Golden ratio

From Justapedia, unleashing the power of collective wisdom
Jump to navigation Jump to search
Golden ratio
Golden ratio line.svg
Line segments in the golden ratio
Representations
Decimal1.618033988749894...[1]
Algebraic form
Continued fraction
Binary1.10011110001101110111...
Hexadecimal1.9E3779B97F4A7C15...
A golden rectangle with long side a and short side b (shaded red, right) and a square with sides of length a (shaded blue, left) combine to form a similar golden rectangle with long side a + b and short side a. This illustrates the relationship

In mathematics, two quantities are in the golden ratio if their ratio is the same as the ratio of their sum to the larger of the two quantities. Expressed algebraically, for quantities and with

where the Greek letter phi ( or ) denotes the golden ratio.[a] The constant satisfies the quadratic equation and is an irrational number with a value of[1]

1.618033988749....

The golden ratio was called the extreme and mean ratio by Euclid,[2] and the divine proportion by Luca Pacioli,[3] and also goes by several other names.[b]

Mathematicians have studied the golden ratio's properties since antiquity. It is the ratio of a regular pentagon's diagonal to its side, and thus appears in the construction of the dodecahedron and icosahedron.[7] A golden rectangle—that is, a rectangle with an aspect ratio of —may be cut into a square and a smaller rectangle with the same aspect ratio. The golden ratio has been used to analyze the proportions of natural objects as well as artificial systems such as financial markets, in some cases based on dubious fits to data.[8] The golden ratio appears in some patterns in nature, including the spiral arrangement of leaves and other parts of vegetation.

Some 20th-century artists and architects, including Le Corbusier and Salvador Dalí, have proportioned their works to approximate the golden ratio, believing it to be aesthetically pleasing. These uses often appear in the form of a golden rectangle.

Calculation

Two quantities and are in the golden ratio if[9]

One method for finding 's closed form starts with the left fraction. Simplifying the fraction and substituting the reciprocal ,

Therefore,

Multiplying by gives

which can be rearranged to

The quadratic formula yields two solutions:

and

Because is a ratio between positive quantities, is necessarily the positive root.[10] The negative root is in fact the negative inverse , which shares many properties with the golden ratio.

History

According to Mario Livio,

Some of the greatest mathematical minds of all ages, from Pythagoras and Euclid in ancient Greece, through the medieval Italian mathematician Leonardo of Pisa and the Renaissance astronomer Johannes Kepler, to present-day scientific figures such as Oxford physicist Roger Penrose, have spent endless hours over this simple ratio and its properties. ... Biologists, artists, musicians, historians, architects, psychologists, and even mystics have pondered and debated the basis of its ubiquity and appeal. In fact, it is probably fair to say that the Golden Ratio has inspired thinkers of all disciplines like no other number in the history of mathematics.[11]

— The Golden Ratio: The Story of Phi, the World's Most Astonishing Number

Ancient Greek mathematicians first studied the golden ratio because of its frequent appearance in geometry;[12] the division of a line into "extreme and mean ratio" (the golden section) is important in the geometry of regular pentagrams and pentagons.[13] According to one story, 5th-century BC mathematician Hippasus discovered that the golden ratio was neither a whole number nor a fraction (an irrational number), surprising Pythagoreans.[14] Euclid's Elements (c. 300 BC) provides several propositions and their proofs employing the golden ratio,[15][c] and contains its first known definition which proceeds as follows:[16]

A straight line is said to have been cut in extreme and mean ratio when, as the whole line is to the greater segment, so is the greater to the lesser.[17][d]

Michael Maestlin, the first to write a decimal approximation of the ratio

The golden ratio was studied peripherally over the next millennium. Abu Kamil (c. 850–930) employed it in his geometric calculations of pentagons and decagons; his writings influenced that of Fibonacci (Leonardo of Pisa) (c. 1170–1250), who used the ratio in related geometry problems but did not observe that it was connected to the Fibonacci numbers.[19]

Luca Pacioli named his book Divina proportione (1509) after the ratio; the book, largely plagiarized from Piero della Francesca, explored its properties including its appearance in some of the Platonic solids.[20][21] Leonardo da Vinci, who illustrated Pacioli's book, called the ratio the sectio aurea ('golden section').[22] Though it is often said that Pacioli advocated the golden ratio's application to yield pleasing, harmonious proportions, Livio points out that the interpretation has been traced to an error in 1799, and that Pacioli actually advocated the Vitruvian system of rational proportions.[23] Pacioli also saw Catholic religious significance in the ratio, which led to his work's title. 16th-century mathematicians such as Rafael Bombelli solved geometric problems using the ratio.[24]

German mathematician Simon Jacob (d. 1564) noted that consecutive Fibonacci numbers converge to the golden ratio;[25] this was rediscovered by Johannes Kepler in 1608.[26] The first known decimal approximation of the (inverse) golden ratio was stated as "about " in 1597 by Michael Maestlin of the University of Tübingen in a letter to Kepler, his former student.[27] The same year, Kepler wrote to Maestlin of the Kepler triangle, which combines the golden ratio with the Pythagorean theorem. Kepler said of these:

Geometry has two great treasures: one is the theorem of Pythagoras, the other the division of a line into extreme and mean ratio. The first we may compare to a mass of gold, the second we may call a precious jewel.[28]

18th-century mathematicians Abraham de Moivre, Nicolaus I Bernoulli, and Leonhard Euler used a golden ratio-based formula which finds the value of a Fibonacci number based on its placement in the sequence; in 1843, this was rediscovered by Jacques Philippe Marie Binet, for whom it was named "Binet's formula".[29] Martin Ohm first used the German term goldener Schnitt ('golden section') to describe the ratio in 1835.[30] James Sully used the equivalent English term in 1875.[31]

By 1910, mathematician Mark Barr began using the Greek letter Phi () as a symbol for the golden ratio.[32][e] It has also been represented by tau (), the first letter of the ancient Greek τομή ('cut' or 'section').[35]

Dan Shechtman demonstrates quasicrystals at the NIST in 1985 using a Zometoy model.

The zome construction system, developed by Steve Baer in the late 1960s, is based on the symmetry system of the icosahedron/dodecahedron, and uses the golden ratio ubiquitously. Between 1973 and 1974, Roger Penrose developed Penrose tiling, a pattern related to the golden ratio both in the ratio of areas of its two rhombic tiles and in their relative frequency within the pattern.[36] This gained in interest after Dan Shechtman's Nobel-winning 1982 discovery of quasicrystals with icosahedral symmetry, which were soon afterward explained through analogies to the Penrose tiling.[37][38][39][40]

Mathematics

Irrationality

The golden ratio is an irrational number. Below are two short proofs of irrationality:

Contradiction from an expression in lowest terms

If were rational, then it would be the ratio of sides of a rectangle with integer sides (the rectangle comprising the entire diagram). But it would also be a ratio of integer sides of the smaller rectangle (the rightmost portion of the diagram) obtained by deleting a square. The sequence of decreasing integer side lengths formed by deleting squares cannot be continued indefinitely because the positive integers have a lower bound, so cannot be rational.

Recall that:

the whole is the longer part plus the shorter part;
the whole is to the longer part as the longer part is to the shorter part.

If we call the whole and the longer part then the second statement above becomes

is to as is to

To say that the golden ratio is rational means that is a fraction where and are integers. We may take to be in lowest terms and and to be positive. But if is in lowest terms, then the equally valued is in still lower terms. That is a contradiction that follows from the assumption that is rational.

By irrationality of 5

Another short proof – perhaps more commonly known – of the irrationality of the golden ratio makes use of the closure of rational numbers under addition and multiplication. If is rational, then is also rational, which is a contradiction if it is already known that the square root of all non-square natural numbers are irrational.

Minimal polynomial

The golden ratio and its negative reciprocal are the two roots of the quadratic polynomial . The golden ratio's negative and reciprocal are the two roots of the quadratic polynomial .

The golden ratio is also an algebraic number and even an algebraic integer. It has minimal polynomial

This quadratic polynomial has two roots, and

The golden ratio is also closely related to the polynomial

which has roots and As the root of a quadratic polynomial, the golden ratio is a constructible number.[41]

Golden ratio conjugate and powers

The conjugate root to the minimal polynomial is

The absolute value of this quantity () corresponds to the length ratio taken in reverse order (shorter segment length over longer segment length, ).

This illustrates the unique property of the golden ratio among positive numbers, that

or its inverse:

The conjugate and the defining quadratic polynomial relationship lead to decimal values that have their fractional part in common with :

The sequence of powers of contains these values more generally, any power of is equal to the sum of the two immediately preceding powers:

As a result, one can easily decompose any power of into a multiple of and a constant. The multiple and the constant are always adjacent Fibonacci numbers. This leads to another property of the positive powers of :

If then:

Continued fraction and square root

Approximations to the reciprocal golden ratio by finite continued fractions, or ratios of Fibonacci numbers

The formula can be expanded recursively to obtain a continued fraction for the golden ratio:[42]

It is in fact the simplest form of a continued fraction, alongside its reciprocal form:

The convergents of these continued fractions ( ... or ...) are ratios of successive Fibonacci numbers. The consistently small terms in its continued fraction explain why the approximants converge so slowly. This makes the golden ratio an extreme case of the Hurwitz inequality for Diophantine approximations, which states that for every irrational , there are infinitely many distinct fractions such that,

This means that the constant cannot be improved without excluding the golden ratio. It is, in fact, the smallest number that must be excluded to generate closer approximations of such Lagrange numbers.[43]

A continued square root form for can be obtained from , yielding:

Relationship to Fibonacci and Lucas numbers

A Fibonacci spiral (top) which approximates the golden spiral, using Fibonacci sequence square sizes up to A golden spiral is also generated (bottom) from stacking squares whose lengths of sides are numbers belonging to the sequence of Lucas numbers, here up to

Fibonacci numbers and Lucas numbers have an intricate relationship with the golden ratio. In the Fibonacci sequence, each number is equal to the sum of the preceding two, starting with the base sequence :

(OEISA000045).

The sequence of Lucas numbers (not to be confused with the generalized Lucas sequences, of which this is part) is like the Fibonacci sequence, in-which each term is the sum of the previous two, however instead starts with :

(OEISA000032).

Exceptionally, the golden ratio is equal to the limit of the ratios of successive terms in the Fibonacci sequence and sequence of Lucas numbers:[44]

In other words, if a Fibonacci and Lucas number is divided by its immediate predecessor in the sequence, the quotient approximates .

For example, and

These approximations are alternately lower and higher than and converge to as the Fibonacci and Lucas numbers increase.

Closed-form expressions for the Fibonacci and Lucas sequences that involve the golden ratio are:

Combining both formulas above, one obtains a formula for that involves both Fibonacci and Lucas numbers:

Between Fibonacci and Lucas numbers one can deduce which simplifies to express the limit of the quotient of Lucas numbers by Fibonacci numbers as equal to the square root of five:

Indeed, much stronger statements are true:

,
.

These values describe as a fundamental unit of the algebraic number field .

Successive powers of the golden ratio obey the Fibonacci recurrence, i.e.

The reduction to a linear expression can be accomplished in one step by using:

This identity allows any polynomial in to be reduced to a linear expression, as in:

Consecutive Fibonacci numbers can also be used to obtain a similar formula for the golden ratio, here by infinite summation:

In particular, the powers of themselves round to Lucas numbers (in order, except for the first two powers, and , are in reverse order):

and so forth.[45] The Lucas numbers also directly generate powers of the golden ratio; for :

Rooted in their interconnecting relationship with the golden ratio is the notion that the sum of third consecutive Fibonacci numbers equals a Lucas number, that is ; and, importantly, that .

Both the Fibonacci sequence and the sequence of Lucas numbers can be used to generate approximate forms of the golden spiral (which is a special form of a logarithmic spiral) using quarter-circles with radii from these sequences, differing only slightly from the true golden logarithmic spiral. Fibonacci spiral is generally the term used for spirals that approximate golden spirals using Fibonacci number-sequenced squares and quarter-circles.

Geometry

The golden ratio features prominently in geometry. For example, it is intrinsically involved in the internal symmetry of the pentagon, and extends to form part of the coordinates of the vertices of a regular dodecahedron, as well as those of a 5-cell. It features in the Kepler triangle and Penrose tilings too, as well as in various other polytopes.

Construction

Dividing a line segment by interior division (top) and exterior division (bottom) according to the golden ratio.

Dividing by interior division

  1. Having a line segment construct a perpendicular at point with half the length of Draw the hypotenuse
  2. Draw an arc with center and radius This arc intersects the hypotenuse at point
  3. Draw an arc with center and radius This arc intersects the original line segment at point Point divides the original line segment into line segments and with lengths in the golden ratio.

Dividing by exterior division

  1. Draw a line segment and construct off the point a segment perpendicular to and with the same length as
  2. Do bisect the line segment with
  3. A circular arc around with radius intersects in point the straight line through points and (also known as the extension of ). The ratio of to the constructed segment is the golden ratio.

Application examples you can see in the articles Pentagon with a given side length, Decagon with given circumcircle and Decagon with a given side length.

Both of the above displayed different algorithms produce geometric constructions that determine two aligned line segments where the ratio of the longer one to the shorter one is the golden ratio.

Golden angle

When two angles that make a full circle have measures in the golden ratio, the smaller is called the golden angle, with measure

This angle occurs in patterns of plant growth as the optimal spacing of leaf shoots around plant stems so that successive leaves do not block sunlight from the leaves below them.[46]

Golden spiral

A golden logarithmic spiral swirls around a golden triangle, touching its three vertices, moving inwardly inside similar fractal golden triangles.

Logarithmic spirals are self-similar spirals where distances covered per turn are in geometric progression. Importantly, isosceles golden triangles can be encased by a golden logarithmic spiral, such that successive turns of a spiral generate new golden triangles inside. This special case of logarithmic spirals is called the golden spiral, and it exhibits continuous growth in golden ratio. That is, for every turn, there is a growth factor of . As mentioned above, these golden spirals can be approximated by quarter-circles generated from Fibonacci and Lucas number-sized squares that are tiled together. In their exact form, they can be described by the polar equation with :

As with any logarithmic spiral, for with at right angles:

Its polar slope can be calculated using alongside from above,

It has a complementary angle, :

Golden spirals can be symmetrically placed inside pentagons and pentagrams as well, such that fractal copies of the underlying geometry are reproduced at all scales.

In triangles, quadrilaterals, and pentagons

Odom's construction
Odom's construction:

George Odom found a construction for involving an equilateral triangle: if an equilateral triangle is inscribed in a circle and the line segment joining the midpoints of two sides is produced to intersect the circle in either of two points, then these three points are in golden proportion.[47]

Kepler triangle
A Kepler triangle has sides shared by squares that have areas in geometric progression: .

The Kepler triangle, named after Johannes Kepler, is the unique right triangle with sides in geometric progression:

.

The Kepler triangle can also be understood as the right triangle formed by three squares whose areas are also in golden geometric progression .

Fittingly, the Pythagorean means for are precisely , , and . It is from these ratios that we are able to geometrically express the fundamental defining quadratic polynomial for with the Pythagorean theorem; that is, .

The inradius of an isosceles triangle is greatest when the triangle is composed of two mirror Kepler triangles, such that their bases lie on the same line.[48] Also, the isosceles triangle of given perimeter with the largest possible semicircle is one from two mirrored Kepler triangles.[49]

For a Kepler triangle with smallest side length , the area and acute internal angles are:

Golden triangle
Golden triangle: the double-red-arched angle is or radians.

A golden triangle is characterized as an isosceles with the property that bisecting the angle produces new acute and obtuse isosceles triangles and that are similar to the original, as well as in leg to base length ratios of and , respectively.[50]

The acute isosceles triangle is sometimes called a sublime triangle, and the ratio of its base to its equal-length sides is .[51] Its apex angle is equal to:

Both base angles of the isosceles golden triangle equal degrees each, since the sum of the angles of a triangle must equal degrees. It is the only triangle to have its three angles in ratio.[52] A regular pentagram contains five acute sublime triangles, and a regular decagon contains ten, as each two vertices connected to the center form acute golden triangles.

The obtuse isosceles triangle is sometimes called a golden gnomon, and the ratio of its base to its other sides is the reciprocal of the golden ratio, .[50] The measure of its apex angle is:

Its two base angles equal each. It is the only triangle whose internal angles are in ratio. It's base angles, being equal to , are the same measure as that of the acute golden triangle's apex angle. Five golden gnomons can be created from adjacent sides of a pentagon whose non-coincident vertices are joined by a diagonal of the pentagon.

Appropriately, the ratio of the area of the obtuse golden gnomon to that of the acute sublime triangle is in golden ratio. Bisecting a base angle inside a sublime triangle produces a golden gnomon, and another a sublime triangle. Bisecting the apex angle of a golden gnomon in ratio produces two new golden triangles, too. Golden triangles that are decomposed further like this into pairs of isosceles and obtuse golden triangles are known as Robinson triangles.[52]

Golden rectangle
To construct a golden rectangle with only a straightedge and compass in four simple steps:
Draw a square.
Draw a line from the midpoint of one side of the square to an opposite corner.
Use that line as the radius to draw an arc that defines the height of the rectangle.
Complete the golden rectangle.

The golden ratio proportions the adjacent side lengths of a golden rectangle in ratio.[53] Stacking golden rectangles produces golden rectangles anew, and removing or adding squares from golden rectangles leaves rectangles still proportioned in ratio. They can be generated by golden spirals, through successive Fibonacci and Lucas number-sized squares and quarter circles. They feature prominently in the icosahedron as well as in the dodecahedron (see section below for more detail).[54]

Golden rhombus

A golden rhombus is a rhombus whose diagonals are in proportion to the golden ratio, most commonly .[55] For a rhombus of such proportions, it's acute angle and obtuse angles are:

The lengths of its short and long diagonals and , in terms of side length are:

Its area, in terms of ,and :

Its inradius, in terms of side :

Golden rhombi feature in the rhombic triacontahedron (see section below). They also are found in the golden rhombohedron, the Bilinski dodecahedron,[56] and the rhombic hexecontahedron.[55]

Pentagon and pentagram
A pentagram colored to distinguish its line segments of different lengths. The four lengths are in golden ratio to one another.

In a regular pentagon the ratio of a diagonal to a side is the golden ratio, while intersecting diagonals section each other in the golden ratio. The golden ratio properties of a regular pentagon can be confirmed by applying Ptolemy's theorem to the quadrilateral formed by removing one of its vertices. If the quadrilateral's long edge and diagonals are and short edges are then Ptolemy's theorem gives which yields,

The diagonal segments of a pentagon form a pentagram, or five-pointed star polygon, whose geometry is quintessentially described by . Primarily, each intersection of edges sections other edges in the golden ratio. The ratio of the length of the shorter segment to the segment bounded by the two intersecting edges (that is, a side of the inverted pentagon in the pentagram's center) is as the four-color illustration shows.

A pentagram has ten isosceles triangles: five are acute sublime triangles, and five are obtuse golden gnomons. In all of them, the ratio of the longer side to the shorter side is These can be decomposed further into pairs of golden Robinson triangles, which become relevant in Penrose tilings.

Otherwise, pentagonal and pentagrammic geometry permits us to calculate the following values for :

Penrose tilings
The kite and dart tiles of the Penrose tiling. The colored arcs divide each edge in the golden ratio; when two tiles share an edge, their arcs must match.

The golden ratio appears prominently in the Penrose tiling, a family of aperiodic tilings of the plane developed by Roger Penrose, inspired by Johannes Kepler's remark that pentagrams, decagons, and other shapes could fill gaps that pentagonal shapes alone leave when tiled together.[57] Several variations of this tiling have been studied, all of whose prototiles exhibit the golden ratio:

  • Penrose's original version of this tiling used four shapes: regular pentagons and pentagrams, "boat" figures with three points of a pentagram, and "diamond" shaped rhombi.[58]
  • The kite and dart Penrose tiling uses kites with three interior angles of 72° and one interior angle of 144°, and darts, concave quadrilaterals with two interior angles of 36°, one of 72°, and one non-convex angle of 216°. Special matching rules restrict how the tiles can meet at any edge, resulting in seven combinations of tiles at any vertex. Both the kites and darts have sides of two lengths, in the golden ratio to each other. The areas of these two tile shapes are also in the golden ratio to each other.[57]
  • The rhombic Penrose tiling contains two types of rhombus, a thin rhombus with angles of 36° and 144°, and a thick rhombus with angles of 72° and 108°. Again, these rhombi can be decomposed into golden Robinson triangles. All side lengths are equal, but the ratio of the length of sides to the short diagonal in the thin rhombus equals , as does the ratio of the sides of to the long diagonal of the thick rhombus. As with the kite and dart tiling, the areas of these two tiles are in the golden ratio to each other.[57]
Original four-tile Penrose tiling
Rhombic Penrose tiling

In the dodecahedron and icosahedron

Cartesian coordinates of the dodecahedron :
(±1, ±1, ±1)
(0, ±φ, ±1/φ)
1/φ, 0, ±φ)
φ, ±1/φ, 0)
A nested cube inside the dodecahedron is represented with dotted lines.

The regular dodecahedron and its dual polyhedron the icosahedron are Platonic solids whose dimensions are related to the golden ratio. An icosahedron is made of regular pentagonal faces, whereas the icosahedron is made of equilateral triangles; both with edges.[59]

For a dodecahedron of side , the radius of a circumscribed and inscribed sphere, and midradius are (, and , respectively):

, , and .

While for an icosahedron of side , the radius of a circumscribed and inscribed sphere, and midradius are:

, , and .

The volume and surface area of the dodecahedron can be expressed in terms of :

and .

As well as for the icosahedron:

and
Three golden rectangles touch all of the vertices of a regular icosahedron.

These geometric values can be calculated from their Cartesian coordinates, which also can be given using formulas involving . The coordinates of the dodecahedron are displayed on the figure above, while those of the icosahedron are the cyclic permutations of:

, ,

Sets of three golden rectangles intersect perpendicularly inside dodecahedra and icosahedra, forming Borromean rings.[60][54] In dodecahedra, pairs of opposing vertices in golden rectangles meet the centers of pentagonal faces, and in icosahedra, they meet at its vertices. In all, the three golden rectangles contain vertices of the icosahedron, or equivalently, intersect the centers of of the dodecahedron's faces.[59]

A cube can be inscribed in a regular dodecahedron, with some of the diagonals of the pentagonal faces of the dodecahedron serving as the cube's edges; therefore, the edge lengths are in the golden ratio. The cube's volume is times that of the dodecahedron's.[61] In fact, golden rectangles inside a dodecahedron are in golden proportions to an inscribed cube, such that edges of a cube and the long edges of a golden rectangle are themselves in ratio. On the other hand, the octahedron, which is the dual polyhedron of the cube, can inscribe an icosahedron, such that an icosahedron's vertices touch the edges of an octahedron at points that divide its edges in golden ratio.[62]

Other polyhedra are related to the dodecahedron and icosahedron or their symmetries, and therefore have corresponding relations to the golden ratio. These include the compound of five cubes, compound of five octahedra, compound of five tetrahedra, the compound of ten tetrahedra, rhombic triacontahedron, icosidodecahedron, truncated icosahedron, truncated dodecahedron, and rhombicosidodecahedron, rhombic enneacontahedron, and Kepler-Poinsot polyhedra, and rhombic hexecontahedron. In four dimensions, the dodecahedron and icosahedron appear as faces of the 120-cell and 600-cell, which again have dimensions related to the golden ratio.

Other properties

The golden ratio's decimal expansion can be calculated via root-finding methods, such as Newton's method or Halley's method, on the equation or on (to compute first). The time needed to compute digits of the golden ratio using Newton's method is essentially , where is the time complexity of multiplying two -digit numbers.[63] This is considerably faster than known algorithms for and . An easily programmed alternative using only integer arithmetic is to calculate two large consecutive Fibonacci numbers and divide them. The ratio of Fibonacci numbers and each over digits, yields over significant digits of the golden ratio. The decimal expansion of the golden ratio [1] has been calculated to an accuracy of ten trillion () digits.[64]

The golden ratio and inverse golden ratio have a set of symmetries that preserve and interrelate them. They are both preserved by the fractional linear transformations – this fact corresponds to the identity and the definition quadratic equation. Further, they are interchanged by the three maps – they are reciprocals, symmetric about and (projectively) symmetric about More deeply, these maps form a subgroup of the modular group isomorphic to the symmetric group on letters, corresponding to the stabilizer of the set of standard points on the projective line, and the symmetries correspond to the quotient map – the subgroup consisting of the identity and the -cycles, in cycle notation fixes the two numbers, while the -cycles interchange these, thus realizing the map.

In the complex plane, the fifth roots of unity (for an integer ) satisfying are the vertices of a pentagon. They do not form a ring of quadratic integers, however the sum of any fifth root of unity and its complex conjugate, is a quadratic integer, an element of Specifically,

This also holds for the remaining tenth roots of unity satisfying

For the gamma function , the only solutions to the equation are and .

When the golden ratio is used as the base of a numeral system (see golden ratio base, sometimes dubbed phinary or -nary), quadratic integers in the ring – that is, numbers of the form for – have terminating representations, but rational fractions have non-terminating representations.

The golden ratio also appears in hyperbolic geometry, as the maximum distance from a point on one side of an ideal triangle to the closer of the other two sides: this distance, the side length of the equilateral triangle formed by the points of tangency of a circle inscribed within the ideal triangle, is [65]

The golden ratio appears in the theory of modular functions as well. For , let

Then

and

where and in the continued fraction should be evaluated as . The function is invariant under , a congruence subgroup of the modular group. Also for positive real numbers and then[66]

and

is a Pisot–Vijayaraghavan number.[67]

Applications and observations

Architecture

The Swiss architect Le Corbusier, famous for his contributions to the modern international style, centered his design philosophy on systems of harmony and proportion. Le Corbusier's faith in the mathematical order of the universe was closely bound to the golden ratio and the Fibonacci series, which he described as "rhythms apparent to the eye and clear in their relations with one another. And these rhythms are at the very root of human activities. They resound in man by an organic inevitability, the same fine inevitability which causes the tracing out of the Golden Section by children, old men, savages and the learned."[68][69]

Le Corbusier explicitly used the golden ratio in his Modulor system for the scale of architectural proportion. He saw this system as a continuation of the long tradition of Vitruvius, Leonardo da Vinci's "Vitruvian Man", the work of Leon Battista Alberti, and others who used the proportions of the human body to improve the appearance and function of architecture.

In addition to the golden ratio, Le Corbusier based the system on human measurements, Fibonacci numbers, and the double unit. He took suggestion of the golden ratio in human proportions to an extreme: he sectioned his model human body's height at the navel with the two sections in golden ratio, then subdivided those sections in golden ratio at the knees and throat; he used these golden ratio proportions in the Modulor system. Le Corbusier's 1927 Villa Stein in Garches exemplified the Modulor system's application. The villa's rectangular ground plan, elevation, and inner structure closely approximate golden rectangles.[70]

Another Swiss architect, Mario Botta, bases many of his designs on geometric figures. Several private houses he designed in Switzerland are composed of squares and circles, cubes and cylinders. In a house he designed in Origlio, the golden ratio is the proportion between the central section and the side sections of the house.[71]

Art

Da Vinci's illustration of a dodecahedron from Pacioli's Divina proportione (1509)

Leonardo da Vinci's illustrations of polyhedra in Pacioli's Divina proportione have led some to speculate that he incorporated the golden ratio in his paintings. But the suggestion that his Mona Lisa, for example, employs golden ratio proportions, is not supported by Leonardo's own writings.[72] Similarly, although Leonardo's Vitruvian Man is often shown in connection with the golden ratio, the proportions of the figure do not actually match it, and the text only mentions whole number ratios.[73][74]

Salvador Dalí, influenced by the works of Matila Ghyka,[75] explicitly used the golden ratio in his masterpiece, The Sacrament of the Last Supper. The dimensions of the canvas are a golden rectangle. A huge dodecahedron, in perspective so that edges appear in golden ratio to one another, is suspended above and behind Jesus and dominates the composition.[72][76]

A statistical study on 565 works of art of different great painters, performed in 1999, found that these artists had not used the golden ratio in the size of their canvases. The study concluded that the average ratio of the two sides of the paintings studied is with averages for individual artists ranging from (Goya) to (Bellini).[77] On the other hand, Pablo Tosto listed over 350 works by well-known artists, including more than 100 which have canvasses with golden rectangle and proportions, and others with proportions like and [78]

Depiction of the proportions in a medieval manuscript. According to Jan Tschichold: "Page proportion 2:3. Margin proportions 1:1:2:3. Text area proportioned in the Golden Section."[79]

Books and design

According to Jan Tschichold,

There was a time when deviations from the truly beautiful page proportions and the Golden Section were rare. Many books produced between 1550 and 1770 show these proportions exactly, to within half a millimeter.[80]

According to some sources, the golden ratio is used in everyday design, for example in the proportions of playing cards, postcards, posters, light switch plates, and widescreen televisions.[81][82][83][84]

Flags

The flag of Togo, whose aspect ratio uses the golden ratio

The aspect ratio (width to height ratio) of the flag of Togo was intended to be the golden ratio, according to its designer.[85]

Music

Ernő Lendvai analyzes Béla Bartók's works as being based on two opposing systems, that of the golden ratio and the acoustic scale,[86] though other music scholars reject that analysis.[87] French composer Erik Satie used the golden ratio in several of his pieces, including Sonneries de la Rose+Croix. The golden ratio is also apparent in the organization of the sections in the music of Debussy's Reflets dans l'eau (Reflections in Water), from Images (1st series, 1905), in which "the sequence of keys is marked out by the intervals 34, 21, 13 and 8, and the main climax sits at the phi position".[88]

The musicologist Roy Howat has observed that the formal boundaries of Debussy's La Mer correspond exactly to the golden section.[89] Trezise finds the intrinsic evidence "remarkable", but cautions that no written or reported evidence suggests that Debussy consciously sought such proportions.[90]

Music theorists including Hans Zender and Heinz Bohlen have experimented with the 833 cents scale, a musical scale based on using the golden ratio as its fundamental musical interval. When measured in cents, a logarithmic scale for musical intervals, the golden ratio is approximately 833.09 cents.[91][92][93]

Nature

Detail of the saucer plant, Aeonium tabuliforme, showing the multiple spiral arrangement (parastichy)

Johannes Kepler wrote that "the image of man and woman stems from the divine proportion. In my opinion, the propagation of plants and the progenitive acts of animals are in the same ratio".[94]

The psychologist Adolf Zeising noted that the golden ratio appeared in phyllotaxis and argued from these patterns in nature that the golden ratio was a universal law.[95][96] Zeising wrote in 1854 of a universal orthogenetic law of "striving for beauty and completeness in the realms of both nature and art".[97]

However, some have argued that many apparent manifestations of the golden ratio in nature, especially in regard to animal dimensions, are fictitious.[98]

Physics

The quasi-one-dimensional Ising ferromagnet CoNb2O6 (cobalt niobate) has 8 predicted excitation states (with E8 symmetry), that when probed with neutron scattering, showed its lowest two were in golden ratio. Specifically, these quantum phase transitions during spin excitation, which occur at near absolute zero temperature, showed pairs of kinks in its ordered-phase to spin-flips in its paramagnetic phase; revealing, just below its critical field, a spin dynamics with sharp modes at low energies approaching the golden mean.[99]

Optimization

There is no known general algorithm to arrange a given number of nodes evenly on a sphere, for any of several definitions of even distribution (see, for example, Thomson problem or Tammes problem). However, a useful approximation results from dividing the sphere into parallel bands of equal surface area and placing one node in each band at longitudes spaced by a golden section of the circle, i.e. This method was used to arrange the 1500 mirrors of the student-participatory satellite Starshine-3.[100]

The golden ratio is a critical element to golden-section search as well.

Disputed observations

Examples of disputed observations of the golden ratio include the following:

Nautilus shells are often erroneously claimed to be golden-proportioned.
  • Some specific proportions in the bodies of many animals (including humans)[101][102] and parts of the shells of mollusks[103] are often claimed to be in the golden ratio. There is a large variation in the real measures of these elements in specific individuals, however, and the proportion in question is often significantly different from the golden ratio.[101] The ratio of successive phalangeal bones of the digits and the metacarpal bone has been said to approximate the golden ratio.[102] The nautilus shell, the construction of which proceeds in a logarithmic spiral, is often cited, usually with the erroneous idea that any logarithmic spiral is related to the golden ratio,[104] but sometimes with the claim that each new chamber is golden-proportioned relative to the previous one.[105] However, measurements of nautilus shells do not support this claim.[106]
  • Historian John Man states that both the pages and text area of the Gutenberg Bible were "based on the golden section shape". However, according to his own measurements, the ratio of height to width of the pages is [107]
  • Studies by psychologists, starting with Gustav Fechner c. 1876,[108] have been devised to test the idea that the golden ratio plays a role in human perception of beauty. While Fechner found a preference for rectangle ratios centered on the golden ratio, later attempts to carefully test such a hypothesis have been, at best, inconclusive.[109][72]
  • In investing, some practitioners of technical analysis use the golden ratio to indicate support of a price level, or resistance to price increases, of a stock or commodity; after significant price changes up or down, new support and resistance levels are supposedly found at or near prices related to the starting price via the golden ratio.[110] The use of the golden ratio in investing is also related to more complicated patterns described by Fibonacci numbers (e.g. Elliott wave principle and Fibonacci retracement). However, other market analysts have published analyses suggesting that these percentages and patterns are not supported by the data.[111]

Egyptian pyramids

The Great Pyramid of Giza (also known as the Pyramid of Cheops or Khufu) has been analyzed by pyramidologists as having a doubled Kepler triangle as its cross-section. If this theory were true, the golden ratio would describe the ratio of distances from the midpoint of one of the sides of the pyramid to its apex, and from the same midpoint to the center of the pyramid's base. However, imprecision in measurement caused in part by the removal of the outer surface of the pyramid makes it impossible to distinguish this theory from other numerical theories of the proportions of the pyramid, based on pi or on whole-number ratios. The consensus of modern scholars is that this pyramid's proportions are not based on the golden ratio, because such a basis would be inconsistent both with what is known about Egyptian mathematics from the time of construction of the pyramid, and with Egyptian theories of architecture and proportion used in their other works.[112][113][114][115]

The Parthenon

Many of the proportions of the Parthenon are alleged to exhibit the golden ratio, but this has largely been discredited.[116]

The Parthenon's façade (c. 432 BC) as well as elements of its façade and elsewhere are said by some to be circumscribed by golden rectangles.[117] Other scholars deny that the Greeks had any aesthetic association with golden ratio. For example, Keith Devlin says, "Certainly, the oft repeated assertion that the Parthenon in Athens is based on the golden ratio is not supported by actual measurements. In fact, the entire story about the Greeks and golden ratio seems to be without foundation."[118] Midhat J. Gazalé affirms that "It was not until Euclid ... that the golden ratio's mathematical properties were studied."[119]

From measurements of 15 temples, 18 monumental tombs, 8 sarcophagi, and 58 grave stelae from the fifth century BC to the second century AD, one researcher concluded that the golden ratio was totally absent from Greek architecture of the classical fifth century BC, and almost absent during the following six centuries.[120] Later sources like Vitruvius (first century BC) exclusively discuss proportions that can be expressed in whole numbers, i.e. commensurate as opposed to irrational proportions.

Modern art

The Section d'Or ('Golden Section') was a collective of painters, sculptors, poets and critics associated with Cubism and Orphism.[121] Active from 1911 to around 1914, they adopted the name both to highlight that Cubism represented the continuation of a grand tradition, rather than being an isolated movement, and in homage to the mathematical harmony associated with Georges Seurat.[122] (Several authors have claimed that Seurat employed the golden ratio in his paintings, but Seurat’s writings and paintings suggest that he employed simple whole-number ratios and any approximation of the golden ratio was coincidental.)[123] The Cubists observed in its harmonies, geometric structuring of motion and form, "the primacy of idea over nature", "an absolute scientific clarity of conception".[124] However, despite this general interest in mathematical harmony, whether the paintings featured in the celebrated 1912 Salon de la Section d'Or exhibition used the golden ratio in any compositions is more difficult to determine. Livio, for example, claims that they did not,[125] and Marcel Duchamp said as much in an interview.[126] On the other hand, an analysis suggests that Juan Gris made use of the golden ratio in composing works that were likely, but not definitively, shown at the exhibition.[126][127][128] Art historian Daniel Robbins has argued that in addition to referencing the mathematical term, the exhibition's name also refers to the earlier Bandeaux d'Or group, with which Albert Gleizes and other former members of the Abbaye de Créteil had been involved.[129]

Piet Mondrian has been said to have used the golden section extensively in his geometrical paintings,[130] though other experts (including critic Yve-Alain Bois) have discredited these claims.[72][131]

See also

References

Explanatory footnotes

  1. ^ If the constraint on and each being greater than zero is lifted, then there are actually two solutions, one positive and one negative, to this equation. is defined as the positive solution. The negative solution is The sum of the two solutions is and the product of the two solutions is
  2. ^ Other names include the golden mean, golden section,[4] golden cut,[5] golden proportion, golden number,[6] medial section, and divine section.
  3. ^ Euclid, Elements, Book II, Proposition 11; Book IV, Propositions 10–11; Book VI, Proposition 30; Book XIII, Propositions 1–6, 8–11, 16–18.
  4. ^ "῎Ακρον καὶ μέσον λόγον εὐθεῖα τετμῆσθαι λέγεται, ὅταν ᾖ ὡς ἡ ὅλη πρὸς τὸ μεῖζον τμῆμα, οὕτως τὸ μεῖζον πρὸς τὸ ἔλαττὸν."[18]
  5. ^ After Classical Greek sculptor Phidias (c. 490–430 BC);[33] Barr later wrote that he thought it unlikely that Phidias actually used the golden ratio.[34]

Citations

  1. ^ a b c Sloane, N. J. A. (ed.). "Sequence A001622 (Decimal expansion of golden ratio phi (or tau) = (1 + sqrt(5))/2)". The On-Line Encyclopedia of Integer Sequences. OEIS Foundation.
  2. ^ Euclid, Elements, Book 6, Definition 3.
  3. ^ Pacioli, Luca. De divina proportione, Luca Paganinem de Paganinus de Brescia (Antonio Capella) 1509, Venice.
  4. ^ Livio 2002, pp. 3, 81.
  5. ^ Summerson John, Heavenly Mansions: And Other Essays on Architecture (New York: W.W. Norton, 1963) p. 37. "And the same applies in architecture, to the rectangles representing these and other ratios (e.g. the 'golden cut'). The sole value of these ratios is that they are intellectually fruitful and suggest the rhythms of modular design."
  6. ^ Herz-Fischler 1998.
  7. ^ Herz-Fischler 1998, pp. 20–25.
  8. ^ Strogatz, Steven (September 24, 2012). "Me, Myself, and Math: Proportion Control". The New York Times.
  9. ^ Schielack, Vincent P. (1987). "The Fibonacci Sequence and the Golden Ratio". The Mathematics Teacher. 80 (5): 357–358. ISSN 0025-5769. This source contains an elementary derivation of the golden ratio's value.{{cite journal}}: CS1 maint: postscript (link)
  10. ^ Peters, J. M. H. (1978). "An Approximate Relation between π and the Golden Ratio". The Mathematical Gazette. 62 (421): 197–198. doi:10.2307/3616690. ISSN 0025-5572.
  11. ^ Livio 2002, p. 6.
  12. ^ Livio 2002, p. 4: "... line division, which Euclid defined for ... purely geometrical purposes ..."
  13. ^ Livio 2002, pp. 7–8.
  14. ^ Livio 2002, pp. 4–5.
  15. ^ Livio 2002, p. 78.
  16. ^ Hemenway, Priya (2005). Divine Proportion: Phi In Art, Nature, and Science. New York: Sterling. pp. 20–21. ISBN 978-1-4027-3522-6.
  17. ^ Livio 2002, p. 3.
  18. ^ Fitzpatrick, Richard (2007). Euclid's Elements of Geometry. Translated by Richard Fitzpatrick. p. 156. ISBN 978-0615179841.
  19. ^ Livio 2002, pp. 88–96.
  20. ^ Mackinnon, Nick (July 1993). "The Portrait of Fra Luca Pacioli". The Mathematical Gazette. 77 (479): 130–219. doi:10.2307/3619717. JSTOR 3619717. S2CID 195006163.
  21. ^ Livio 2002, pp. 131–132.
  22. ^ Baravalle, H. V. (1948). "The geometry of the pentagon and the golden section". Mathematics Teacher. 41: 22–31. doi:10.5951/MT.41.1.0022.
  23. ^ Livio 2002, pp. 134–135.
  24. ^ Livio 2002, p. 141.
  25. ^ Schreiber, Peter (1995). "A Supplement to J. Shallit's Paper "Origins of the Analysis of the Euclidean Algorithm"". Historia Mathematica. 22 (4): 422–424. doi:10.1006/hmat.1995.1033.
  26. ^ Livio 2002, pp. 151–152.
  27. ^ O'Connor, John J.; Robertson, Edmund F. (2001). "The Golden Ratio". MacTutor History of Mathematics archive. Retrieved 2007-09-18.
  28. ^ Fink, Karl; Beman, Wooster Woodruff; Smith, David Eugene (1903). A Brief History of Mathematics: An Authorized Translation of Dr. Karl Fink's Geschichte der Elementar-Mathematik (2nd ed.). Chicago: Open Court Publishing Co. p. 223.
  29. ^ Beutelspacher, Albrecht; Petri, Bernhard (1996). "Fibonacci-Zahlen". Der Goldene Schnitt (in German). Vieweg+Teubner Verlag. pp. 87–98. doi:10.1007/978-3-322-85165-9_6.
  30. ^ Herz-Fischler 1998, pp. 167–170.
  31. ^ Posamentier & Lehmann 2011, p. 8.
  32. ^ Posamentier & Lehmann 2011, p. 285.
  33. ^ Cook, Theodore Andrea (1914). The Curves of Life. London: Constable and Company Ltd. p. 420.
  34. ^ Barr, Mark (1929). "Parameters of beauty". Architecture (NY). Vol. 60. p. 325. Reprinted: "Parameters of beauty". Think. Vol. 10–11. International Business Machines Corporation. 1944.
  35. ^ Livio 2002, p. 5.
  36. ^ Gardner, Martin (2001). "7. Penrose Tiles". The Colossal Book of Mathematics. Norton. pp. 73–93.
  37. ^ Gratias, Denis; Quiquandon, Marianne (November 2019). "Discovery of quasicrystals: The early days". Comptes Rendus Physique. 20 (7–8): 803–816. Bibcode:2019CRPhy..20..803G. doi:10.1016/j.crhy.2019.05.009. S2CID 182005594.
  38. ^ Jaric, Marko V. (2012). Introduction to the Mathematics of Quasicrystals. Elsevier. p. x. ISBN 978-0323159470. Although at the time of the discovery of quasicrystals the theory of quasiperiodic functions had been known for nearly sixty years, it was the mathematics of aperiodic Penrose tilings, mostly developed by Nicolaas de Bruijn, that provided the major influence on the new field.
  39. ^ Livio 2002, pp. 203–209.
  40. ^ Goldman, Alan I.; Anderegg, James W.; Besser, Matthew F.; Chang, Sheng-Liang; Delaney, Drew W.; Jenks, Cynthia J.; Kramer, Matthew J.; Lograsso, Thomas A.; Lynch, David W.; McCallum, R. William; Shield, Jeffrey E.; Sordelet, Daniel J.; Thiel, Patricia A. (May–June 1996). "Quasicrystalline materials". American Scientist. 84 (3): 230–241. JSTOR 29775669.
  41. ^ Martin, George E. (1998). Geometric Constructions. Undergraduate Texts in Mathematics. Springer-Verlag, New York. pp. 13–14. doi:10.1007/978-1-4612-0629-3. ISBN 0-387-98276-0. MR 1483895.
  42. ^ Hailperin, Max; Kaiser, Barbara K.; Knight, Karl W. (1998). Concrete Abstractions: An Introduction to Computer Science Using Scheme. Brooks/Cole Pub. Co. p. 63. ISBN 978-0-534-95211-2.
  43. ^ Hardy, G. H.; Wright, E. M. (1960) [1938]. "§11.8. The measure of the closest approximations to an arbitrary irrational". An Introduction to the Theory of Numbers (4th ed.). Oxford University Press. pp. 163–164. ISBN 978-0-19-853310-8.
  44. ^ Tattersall, James Joseph (2005). Elementary number theory in nine chapters (2nd ed.). Cambridge University Press. p. 28. ISBN 978-0-521-85014-8.
  45. ^ Parker, Matt (2014). "13". Things to Make and Do in the Fourth Dimension. Farrar, Straus and Giroux. p. 284. ISBN 978-0-374-53563-6.
  46. ^ King, S.; Beck, F.; Lüttge, U. (June 2004). "On the mystery of the golden angle in phyllotaxis". Plant, Cell and Environment. 27 (6): 685–695. doi:10.1111/j.1365-3040.2004.01185.x.
  47. ^ Odom, George; van de Craats, Jan (August–September 1986). "E3007: The golden ratio from an equilateral triangle and its circumcircle". Problems and solutions. The American Mathematical Monthly. 93 (7): 572. doi:10.2307/2323047. JSTOR 2323047.
  48. ^ Busard, Hubert L. L. (April–June 1968). "L'algèbre au Moyen Âge : le "Liber mensurationum" d'Abû Bekr". Journal des Savants (in French and Latin). 1968 (2): 65–124. doi:10.3406/jds.1968.1175. Archived from the original on 2022-01-12. Retrieved 2022-01-12. See problem 51, reproduced on p. 98
  49. ^ Bruce, Ian (1994). "Another instance of the golden right triangle" (PDF). Fibonacci Quarterly. 32 (3): 232–233. MR 1285752. Archived (PDF) from the original on 2022-01-29. Retrieved 2022-01-29.
  50. ^ a b Loeb, Arthur (1992). "The Golden Triangle". Concepts and Images: Visual Mathematics. Boston: Birkhäuser Boston. pp. 179–192. doi:10.1007/978-1-4612-0343-8_20. ISBN 0-8176-3620-X.
  51. ^ Fletcher, Rachel (2006). "The golden section". Nexus Network Journal. 8 (1): 67–89. doi:10.1007/s00004-006-0004-z. MR 2252091. S2CID 120991151.
  52. ^ a b Tilings Encyclopedia. 1970. Archived from the original on 2009-05-24.
  53. ^ Posamentier & Lehmann 2011, p. 11.
  54. ^ a b Burger, Edward B.; Starbird, Michael P. (2005). The Heart of Mathematics: An Invitation to Effective Thinking. Springer. p. 382. ISBN 9781931914413.
  55. ^ a b Grünbaum, Branko (1996). "A new rhombic hexecontahedron" (PDF). Geombinatorics. 6 (1): 15–18. MR 1392793.
  56. ^ Senechal, Marjorie (2006). "Donald and the golden rhombohedra". In Davis, Chandler; Ellers, Erich W. (eds.). The Coxeter Legacy. American Mathematical Society, Providence, RI. pp. 159–177. ISBN 0-8218-3722-2. MR 2209027.
  57. ^ a b c Grünbaum, Branko; Shephard, G. C. (1987). Tilings and Patterns. New York: W. H. Freeman. pp. 537–547. ISBN 978-0-7167-1193-3.
  58. ^ Penrose, Roger (1978). "Pentaplexity" (PDF). Eureka. Vol. 39. p. 32.
  59. ^ a b Livio (2002, pp. 70–72)
  60. ^ Gunn, Charles; Sullivan, John M. (2008). "The Borromean Rings: A video about the New IMU logo". In Sarhangi, Reza; Séquin, Carlo H. (eds.). Bridges Leeuwarden: Mathematics, Music, Art, Architecture, Culture. London: Tarquin Publications. pp. 63–70. ISBN 978-0-9665201-9-4.; Video at "The Borromean Rings: A new logo for the IMU", International Mathematical Union
  61. ^ Hume, Alfred (December 1900). "Some propositions on the regular dodecahedron". The American Mathematical Monthly. 7 (12): 293–295. doi:10.2307/2969130. JSTOR 2969130.
  62. ^ Coxeter, H.S.M.; du Val, Patrick; Flather, H.T.; Petrie, J.F. (1938). The Fifty-Nine Icosahedra. Vol. 6. University of Toronto Studies (Mathematical Series). p. 4. Just as a tetrahedron can be inscribed in a cube, so a cube can be inscribed in a dodecahedron. By reciprocation, this leads to an octahedron circumscribed about an icosahedron. In fact, each of the twelve vertices of the icosahedron divides an edge of the octahedron according to the "golden section.
  63. ^ Muller, J. M. (2006). Elementary functions : algorithms and implementation (2nd ed.). Boston: Birkhäuser. p. 93. ISBN 978-0817643720.
  64. ^ Yee, Alexander J. (13 March 2021). "Records Set by y-cruncher". numberword.org. Two independent computations done by Clifford Spielman.
  65. ^ Horocycles exinscrits : une propriété hyperbolique remarquable, cabri.net, retrieved 2009-07-21.
  66. ^ Brendt, B. et al. "The Rogers–Ramanujan Continued Fraction"
  67. ^ Duffin, R. J. (1978). "Algorithms for localizing roots of a polynomial and the Pisot Vijayaraghavan numbers". Pacific Journal of Mathematics. 74 (1): 47–56. doi:10.2140/pjm.1978.74.47. MR 0480414.
  68. ^ Le Corbusier, The Modulor p. 25, as cited in Padovan, Richard, Proportion: Science, Philosophy, Architecture (1999), p. 316, Taylor and Francis, ISBN 0-419-22780-6
  69. ^ Frings, Marcus, The Golden Section in Architectural Theory, Nexus Network Journal vol. 4 no. 1 (Winter 2002).
  70. ^ Le Corbusier, The Modulor, p. 35, as cited in Padovan, Richard, Proportion: Science, Philosophy, Architecture (1999), p. 320. Taylor & Francis. ISBN 0-419-22780-6: "Both the paintings and the architectural designs make use of the golden section".
  71. ^ Urwin, Simon. Analysing Architecture (2003) pp. 154–155, ISBN 0-415-30685-X
  72. ^ a b c d Livio, Mario (November 1, 2002). "The golden ratio and aesthetics". Plus Magazine. Retrieved November 26, 2018.
  73. ^ Devlin, Keith (May 2007). "The Myth That Will Not Go Away". Retrieved September 26, 2013. Part of the process of becoming a mathematics writer is, it appears, learning that you cannot refer to the golden ratio without following the first mention by a phrase that goes something like 'which the ancient Greeks and others believed to have divine and mystical properties.' Almost as compulsive is the urge to add a second factoid along the lines of 'Leonardo Da Vinci believed that the human form displays the golden ratio.' There is not a shred of evidence to back up either claim, and every reason to assume they are both false. Yet both claims, along with various others in a similar vein, live on.
  74. ^ Simanek, Donald E. "Fibonacci Flim-Flam". Archived from the original on January 9, 2010. Retrieved April 9, 2013.
  75. ^ Salvador Dalí (2008). The Dali Dimension: Decoding the Mind of a Genius (DVD). Media 3.14-TVC-FGSD-IRL-AVRO.
  76. ^ Hunt, Carla Herndon and Gilkey, Susan Nicodemus. Teaching Mathematics in the Block pp. 44, 47, ISBN 1-883001-51-X
  77. ^ Olariu, Agata, Golden Section and the Art of Painting Available online
  78. ^ Tosto, Pablo, La composición áurea en las artes plásticas – El número de oro, Librería Hachette, 1969, pp. 134–144
  79. ^ Jan Tschichold. The Form of the Book, p. 43 Fig 4. "Framework of ideal proportions in a medieval manuscript without multiple columns. Determined by Jan Tschichold 1953. Page proportion 2:3. margin proportions 1:1:2:3, Text area proportioned in the Golden Section. The lower outer corner of the text area is fixed by a diagonal as well."
  80. ^ Tschichold, Jan (1991). The Form of the Book. Hartley & Marks. pp. 27–28. ISBN 0-88179-116-4.
  81. ^ Jones, Ronald (1971). "The golden section: A most remarkable measure". The Structurist. 11: 44–52. Who would suspect, for example, that the switch plate for single light switches are standardized in terms of a Golden Rectangle?
  82. ^ Johnson, Art (1999). Famous problems and their mathematicians. Libraries Unlimited. p. 45. ISBN 978-1-56308-446-1. The Golden Ratio is a standard feature of many modern designs, from postcards and credit cards to posters and light-switch plates.
  83. ^ Stakhov, Alexey P.; Olsen, Scott (2009). The Mathematics of Harmony: From Euclid to Contemporary Mathematics and Computer Science. Singapore: World Scientific. p. 21. ISBN 978-981-277-582-5. A credit card has a form of the golden rectangle
  84. ^ Cox, Simon (2004). Cracking the Da Vinci code: the unauthorized guide to the facts behind Dan Brown's bestselling novel. Barnes & Noble Books. p. 62. ISBN 978-0-7607-5931-8. The Golden Ratio also crops up in some very unlikely places: widescreen televisions, postcards, credit cards and photographs all commonly conform to its proportions.
  85. ^ Posamentier & Lehmann 2011, chapter 4, footnote 12: "The Togo flag was designed by the artist Paul Ahyi (1930–2010), who claims to have attempted to have the flag constructed in the shape of a golden rectangle".
  86. ^ Lendvai, Ernő (1971). Béla Bartók: An Analysis of His Music. London: Kahn and Averill.
  87. ^ Livio 2002, p. 190.
  88. ^ Smith, Peter F. The Dynamics of Delight: Architecture and Aesthetics (New York: Routledge, 2003) p. 83, ISBN 0-415-30010-X
  89. ^ Howat, Roy (1983). Debussy in Proportion: A Musical Analysis. Cambridge University Press. ISBN 978-0-521-31145-8.
  90. ^ Trezise, Simon (1994). Debussy: La Mer. Cambridge University Press. p. 53. ISBN 978-0-521-44656-3.
  91. ^ Mongoven, Casey (2010). "A style of music characterized by Fibonacci and the golden ratio". Congressus Numerantium. 201: 127–138. MR 2598352.
  92. ^ Hasegawa, Robert (2011). "Gegenstrebige Harmonik in the Music of Hans Zender". Perspectives of New Music. Project Muse. 49 (1): 207–234. doi:10.1353/pnm.2011.0000. JSTOR 10.7757/persnewmusi.49.1.0207.
  93. ^ Smethurst, Reilly (2016). "Two Non-Octave Tunings by Heinz Bohlen: A Practical Proposal". In Torrence, Eve; Torrence, Bruce; Séquin, Carlo; McKenna, Douglas; Fenyvesi, Kristóf; Sarhangi, Reza (eds.). Proceedings of Bridges 2016: Mathematics, Music, Art, Architecture, Education, Culture. Phoenix, Arizona: Tessellations Publishing. pp. 519–522. ISBN 978-1-938664-19-9.
  94. ^ Livio 2002, p. 154.
  95. ^ Padovan, Richard (1999). Proportion. Taylor & Francis. pp. 305–306. ISBN 978-0-419-22780-9.
  96. ^ Padovan, Richard (2002). "Proportion: Science, Philosophy, Architecture". Nexus Network Journal. 4 (1): 113–122. doi:10.1007/s00004-001-0008-7.
  97. ^ Zeising, Adolf (1854). Neue Lehre van den Proportionen des meschlischen Körpers. preface.
  98. ^ Pommersheim, James E., Tim K. Marks, and Erica L. Flapan, eds. 2010. "Number Theory: A Lively Introduction with Proofs, Applications, and Stories". John Wiley and Sons: 82.
  99. ^ Coldea, R.; Tennant, D.A.; Wheeler, E.M.; Wawrzynksa, E.; Prabhakaran, D.; Telling, M.; Habicht, K.; Smeibidl, P.; Keifer, K. (2010). "Quantum Criticality in an Ising Chain: Experimental Evidence for Emergent E8 Symmetry". Science. 327 (5962): 177–180. arXiv:1103.3694. Bibcode:2010Sci...327..177C. doi:10.1126/science.1180085. PMID 20056884. S2CID 206522808.
  100. ^ "A Disco Ball in Space". NASA. 2001-10-09. Retrieved 2007-04-16.
  101. ^ a b Pheasant, Stephen (1998). Bodyspace. London: Taylor & Francis. ISBN 978-0-7484-0067-6.
  102. ^ a b van Laack, Walter (2001). A Better History Of Our World: Volume 1 The Universe. Aachen: van Laach GmbH.
  103. ^ Dunlap, Richard A. (1997). The Golden Ratio and Fibonacci Numbers. World Scientific. p. 130. ISBN 9789814496940.
  104. ^ Falbo, Clement (March 2005). "The golden ratio—a contrary viewpoint". The College Mathematics Journal. 36 (2): 123–134. doi:10.1080/07468342.2005.11922119. JSTOR 30044835. S2CID 14816926.
  105. ^ Moscovich, Ivan (2004). The Hinged Square & Other Puzzles. Ivan Moscovich's Masterind Collection. New York: Sterling. p. 122.
  106. ^ Peterson, Ivars (1 April 2005). "Sea shell spirals". Science News. Archived from the original on 3 October 2012. Retrieved 10 November 2008.
  107. ^ Man, John, Gutenberg: How One Man Remade the World with Word (2002) pp. 166–167, Wiley, ISBN 0-471-21823-5. "The half-folio page (30.7 × 44.5 cm) was made up of two rectangles—the whole page and its text area—based on the so called 'golden section', which specifies a crucial relationship between short and long sides, and produces an irrational number, as pi is, but is a ratio of about 5:8."
  108. ^ Fechner, Gustav (1876). Vorschule der Ästhetik. Leipzig: Breitkopf & Härtel. pp. 190–202.
  109. ^ Livio 2002, p. 7.
  110. ^ For instance, Osler writes that "38.2 percent and 61.8 percent retracements of recent rises or declines are common," in Osler, Carol (2000). "Support for Resistance: Technical Analysis and Intraday Exchange Rates" (PDF). Federal Reserve Bank of New York Economic Policy Review. 6 (2): 53–68.
  111. ^ Roy Batchelor and Richard Ramyar, "Magic numbers in the Dow," 25th International Symposium on Forecasting, 2005, p. 13, 31. "Not since the 'big is beautiful' days have giants looked better", Tom Stevenson, The Daily Telegraph, Apr. 10, 2006, and "Technical failure", The Economist, Sep. 23, 2006, are both popular-press accounts of Batchelor and Ramyar's research.
  112. ^ Herz-Fischler, Roger (2000). The Shape of the Great Pyramid. Waterloo, Ontario: Wilfrid Laurier University Press. ISBN 0-88920-324-5. MR 1788996. The entire book surveys many alternative theories for this pyramid's shape. See Chapter 11, "Kepler triangle theory", pp. 80–91, for material specific to the Kepler triangle, and p. 166 for the conclusion that the Kepler triangle theory can be eliminated by the principle that "A theory must correspond to a level of mathematics consistent with what was known to the ancient Egyptians." See note 3, p. 229, for the history of Kepler's work with this triangle.
  113. ^ Rossi, Corinna (2004). Architecture and Mathematics in Ancient Egypt. Cambridge University Press. pp. 67–68. ISBN 978-0-521-82954-0. there is no direct evidence in any ancient Egyptian written mathematical source of any arithmetic calculation or geometrical construction which could be classified as the Golden Section ... convergence to , and itself as a number, do not fit with the extant Middle Kingdom mathematical sources; see also extensive discussion of multiple alternative theories for the shape of the pyramid and other Egyptian architecture, pp. 7–56
  114. ^ Rossi, Corinna; Tout, Christopher A. (2002). "Were the Fibonacci series and the Golden Section known in ancient Egypt?". Historia Mathematica. 29 (2): 101–113. doi:10.1006/hmat.2001.2334. MR 1896969.
  115. ^ Markowsky, George (January 1992). "Misconceptions about the Golden Ratio" (PDF). The College Mathematics Journal. Mathematical Association of America. 23 (1): 2–19. doi:10.2307/2686193. JSTOR 2686193. Archived (PDF) from the original on 2020-12-11. Retrieved 2012-06-29. It does not appear that the Egyptians even knew of the existence of much less incorporated it in their buildings
  116. ^ Livio 2002, pp. 74–75.
  117. ^ Van Mersbergen, Audrey M., "Rhetorical Prototypes in Architecture: Measuring the Acropolis with a Philosophical Polemic", Communication Quarterly, Vol. 46 No. 2, 1998, pp. 194–213.
  118. ^ Devlin, Keith J. (2009) [2005]. The Math Instinct: Why You're a Mathematical Genius (Along with Lobsters, Birds, Cats, and Dogs). New York: Basic Books. p. 54. ISBN 978-1-56025-672-4.
  119. ^ Gazalé, Midhat J., Gnomon: From Pharaohs to Fractals, Princeton University Press, 1999, p. 125. ISBN 0-691-00514-1
  120. ^ Patrice Foutakis, "Did the Greeks Build According to the Golden Ratio?", Cambridge Archaeological Journal, vol. 24, n° 1, February 2014, pp. 71–86.
  121. ^ Le Salon de la Section d'Or, October 1912, Mediation Centre Pompidou
  122. ^ Jeunes Peintres ne vous frappez pas !, La Section d'Or: Numéro spécial consacré à l'Exposition de la "Section d'Or", première année, n° 1, 9 octobre 1912, pp. 1–7 Archived 2020-10-30 at the Wayback Machine, Bibliothèque Kandinsky
  123. ^ Herz-Fischler, Roger (1983). "An Examination of Claims Concerning Seurat and the Golden Number" (PDF). Gazette des Beaux-Arts. 101: 109–112.
  124. ^ Herbert, Robert, Neo-Impressionism, New York: The Solomon R. Guggenheim Foundation, 1968[page needed]
  125. ^ Livio 2002, p. 169.
  126. ^ a b Camfield, William A. (March 1965). "Juan Gris and the golden section". The Art Bulletin. 47 (1): 128–134. doi:10.1080/00043079.1965.10788819.
  127. ^ Green, Christopher (1992). Juan Gris. Yale University Press. pp. 37–38. ISBN 0300053746.
  128. ^ Cottington, David (2004). Cubism and Its Histories. Critical Perspectives in Art History. Manchester University Press. p. 112, 142. ISBN 0719050049.
  129. ^ Allard, Roger (June 1911). "Sur quelques peintres". Les Marches du Sud-Ouest: 57–64. Reprinted in Antliff, Mark; Leighten, Patricia, eds. (2008). A Cubism Reader, Documents and Criticism, 1906–1914. The University of Chicago Press. pp. 178–191.
  130. ^ Bouleau, Charles (1963). The Painter's Secret Geometry: A Study of Composition in Art. Harcourt, Brace & World. pp. 247–248. ISBN 0-87817-259-9.
  131. ^ Livio 2002, pp. 177–178.

Works cited

Further reading

External links