Symmetry (geometry)

Last updated
A drawing of a butterfly with bilateral symmetry, with left and right sides as mirror images of each other. Simetria-bilateria.svg
A drawing of a butterfly with bilateral symmetry, with left and right sides as mirror images of each other.

In geometry, an object has symmetry if there is an operation or transformation (such as translation, scaling, rotation or reflection) that maps the figure/object onto itself (i.e., the object has an invariance under the transform). [1] Thus, a symmetry can be thought of as an immunity to change. [2] For instance, a circle rotated about its center will have the same shape and size as the original circle, as all points before and after the transform would be indistinguishable. A circle is thus said to be symmetric under rotation or to have rotational symmetry. If the isometry is the reflection of a plane figure about a line, then the figure is said to have reflectional symmetry or line symmetry; [3] it is also possible for a figure/object to have more than one line of symmetry. [4]

Contents

The types of symmetries that are possible for a geometric object depend on the set of geometric transforms available, and on what object properties should remain unchanged after a transformation. Because the composition of two transforms is also a transform and every transform has, by definition, an inverse transform that undoes it, the set of transforms under which an object is symmetric form a mathematical group, the symmetry group of the object. [5]

Euclidean symmetries in general

The most common group of transforms applied to objects are termed the Euclidean group of "isometries", which are distance-preserving transformations in space commonly referred to as two-dimensional or three-dimensional (i.e., in plane geometry or solid geometry Euclidean spaces). These isometries consist of reflections, rotations, translations, and combinations of these basic operations. [6] Under an isometric transformation, a geometric object is said to be symmetric if, after transformation, the object is indistinguishable from the object before the transformation. [7] A geometric object is typically symmetric only under a subset or "subgroup" of all isometries. The kinds of isometry subgroups are described below, followed by other kinds of transform groups, and by the types of object invariance that are possible in geometry.

By the Cartan–Dieudonné theorem, an orthogonal transformation in n-dimensional space can be represented by the composition of at most n reflections.

Basic isometries by dimension
1D2D3D4D
ReflectionsPointAffinePointAffinePointAffinePointAffine
1ReflectionReflectionReflectionReflection
2TranslationRotationTranslationRotationTranslationRotationTranslation
3Transflection Rotoreflection TransflectionRotoreflectionTransflection
4Rotary translation Double rotation Rotary translation
5Rotary transflection

Reflectional symmetry

Reflectional symmetry, linear symmetry, mirror symmetry, mirror-image symmetry, or bilateral symmetry is symmetry with respect to reflection. [8]

In one dimension, there is a point of symmetry about which reflection takes place; in two dimensions, there is an axis of symmetry (a.k.a., line of symmetry), and in three dimensions there is a plane of symmetry. [3] [9] An object or figure for which every point has a one-to-one mapping onto another, equidistant from and on opposite sides of a common plane is called mirror symmetric (for more, see mirror image).

The axis of symmetry of a two-dimensional figure is a line such that, if a perpendicular is constructed, any two points lying on the perpendicular at equal distances from the axis of symmetry are identical. Another way to think about it is that if the shape were to be folded in half over the axis, the two halves would be identical as mirror images of each other. For example. a square has four axes of symmetry, because there are four different ways to fold it and have the edges match each other. Another example would be that of a circle, which has infinitely many axes of symmetry passing through its center for the same reason. [10]

If the letter T is reflected along a vertical axis, it appears the same. This is sometimes called vertical symmetry. Thus one can describe this phenomenon unambiguously by saying that "T has a vertical symmetry axis", or that "T has left-right symmetry".

The triangles with reflection symmetry are isosceles, the quadrilaterals with this symmetry are kites and isosceles trapezoids. [11]

For each line or plane of reflection, the symmetry group is isomorphic with Cs (see point groups in three dimensions for more), one of the three types of order two (involutions), hence algebraically isomorphic to C2. The fundamental domain is a half-plane or half-space. [12]

Point reflection and other involutive isometries

In 2 dimensions, a point reflection is a 180 degree rotation. Point Reflection.png
In 2 dimensions, a point reflection is a 180 degree rotation.

Reflection symmetry can be generalized to other isometries of m-dimensional space which are involutions, such as

(x1, ..., xm) ↦ (−x1, ..., −xk, xk+1, ..., xm)

in a certain system of Cartesian coordinates. This reflects the space along an (mk)-dimensional affine subspace. [13] If k = m, then such a transformation is known as a point reflection, or an inversion through a point. On the plane (m = 2), a point reflection is the same as a half-turn (180°) rotation; see below. Antipodal symmetry is an alternative name for a point reflection symmetry through the origin. [14]

Such a "reflection" preserves orientation if and only if k is an even number. [15] This implies that for m = 3 (as well as for other odd m), a point reflection changes the orientation of the space, like a mirror-image symmetry. That explains why in physics, the term P-symmetry (P stands for parity) is used for both point reflection and mirror symmetry. Since a point reflection in three dimensions changes a left-handed coordinate system into a right-handed coordinate system, symmetry under a point reflection is also called a left-right symmetry. [16]

Rotational symmetry

The triskelion has 3-fold rotational symmetry. The armoured triskelion on the flag of the Isle of Man.svg
The triskelion has 3-fold rotational symmetry.

Rotational symmetry is symmetry with respect to some or all rotations in m-dimensional Euclidean space. Rotations are direct isometries, which are isometries that preserve orientation. [17] Therefore, a symmetry group of rotational symmetry is a subgroup of the special Euclidean group E+(m).

Symmetry with respect to all rotations about all points implies translational symmetry with respect to all translations (because translations are compositions of rotations about distinct points), [18] and the symmetry group is the whole E+(m). This does not apply for objects because it makes space homogeneous, but it may apply for physical laws.

For symmetry with respect to rotations about a point, one can take that point as origin. These rotations form the special orthogonal group SO(m), which can be represented by the group of m × m orthogonal matrices with determinant  1. For m = 3, this is the rotation group SO(3). [19]

Phrased slightly differently, the rotation group of an object is the symmetry group within E+(m), the group of rigid motions; [20] that is, the intersection of the full symmetry group and the group of rigid motions. For chiral objects, it is the same as the full symmetry group.

Laws of physics are SO(3)-invariant if they do not distinguish different directions in space. Because of Noether's theorem, rotational symmetry of a physical system is equivalent to the angular momentum conservation law. [21] For more, see rotational invariance.

Translational symmetry

A frieze pattern with translational symmetry Frieze hop.png
A frieze pattern with translational symmetry

Translational symmetry leaves an object invariant under a discrete or continuous group of translations . [22] The illustration on the right shows four congruent footprints generated by translations along the arrow. If the line of footprints were to extend to infinity in both directions, then they would have a discrete translational symmetry; any translation that mapped one footprint onto another would leave the whole line unchanged.

Glide reflection symmetry

A frieze pattern with glide reflection symmetry Frieze step.png
A frieze pattern with glide reflection symmetry

In 2D, a glide reflection symmetry (also called a glide plane symmetry in 3D, and a transflection in general) means that a reflection in a line or plane combined with a translation along the line or in the plane, results in the same object (such as in the case of footprints). [2] [23] The composition of two glide reflections results in a translation symmetry with twice the translation vector. The symmetry group comprising glide reflections and associated translations is the frieze group p11g, and is isomorphic with the infinite cyclic group Z.

Rotoreflection symmetry

A pentagonal antiprism with marked edges shows rotoreflectional symmetry, with an order of 10. Rotoreflection example antiprism.png
A pentagonal antiprism with marked edges shows rotoreflectional symmetry, with an order of 10.

In 3D, a rotary reflection, rotoreflection or improper rotation is a rotation about an axis combined with reflection in a plane perpendicular to that axis. [24] The symmetry groups associated with rotoreflections include:

For more, see point groups in three dimensions.

Helical symmetry

In 3D geometry and higher, a screw axis (or rotary translation) is a combination of a rotation and a translation along the rotation axis. [25]

Helical symmetry is the kind of symmetry seen in everyday objects such as springs, Slinky toys, drill bits, and augers. The concept of helical symmetry can be visualized as the tracing in three-dimensional space that results from rotating an object at a constant angular speed, while simultaneously translating at a constant linear speed along its axis of rotation. At any point in time, these two motions combine to give a coiling angle that helps define the properties of the traced helix. [26] When the tracing object rotates quickly and translates slowly, the coiling angle will be close to 0°. Conversely, if the object rotates slowly and translates quickly, the coiling angle will approach 90°.

A continuous helix Helix.svg
A continuous helix

Three main classes of helical symmetry can be distinguished, based on the interplay of the angle of coiling and translation symmetries along the axis:

A regular skew-apeirogon has a discrete (3-fold here) screw-axis symmetry, drawn in perspective. Triangular helix.png
A regular skew-apeirogon has a discrete (3-fold here) screw-axis symmetry, drawn in perspective.
The Boerdijk-Coxeter helix, constructed by augmented regular tetrahedra, is an example of a screw axis symmetry that is nonperiodic. Coxeter helix 3 colors.png
The Boerdijk–Coxeter helix, constructed by augmented regular tetrahedra, is an example of a screw axis symmetry that is nonperiodic.

Double rotation symmetry

A 4D clifford torus, stereographically projected into 3D, looks like a torus. A double rotation can be seen as a helical path. Torus vectors oblique.jpg
A 4D clifford torus, stereographically projected into 3D, looks like a torus. A double rotation can be seen as a helical path.

In 4D, a double rotation symmetry can be generated as the composite of two orthogonal rotations. [29] It is similar to 3D screw axis which is the composite of a rotation and an orthogonal translation.

Non-isometric symmetries

A wider definition of geometric symmetry allows operations from a larger group than the Euclidean group of isometries. Examples of larger geometric symmetry groups are:

In Felix Klein's Erlangen program, each possible group of symmetries defines a geometry in which objects that are related by a member of the symmetry group are considered to be equivalent. [32] For example, the Euclidean group defines Euclidean geometry, whereas the group of Möbius transformations defines projective geometry.

Scale symmetry and fractals

A Julia set has scale symmetry Julia set (ice).png
A Julia set has scale symmetry

Scale symmetry means that if an object is expanded or reduced in size, the new object has the same properties as the original. [33] This self-similarity is seen in many natural structures such as cumulus clouds, lightning, ferns and coastlines, over a wide range of scales. It is generally not found in gravitationally bound structures, for example the shape of the legs of an elephant and a mouse (so-called allometric scaling). Similarly, if a soft wax candle were enlarged to the size of a tall tree, it would immediately collapse under its own weight.

A more subtle form of scale symmetry is demonstrated by fractals. As conceived by Benoît Mandelbrot, fractals are a mathematical concept in which the structure of a complex form looks similar at any degree of magnification, [34] well seen in the Mandelbrot set. A coast is an example of a naturally occurring fractal, since it retains similar-appearing complexity at every level from the view of a satellite to a microscopic examination of how the water laps up against individual grains of sand. The branching of trees, which enables small twigs to stand in for full trees in dioramas, is another example.

Because fractals can generate the appearance of patterns in nature, they have a beauty and familiarity not typically seen with mathematically generated functions. Fractals have also found a place in computer-generated movie effects, where their ability to create complex curves with fractal symmetries results in more realistic virtual worlds.

Abstract symmetry

Klein's view

With every geometry, Felix Klein associated an underlying group of symmetries. The hierarchy of geometries is thus mathematically represented as a hierarchy of these groups, and hierarchy of their invariants. For example, lengths, angles and areas are preserved with respect to the Euclidean group of symmetries, while only the incidence structure and the cross-ratio are preserved under the most general projective transformations. A concept of parallelism, which is preserved in affine geometry, is not meaningful in projective geometry. Then, by abstracting the underlying groups of symmetries from the geometries, the relationships between them can be re-established at the group level. Since the group of affine geometry is a subgroup of the group of projective geometry, any notion invariant in projective geometry is a priori meaningful in affine geometry; but not the other way round. If you add required symmetries, you have a more powerful theory but fewer concepts and theorems (which will be deeper and more general).

Thurston's view

William Thurston introduced a similar version of symmetries in geometry. A model geometry is a simply connected smooth manifold X together with a transitive action of a Lie group G on X with compact stabilizers. The Lie group can be thought of as the group of symmetries of the geometry.

A model geometry is called maximal if G is maximal among groups acting smoothly and transitively on X with compact stabilizers, i.e. if it is the maximal group of symmetries. Sometimes this condition is included in the definition of a model geometry.

A geometric structure on a manifold M is a diffeomorphism from M to X/Γ for some model geometry X, where Γ is a discrete subgroup of G acting freely on X. If a given manifold admits a geometric structure, then it admits one whose model is maximal.

A 3-dimensional model geometry X is relevant to the geometrization conjecture if it is maximal and if there is at least one compact manifold with a geometric structure modelled on X. Thurston classified the 8 model geometries satisfying these conditions; they are listed below and are sometimes called Thurston geometries. (There are also uncountably many model geometries without compact quotients.)

See also

Related Research Articles

<span class="mw-page-title-main">Cartesian coordinate system</span> Most common coordinate system (geometry)

In geometry, a Cartesian coordinate system in a plane is a coordinate system that specifies each point uniquely by a pair of real numbers called coordinates, which are the signed distances to the point from two fixed perpendicular oriented lines, called coordinate lines, coordinate axes or just axes of the system. The point where they meet is called the origin and has (0, 0) as coordinates.

<span class="mw-page-title-main">Euclidean space</span> Fundamental space of geometry

Euclidean space is the fundamental space of geometry, intended to represent physical space. Originally, in Euclid's Elements, it was the three-dimensional space of Euclidean geometry, but in modern mathematics there are Euclidean spaces of any positive integer dimension n, which are called Euclidean n-spaces when one wants to specify their dimension. For n equal to one or two, they are commonly called respectively Euclidean lines and Euclidean planes. The qualifier "Euclidean" is used to distinguish Euclidean spaces from other spaces that were later considered in physics and modern mathematics.

<span class="mw-page-title-main">Symmetry group</span> Group of transformations under which the object is invariant

In group theory, the symmetry group of a geometric object is the group of all transformations under which the object is invariant, endowed with the group operation of composition. Such a transformation is an invertible mapping of the ambient space which takes the object to itself, and which preserves all the relevant structure of the object. A frequent notation for the symmetry group of an object X is G = Sym(X).

<span class="mw-page-title-main">Affine transformation</span> Geometric transformation that preserves lines but not angles nor the origin

In Euclidean geometry, an affine transformation or affinity is a geometric transformation that preserves lines and parallelism, but not necessarily Euclidean distances and angles.

<span class="mw-page-title-main">Hyperbolic geometry</span> Non-Euclidean geometry

In mathematics, hyperbolic geometry is a non-Euclidean geometry. The parallel postulate of Euclidean geometry is replaced with:

In mathematics, the Erlangen program is a method of characterizing geometries based on group theory and projective geometry. It was published by Felix Klein in 1872 as Vergleichende Betrachtungen über neuere geometrische Forschungen. It is named after the University Erlangen-Nürnberg, where Klein worked.

<span class="mw-page-title-main">Affine geometry</span> Euclidean geometry without distance and angles

In mathematics, affine geometry is what remains of Euclidean geometry when ignoring the metric notions of distance and angle.

In geometry, an improper rotation is an isometry in Euclidean space that is a combination of a rotation about an axis and a reflection in a plane perpendicular to that axis. Reflection and inversion are each special case of improper rotation. Any improper rotation is an affine transformation and, in cases that keep the coordinate origin fixed, a linear transformation. It is used as a symmetry operation in the context of geometric symmetry, molecular symmetry and crystallography, where an object that is unchanged by a combination of rotation and reflection is said to have improper rotation symmetry.

<span class="mw-page-title-main">Reflection (mathematics)</span> Mapping from a Euclidean space to itself

In mathematics, a reflection is a mapping from a Euclidean space to itself that is an isometry with a hyperplane as a set of fixed points; this set is called the axis or plane of reflection. The image of a figure by a reflection is its mirror image in the axis or plane of reflection. For example the mirror image of the small Latin letter p for a reflection with respect to a vertical axis would look like q. Its image by reflection in a horizontal axis would look like b. A reflection is an involution: when applied twice in succession, every point returns to its original location, and every geometrical object is restored to its original state.

<span class="mw-page-title-main">Rotation (mathematics)</span> Motion of a certain space that preserves at least one point

Rotation in mathematics is a concept originating in geometry. Any rotation is a motion of a certain space that preserves at least one point. It can describe, for example, the motion of a rigid body around a fixed point. Rotation can have a sign (as in the sign of an angle): a clockwise rotation is a negative magnitude so a counterclockwise turn has a positive magnitude. A rotation is different from other types of motions: translations, which have no fixed points, and (hyperplane) reflections, each of them having an entire (n − 1)-dimensional flat of fixed points in a n-dimensional space.

<span class="mw-page-title-main">Glide reflection</span> Geometric transformation combining reflection and translation

In geometry, a glide reflection or transflection is a geometric transformation that consists of a reflection across a hyperplane and a translation ("glide") in a direction parallel to that hyperplane, combined into a single transformation. Because the distances between points are not changed under glide reflection, it is a motion or isometry. When the context is the two-dimensional Euclidean plane, the hyperplane of reflection is a straight line called the glide line or glide axis. When the context is three-dimensional space, the hyperplane of reflection is a plane called the glide plane. The displacement vector of the translation is called the glide vector.

<span class="mw-page-title-main">Space group</span> Symmetry group of a configuration in space

In mathematics, physics and chemistry, a space group is the symmetry group of a repeating pattern in space, usually in three dimensions. The elements of a space group are the rigid transformations of the pattern that leave it unchanged. In three dimensions, space groups are classified into 219 distinct types, or 230 types if chiral copies are considered distinct. Space groups are discrete cocompact groups of isometries of an oriented Euclidean space in any number of dimensions. In dimensions other than 3, they are sometimes called Bieberbach groups.

<span class="mw-page-title-main">Transformation geometry</span> Branch of mathematics concerned with movement of shapes and sets

In mathematics, transformation geometry is the name of a mathematical and pedagogic take on the study of geometry by focusing on groups of geometric transformations, and properties that are invariant under them. It is opposed to the classical synthetic geometry approach of Euclidean geometry, that focuses on proving theorems.

<span class="mw-page-title-main">Euclidean group</span> Isometry group of Euclidean space

In mathematics, a Euclidean group is the group of (Euclidean) isometries of a Euclidean space ; that is, the transformations of that space that preserve the Euclidean distance between any two points (also called Euclidean transformations). The group depends only on the dimension n of the space, and is commonly denoted E(n) or ISO(n).

<span class="mw-page-title-main">Rotational symmetry</span> Property of objects which appear unchanged after a partial rotation

Rotational symmetry, also known as radial symmetry in geometry, is the property a shape has when it looks the same after some rotation by a partial turn. An object's degree of rotational symmetry is the number of distinct orientations in which it looks exactly the same for each rotation.

<span class="mw-page-title-main">Point group</span> Group of geometric symmetries with at least one fixed point

In geometry, a point group is a mathematical group of symmetry operations (isometries in a Euclidean space) that have a fixed point in common. The coordinate origin of the Euclidean space is conventionally taken to be a fixed point, and every point group in dimension d is then a subgroup of the orthogonal group O(d). Point groups are used to describe the symmetries of geometric figures and physical objects such as molecules.

<span class="mw-page-title-main">Chirality (mathematics)</span> Property of an object that is not congruent to its mirror image

In geometry, a figure is chiral if it is not identical to its mirror image, or, more precisely, if it cannot be mapped to its mirror image by rotations and translations alone. An object that is not chiral is said to be achiral.

In a group, the conjugate by g of h is ghg−1.

Geometry is a branch of mathematics concerned with questions of shape, size, relative position of figures, and the properties of space. Geometry is one of the oldest mathematical sciences.

<span class="mw-page-title-main">Point reflection</span> Geometric symmetry operation

In geometry, a point reflection is a transformation of affine space in which every point is reflected across a specific fixed point. When dealing with crystal structures and in the physical sciences the terms inversion symmetry, inversion center or centrosymmetric are more commonly used.

References

  1. Martin, G. (1996). Transformation Geometry: An Introduction to Symmetry. Springer. p. 28.
  2. 1 2 "Symmetry | Thinking about Geometry | Underground Mathematics". undergroundmathematics.org. Retrieved 2019-12-06.
  3. 1 2 "Symmetry - MathBitsNotebook(Geo - CCSS Math)". mathbitsnotebook.com. Retrieved 2019-12-06.
  4. Freitag, Mark (2013). Mathematics for Elementary School Teachers: A Process Approach. Cengage Learning. p. 721.
  5. Miller, Willard Jr. (1972). Symmetry Groups and Their Applications. New York: Academic Press. OCLC   589081. Archived from the original on 2010-02-17. Retrieved 2009-09-28.
  6. "Higher Dimensional Group Theory". Archived from the original on 2012-07-23. Retrieved 2013-04-16.
  7. "2.6 Reflection Symmetry". CK-12 Foundation. Retrieved 2019-12-06.
  8. Weyl, Hermann (1982) [1952]. Symmetry. Princeton: Princeton University Press. ISBN   0-691-02374-3.
  9. Cowin, Stephen C.; Doty, Stephen B. (2007). Tissue Mechanics . Springer. p.  152. ISBN   9780387368252.
  10. Caldecott, Stratford (2009). Beauty for Truth's Sake: On the Re-enchantment of Education. Brazos Press. p. 70.
  11. Bassarear, Tom (2011). Mathematics for Elementary School Teachers (5 ed.). Cengage Learning. p. 499.
  12. Johnson, N. W. Johnson (2018). "11: Finite symmetry groups". Geometries and Transformations. Cambridge University Press.
  13. Hertrich-Jeromin, Udo (2003). Introduction to Möbius Differential Geometry. Cambridge University Press.
  14. Dieck, Tammo (2008). Algebraic Topology . European Mathematical Society. pp.  261. ISBN   9783037190487.
  15. William H. Barker, Roger Howe Continuous Symmetry: From Euclid to Klein (Google eBook) American Mathematical Soc
  16. W.M. Gibson & B.R. Pollard (1980). Symmetry principles in elementary particle physics. Cambridge University Press. pp. 120–122. ISBN   0-521-29964-0.
  17. Vladimir G. Ivancevic, Tijana T. Ivancevic (2005) Natural Biodynamics World Scientific
  18. Singer, David A. (1998). Geometry: Plane and Fancy . Springer Science & Business Media.
  19. Joshi, A. W. (2007). Elements of Group Theory for Physicists. New Age International. pp. 111ff.
  20. Hartshorne, Robin (2000). Geometry: Euclid and Beyond. Springer Science & Business Media.
  21. Kosmann-Schwarzbach, Yvette (2010). The Noether theorems: Invariance and conservation laws in the twentieth century. Sources and Studies in the History of Mathematics and Physical Sciences. Springer-Verlag. ISBN   978-0-387-87867-6.
  22. Stenger, Victor J. (2000) and Mahou Shiro (2007). Timeless Reality. Prometheus Books. Especially chapter 12. Nontechnical.
  23. Martin, George E. (1982), Transformation Geometry: An Introduction to Symmetry, Undergraduate Texts in Mathematics, Springer, p. 64, ISBN   9780387906362 .
  24. Robert O. Gould, Steffen Borchardt-Ott (2011)Crystallography: An Introduction Springer Science & Business Media
  25. Bottema, O, and B. Roth, Theoretical Kinematics, Dover Publications (September 1990)
  26. George R. McGhee (2006) The Geometry of Evolution: Adaptive Landscapes and Theoretical Morphospaces Cambridge University Press p.64
  27. Anna Ursyn(2012) Biologically-inspired Computing for the Arts: Scientific Data Through Graphics IGI Global Snippet p.209 [ clarification needed ]
  28. Sinden, Richard R. (1994). DNA structure and function. Gulf Professional Publishing. p. 101. ISBN   9780126457506.
  29. Charles Howard Hinton (1906) The Fourth Dimension (Google eBook) S. Sonnenschein & Company p.223
  30. H.S.M. Coxeter (1961,9) Introduction to Geometry, §5 Similarity in the Euclidean Plane, pp. 67–76, §7 Isometry and Similarity in Euclidean Space, pp 96–104, John Wiley & Sons.
  31. William Thurston. Three-dimensional geometry and topology. Vol. 1. Edited by Silvio Levy. Princeton Mathematical Series, 35. Princeton University Press, Princeton, NJ, 1997. x+311 pp. ISBN   0-691-08304-5
  32. Klein, Felix, 1872. "Vergleichende Betrachtungen über neuere geometrische Forschungen" ('A comparative review of recent researches in geometry'), Mathematische Annalen, 43 (1893) pp. 63–100 (Also: Gesammelte Abh. Vol. 1, Springer, 1921, pp. 460–497).
    An English translation by Mellen Haskell appeared in Bull. N. Y. Math. Soc 2 (1892–1893): 215–249.
  33. Tian Yu Cao Conceptual Foundations of Quantum Field Theory Cambridge University Press p.154-155
  34. Gouyet, Jean-François (1996). Physics and fractal structures. Paris/New York: Masson Springer. ISBN   978-0-387-94153-0.