and the Algorithm: How Math Can Save Art" (htt
p://time.com/2884058/math-art-van-gogh-duke/). Time Magazine. Retrieved 4 September
2015.
175. Sipics, Michelle (18 May 2009). "The Van Gogh Project: Art Meets Mathematics in Ongoing
International Study" (https://web.archive.org/web/20150907222302/http://www.siam.org/new
s/news.php?id=1568). Society for Industrial and Applied Mathematics. Archived from the
original (https://www.siam.org/news/news.php?id=1568) on 7 September 2015. Retrieved
4 September 2015.
176. Emmerling, Leonhard (2003). Jackson Pollock, 1912–1956 (https://books.google.com/book
s?id=K-ZZmvjJ_-IC&pg=PA63). Taschen. p. 63. ISBN 978-3-8228-2132-9.
177. Taylor, Richard P.; Micolich, Adam P.; Jonas, David (June 1999). "Fractal analysis of
Pollock's drip paintings" (https://doi.org/10.1038%2F20833). Nature. 399 (6735): 422.
Bibcode:1999Natur.399..422T (https://ui.adsabs.harvard.edu/abs/1999Natur.399..422T).
doi:10.1038/20833 (https://doi.org/10.1038%2F20833). S2CID 204993516 (https://api.sema
nticscholar.org/CorpusID:204993516).
178. Taylor, Richard; Micolich, Adam P.; Jonas, David (October 1999). "Fractal Expressionism:
Can Science Be Used To Further Our Understanding Of Art?" (https://archive.today/2012080
5084052/http://phys.unsw.edu.au/phys_about/PHYSICS!/FRACTAL_EXPRESSIONISM/frac
tal_taylor.html). Physics World. 12 (10): 25–28. doi:10.1088/2058-7058/12/10/21 (https://doi.
org/10.1088%2F2058-7058%2F12%2F10%2F21). Archived from the original (http://phys.un
sw.edu.au/phys_about/PHYSICS!/FRACTAL_EXPRESSIONISM/fractal_taylor.html) 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."
179. King, M. (2002). "From Max Ernst to Ernst Mach: epistemology in art and science" (https://w
ww.herts.ac.uk/__data/assets/pdf_file/0013/12307/WPIAAD_vol2_king.pdf) (PDF).
Retrieved 17 September 2015.
180. Dodgson, N. A. (2012). "Mathematical characterisation of Bridget Riley's stripe paintings" (htt
p://www.cl.cam.ac.uk/~nad10/pubs/jma12.pdf) (PDF). Journal of Mathematics and the Arts. 5
(2–3): 89–106. doi:10.1080/17513472.2012.679468 (https://doi.org/10.1080%2F17513472.2
012.679468). S2CID 10349985 (https://api.semanticscholar.org/CorpusID:10349985). "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."
181. Cucker, Felipe (2013). Manifold Mirrors: The Crossing Paths of the Arts and Mathematics.
Cambridge University Press. pp. 116–120. ISBN 978-0-521-72876-8.
182. Treibergs, Andrejs (24 July 2001). "The Geometry of Perspective Drawing on the Computer"
(http://www.math.utah.edu/~treiberg/Perspect/Perspect.htm). University of Utah. Retrieved
5 September 2015.
183. Gamwell, Lynn (2015). Mathematics and Art: A Cultural History. Princeton University Press.
p. xviii. ISBN 978-0-691-16528-8.
184. Malkevitch, Joseph. "Mathematics and Art. 6. Origami" (https://www.ams.org/samplings/featu
re-column/fcarc-art6). American Mathematical Society. Retrieved 1 September 2015.
185. Rao, T. Sundara (1893). Geometric Exercises in Paper Folding. Addison.
186. Justin, J. (June 1986). "Mathematics of Origami, part 9". British Origami: 28–30..
187. Alsina, Claudi; Nelsen, Roger (2010). Charming Proofs: A Journey Into Elegant Mathematics
(https://books.google.com/books?id=mIT5-BN_L0oC&pg=PA57). Dolciani Mathematical
Expositions. Vol. 42. Mathematical Association of America. p. 57. ISBN 978-0-88385-348-1.
188. Alperin, Roger C.; Lang, Robert J. (2009). "One-, Two-, and Multi-Fold Origami Axioms" (htt
p://www.math.sjsu.edu/~alperin/AlperinLang.pdf) (PDF). 4OSME.
189. The World of Geometric Toys (http://www1.ttcn.ne.jp/~a-nishi/), Origami Spring (http://www1.t
tcn.ne.jp/~a-nishi/spring/z_spring.html), August, 2007.
190. Cucker, Felipe (2013). Manifold Mirrors: The Crossing Paths of the Arts and Mathematics.
Cambridge University Press. pp. 163–166. ISBN 978-0-521-72876-8.
191. Gamwell, Lynn (2015). Mathematics and Art: A Cultural History. Princeton University Press.
pp. 406–410. ISBN 978-0-691-16528-8.
192. Ghyka, Matila (2003). The Geometry of Art and Life (https://archive.org/details/geometryofartli
f00mati). Dover. pp. ix–xi. ISBN 978-0-486-23542-4.
193. Lawlor, Robert (1982). Sacred Geometry: Philosophy and Practice (https://archive.org/detail
s/sacredgeometryph00lawl). Thames & Hudson. ISBN 978-0-500-81030-9.
194. Calter, Paul (1998). "Celestial Themes in Art & Architecture" (https://www.dartmouth.edu/~m
atc/math5.geometry/unit10/unit10.html). Dartmouth College. Retrieved 5 September 2015.
195. Maddocks, Fiona (21 Nov 2014). "The 10 best works by William Blake" (https://www.theguar
dian.com/culture/2014/nov/21/the-10-best-works-by-william-blake). The Guardian. Retrieved
25 December 2019.
196. "William Blake, Newton, 1795–c.1805" (https://web.archive.org/web/20190328134113/http
s://www.tate.org.uk/art/artworks/blake-newton-n05058). Tate. October 2018. Archived from
the original (https://www.tate.org.uk/art/artworks/blake-newton-n05058) on 28 March 2019.
197. Livio, Mario (November 2002). "The golden ratio and aesthetics" (https://plus.maths.org/cont
ent/golden-ratio-and-aesthetics). Retrieved 26 June 2015.

External links
Bridges Organization (http://www.bridgesmathart.org/) conference on connections between
art and mathematics
Bridging the Gap Between Math and Art (http://www.scientificamerican.com/slideshow/bridgi
ng-the-gap/) – Slide Show from Scientific American
Discovering the Art of Mathematics (https://www.artofmathematics.org/)
Mathematics and Art (https://www.ams.org/samplings/feature-column/fcarc-art1) – AMS
Mathematics and Art (http://www.cut-the-knot.org/ctk/ArtMath.shtml) – Cut-the-Knot
Mathematical Imagery (https://www.ams.org/mathimagery/) – American Mathematical
Society
Mathematics in Art and Architecture (https://web.archive.org/web/20150507151115/http://ww
w.math.nus.edu.sg/aslaksen/teaching/math-art-arch.shtml) – National University of
Singapore
Mathematical Art (http://virtualmathmuseum.org/mathart/MathematicalArt.html) – Virtual Math
Museum
When art and math collide (https://www.sciencenews.org/article/when-art-and-math-collide)
– Science News
Why the history of maths is also the history of art (https://www.theguardian.com/science/alex
s-adventures-in-numberland/2015/dec/02/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=1173728804"

You might also like