Spin structure

Last updated

In differential geometry, a spin structure on an orientable Riemannian manifold (M, g) allows one to define associated spinor bundles, giving rise to the notion of a spinor in differential geometry.

Contents

Spin structures have wide applications to mathematical physics, in particular to quantum field theory where they are an essential ingredient in the definition of any theory with uncharged fermions. They are also of purely mathematical interest in differential geometry, algebraic topology, and K theory. They form the foundation for spin geometry.

Overview

In geometry and in field theory, mathematicians ask whether or not a given oriented Riemannian manifold (M,g) admits spinors. One method for dealing with this problem is to require that M have a spin structure. [1] [2] [3] This is not always possible since there is potentially a topological obstruction to the existence of spin structures. Spin structures will exist if and only if the second Stiefel–Whitney class w2(M) ∈ H2(M, Z2) of M vanishes. Furthermore, if w2(M) = 0, then the set of the isomorphism classes of spin structures on M is acted upon freely and transitively by H1(M, Z2) . As the manifold M is assumed to be oriented, the first Stiefel–Whitney class w1(M) ∈ H1(M, Z2) of M vanishes too. (The Stiefel–Whitney classes wi(M) ∈ Hi(M, Z2) of a manifold M are defined to be the Stiefel–Whitney classes of its tangent bundle TM.)

The bundle of spinors πS: SM over M is then the complex vector bundle associated with the corresponding principal bundle πP: PM of spin frames over M and the spin representation of its structure group Spin(n) on the space of spinors Δn. The bundle S is called the spinor bundle for a given spin structure on M.

A precise definition of spin structure on manifold was possible only after the notion of fiber bundle had been introduced; André Haefliger (1956) found the topological obstruction to the existence of a spin structure on an orientable Riemannian manifold and Max Karoubi (1968) extended this result to the non-orientable pseudo-Riemannian case. [4] [5]

Spin structures on Riemannian manifolds

Definition

A spin structure on an orientable Riemannian manifold with an oriented vector bundle is an equivariant lift of the orthonormal frame bundle with respect to the double covering . In other words, a pair is a spin structure on the SO(n)-principal bundle when

a) is a principal Spin(n)-bundle over , and
b) is an equivariant 2-fold covering map such that

andfor all and .

Two spin structures and on the same oriented Riemannian manifold are called "equivalent" if there exists a Spin(n)-equivariant map such that

and for all and .

In this case and are two equivalent double coverings.

The definition of spin structure on as a spin structure on the principal bundle is due to André Haefliger (1956).

Obstruction

Haefliger [1] found necessary and sufficient conditions for the existence of a spin structure on an oriented Riemannian manifold (M,g). The obstruction to having a spin structure is a certain element [k] of H2(M, Z2) . For a spin structure the class [k] is the second Stiefel–Whitney class w2(M) ∈ H2(M, Z2) of M. Hence, a spin structure exists if and only if the second Stiefel–Whitney class w2(M) ∈ H2(M, Z2) of M vanishes.

Spin structures on vector bundles

Let M be a paracompact topological manifold and E an oriented vector bundle on M of dimension n equipped with a fibre metric. This means that at each point of M, the fibre of E is an inner product space. A spinor bundle of E is a prescription for consistently associating a spin representation to every point of M. There are topological obstructions to being able to do it, and consequently, a given bundle E may not admit any spinor bundle. In case it does, one says that the bundle E is spin.

This may be made rigorous through the language of principal bundles. The collection of oriented orthonormal frames of a vector bundle form a frame bundle PSO(E), which is a principal bundle under the action of the special orthogonal group SO(n). A spin structure for PSO(E) is a lift of PSO(E) to a principal bundle PSpin(E) under the action of the spin group Spin(n), by which we mean that there exists a bundle map  : PSpin(E) → PSO(E) such that

, for all pPSpin(E) and g Spin(n),

where ρ : Spin(n) → SO(n) is the mapping of groups presenting the spin group as a double-cover of SO(n).

In the special case in which E is the tangent bundle TM over the base manifold M, if a spin structure exists then one says that M is a spin manifold. Equivalently M is spin if the SO(n) principal bundle of orthonormal bases of the tangent fibers of M is a Z2 quotient of a principal spin bundle.

If the manifold has a cell decomposition or a triangulation, a spin structure can equivalently be thought of as a homotopy-class of trivialization of the tangent bundle over the 1-skeleton that extends over the 2-skeleton. If the dimension is lower than 3, one first takes a Whitney sum with a trivial line bundle.

Obstruction and classification

For an orientable vector bundle a spin structure exists on if and only if the second Stiefel–Whitney class vanishes. This is a result of Armand Borel and Friedrich Hirzebruch. [6] Furthermore, in the case is spin, the number of spin structures are in bijection with . These results can be easily proven [7] pg 110-111 using a spectral sequence argument for the associated principal -bundle . Notice this gives a fibration

hence the Serre spectral sequence can be applied. From general theory of spectral sequences, there is an exact sequence

where

In addition, and for some filtration on , hence we get a map

giving an exact sequence

Now, a spin structure is exactly a double covering of fitting into a commutative diagram

where the two left vertical maps are the double covering maps. Now, double coverings of are in bijection with index subgroups of , which is in bijection with the set of group morphisms . But, from Hurewicz theorem and change of coefficients, this is exactly the cohomology group . Applying the same argument to , the non-trivial covering corresponds to , and the map to is precisely the of the second Stiefel–Whitney class, hence . If it vanishes, then the inverse image of under the map

is the set of double coverings giving spin structures. Now, this subset of can be identified with , showing this latter cohomology group classifies the various spin structures on the vector bundle . This can be done by looking at the long exact sequence of homotopy groups of the fibration

and applying , giving the sequence of cohomology groups

Because is the kernel, and the inverse image of is in bijection with the kernel, we have the desired result.

Remarks on classification

When spin structures exist, the inequivalent spin structures on a manifold have a one-to-one correspondence (not canonical) with the elements of H1(M,Z2), which by the universal coefficient theorem is isomorphic to H1(M,Z2). More precisely, the space of the isomorphism classes of spin structures is an affine space over H1(M,Z2).

Intuitively, for each nontrivial cycle on M a spin structure corresponds to a binary choice of whether a section of the SO(N) bundle switches sheets when one encircles the loop. If w2 [8] vanishes then these choices may be extended over the two-skeleton, then (by obstruction theory) they may automatically be extended over all of M. In particle physics this corresponds to a choice of periodic or antiperiodic boundary conditions for fermions going around each loop. Note that on a complex manifold the second Stiefel-Whitney class can be computed as the first chern class .

Examples

  1. A genus g Riemann surface admits 22g inequivalent spin structures; see theta characteristic.
  2. If H2(M,Z2) vanishes, M is spin. For example, Sn is spin for all . (Note that S2 is also spin, but for different reasons; see below.)
  3. The complex projective plane CP2 is not spin.
  4. More generally, all even-dimensional complex projective spaces CP2n are not spin.
  5. All odd-dimensional complex projective spaces CP2n+1 are spin.
  6. All compact, orientable manifolds of dimension 3 or less are spin.
  7. All Calabi–Yau manifolds are spin.

Properties

SpinC structures

A spinC structure is analogous to a spin structure on an oriented Riemannian manifold, [9] but uses the SpinC group, which is defined instead by the exact sequence

To motivate this, suppose that κ : Spin(n) → U(N) is a complex spinor representation. The center of U(N) consists of the diagonal elements coming from the inclusion i : U(1) → U(N), i.e., the scalar multiples of the identity. Thus there is a homomorphism

This will always have the element (−1,−1) in the kernel. Taking the quotient modulo this element gives the group SpinC(n). This is the twisted product

where U(1) = SO(2) = S1. In other words, the group SpinC(n) is a central extension of SO(n) by S1.

Viewed another way, SpinC(n) is the quotient group obtained from Spin(n) × Spin(2) with respect to the normal Z2 which is generated by the pair of covering transformations for the bundles Spin(n) → SO(n) and Spin(2) → SO(2) respectively. This makes the SpinC group both a bundle over the circle with fibre Spin(n), and a bundle over SO(n) with fibre a circle. [10] [11]

The fundamental group π1(SpinC(n)) is isomorphic to Z if n ≠ 2, and to ZZ if n = 2.

If the manifold has a cell decomposition or a triangulation, a spinC structure can be equivalently thought of as a homotopy class of complex structure over the 2-skeleton that extends over the 3-skeleton. Similarly to the case of spin structures, one takes a Whitney sum with a trivial line bundle if the manifold is odd-dimensional.

Yet another definition is that a spinC structure on a manifold N is a complex line bundle L over N together with a spin structure on TNL.

Obstruction

A spinC structure exists when the bundle is orientable and the second Stiefel–Whitney class of the bundle E is in the image of the map H2(M, Z) → H2(M, Z/2Z) (in other words, the third integral Stiefel–Whitney class vanishes). In this case one says that E is spinC. Intuitively, the lift gives the Chern class of the square of the U(1) part of any obtained spinC bundle. By a theorem of Hopf and Hirzebruch, closed orientable 4-manifolds always admit a spinC structure.

Classification

When a manifold carries a spinC structure at all, the set of spinC structures forms an affine space. Moreover, the set of spinC structures has a free transitive action of H2(M, Z). Thus, spinC-structures correspond to elements of H2(M, Z) although not in a natural way.

Geometric picture

This has the following geometric interpretation, which is due to Edward Witten. When the spinC structure is nonzero this square root bundle has a non-integral Chern class, which means that it fails the triple overlap condition. In particular, the product of transition functions on a three-way intersection is not always equal to one, as is required for a principal bundle. Instead it is sometimes 1.

This failure occurs at precisely the same intersections as an identical failure in the triple products of transition functions of the obstructed spin bundle. Therefore, the triple products of transition functions of the full spinc bundle, which are the products of the triple product of the spin and U(1) component bundles, are either 12 = 1 or (−1)2 = 1 and so the spinC bundle satisfies the triple overlap condition and is therefore a legitimate bundle.

The details

The above intuitive geometric picture may be made concrete as follows. Consider the short exact sequence 0 → ZZZ2 → 0, where the second arrow is multiplication by 2 and the third is reduction modulo 2. This induces a long exact sequence on cohomology, which contains

where the second arrow is induced by multiplication by 2, the third is induced by restriction modulo 2 and the fourth is the associated Bockstein homomorphism β.

The obstruction to the existence of a spin bundle is an element w2 of H2(M,Z2). It reflects the fact that one may always locally lift an SO(n) bundle to a spin bundle, but one needs to choose a Z2 lift of each transition function, which is a choice of sign. The lift does not exist when the product of these three signs on a triple overlap is −1, which yields the Čech cohomology picture of w2.

To cancel this obstruction, one tensors this spin bundle with a U(1) bundle with the same obstruction w2. Notice that this is an abuse of the word bundle, as neither the spin bundle nor the U(1) bundle satisfies the triple overlap condition and so neither is actually a bundle.

A legitimate U(1) bundle is classified by its Chern class, which is an element of H2(M,Z). Identify this class with the first element in the above exact sequence. The next arrow doubles this Chern class, and so legitimate bundles will correspond to even elements in the second H2(M, Z), while odd elements will correspond to bundles that fail the triple overlap condition. The obstruction then is classified by the failure of an element in the second H2(M,Z) to be in the image of the arrow, which, by exactness, is classified by its image in H2(M,Z2) under the next arrow.

To cancel the corresponding obstruction in the spin bundle, this image needs to be w2. In particular, if w2 is not in the image of the arrow, then there does not exist any U(1) bundle with obstruction equal to w2 and so the obstruction cannot be cancelled. By exactness, w2 is in the image of the preceding arrow only if it is in the kernel of the next arrow, which we recall is the Bockstein homomorphism β. That is, the condition for the cancellation of the obstruction is

where we have used the fact that the third integral Stiefel–Whitney class W3 is the Bockstein of the second Stiefel–Whitney class w2 (this can be taken as a definition of W3).

Integral lifts of Stiefel–Whitney classes

This argument also demonstrates that second Stiefel–Whitney class defines elements not only of Z2 cohomology but also of integral cohomology in one higher degree. In fact this is the case for all even Stiefel–Whitney classes. It is traditional to use an uppercase W for the resulting classes in odd degree, which are called the integral Stiefel–Whitney classes, and are labeled by their degree (which is always odd).

Examples

  1. All oriented smooth manifolds of dimension 4 or less are spinC. [12]
  2. All almost complex manifolds are spinC.
  3. All spin manifolds are spinC.

Application to particle physics

In particle physics the spin–statistics theorem implies that the wavefunction of an uncharged fermion is a section of the associated vector bundle to the spin lift of an SO(N) bundle E. Therefore, the choice of spin structure is part of the data needed to define the wavefunction, and one often needs to sum over these choices in the partition function. In many physical theories E is the tangent bundle, but for the fermions on the worldvolumes of D-branes in string theory it is a normal bundle.

In quantum field theory charged spinors are sections of associated spinc bundles, and in particular no charged spinors can exist on a space that is not spinc. An exception arises in some supergravity theories where additional interactions imply that other fields may cancel the third Stiefel–Whitney class. The mathematical description of spinors in supergravity and string theory is a particularly subtle open problem, which was recently addressed in references. [13] [14] It turns out that the standard notion of spin structure is too restrictive for applications to supergravity and string theory, and that the correct notion of spinorial structure for the mathematical formulation of these theories is a "Lipschitz structure". [13] [15]

See also

Related Research Articles

<span class="mw-page-title-main">Special unitary group</span> Group of unitary matrices with determinant of 1

In mathematics, the special unitary group of degree n, denoted SU(n), is the Lie group of n × n unitary matrices with determinant 1.

In mathematics, complex geometry is the study of geometric structures and constructions arising out of, or described by, the complex numbers. In particular, complex geometry is concerned with the study of spaces such as complex manifolds and complex algebraic varieties, functions of several complex variables, and holomorphic constructions such as holomorphic vector bundles and coherent sheaves. Application of transcendental methods to algebraic geometry falls in this category, together with more geometric aspects of complex analysis.

In differential geometry, the Ricci curvature tensor, named after Gregorio Ricci-Curbastro, is a geometric object which is determined by a choice of Riemannian or pseudo-Riemannian metric on a manifold. It can be considered, broadly, as a measure of the degree to which the geometry of a given metric tensor differs locally from that of ordinary Euclidean space or pseudo-Euclidean space.

In Riemannian geometry, the sectional curvature is one of the ways to describe the curvature of Riemannian manifolds. The sectional curvature Kp) depends on a two-dimensional linear subspace σp of the tangent space at a point p of the manifold. It can be defined geometrically as the Gaussian curvature of the surface which has the plane σp as a tangent plane at p, obtained from geodesics which start at p in the directions of σp. The sectional curvature is a real-valued function on the 2-Grassmannian bundle over the manifold.

In mathematics, in particular in algebraic topology, differential geometry and algebraic geometry, the Chern classes are characteristic classes associated with complex vector bundles. They have since become fundamental concepts in many branches of mathematics and physics, such as string theory, Chern–Simons theory, knot theory, Gromov–Witten invariants. Chern classes were introduced by Shiing-Shen Chern.

In mathematics, a principal bundle is a mathematical object that formalizes some of the essential features of the Cartesian product of a space with a group . In the same way as with the Cartesian product, a principal bundle is equipped with

  1. An action of on , analogous to for a product space.
  2. A projection onto . For a product space, this is just the projection onto the first factor, .

In mathematics, the Pontryagin classes, named after Lev Pontryagin, are certain characteristic classes of real vector bundles. The Pontryagin classes lie in cohomology groups with degrees a multiple of four.

In mathematics and especially differential geometry, a Kähler manifold is a manifold with three mutually compatible structures: a complex structure, a Riemannian structure, and a symplectic structure. The concept was first studied by Jan Arnoldus Schouten and David van Dantzig in 1930, and then introduced by Erich Kähler in 1933. The terminology has been fixed by André Weil. Kähler geometry refers to the study of Kähler manifolds, their geometry and topology, as well as the study of structures and constructions that can be performed on Kähler manifolds, such as the existence of special connections like Hermitian Yang–Mills connections, or special metrics such as Kähler–Einstein metrics.

<span class="mw-page-title-main">Cobordism</span>

In mathematics, cobordism is a fundamental equivalence relation on the class of compact manifolds of the same dimension, set up using the concept of the boundary of a manifold. Two manifolds of the same dimension are cobordant if their disjoint union is the boundary of a compact manifold one dimension higher.

<span class="mw-page-title-main">Spin group</span> Double cover Lie group of the special orthogonal group

In mathematics the spin group, denoted Spin(n), is a Lie group whose underlying manifold is the double cover of the special orthogonal group SO(n) = SO(n, R), such that there exists a short exact sequence of Lie groups (when n ≠ 2)

<span class="mw-page-title-main">Holonomy</span> Concept in differential geometry

In differential geometry, the holonomy of a connection on a smooth manifold is a general geometrical consequence of the curvature of the connection measuring the extent to which parallel transport around closed loops fails to preserve the geometrical data being transported. For flat connections, the associated holonomy is a type of monodromy and is an inherently global notion. For curved connections, holonomy has nontrivial local and global features.

In mathematics, in particular in algebraic topology and differential geometry, the Stiefel–Whitney classes are a set of topological invariants of a real vector bundle that describe the obstructions to constructing everywhere independent sets of sections of the vector bundle. Stiefel–Whitney classes are indexed from 0 to n, where n is the rank of the vector bundle. If the Stiefel–Whitney class of index i is nonzero, then there cannot exist everywhere linearly independent sections of the vector bundle. A nonzero nth Stiefel–Whitney class indicates that every section of the bundle must vanish at some point. A nonzero first Stiefel–Whitney class indicates that the vector bundle is not orientable. For example, the first Stiefel–Whitney class of the Möbius strip, as a line bundle over the circle, is not zero, whereas the first Stiefel–Whitney class of the trivial line bundle over the circle, , is zero.

In mathematics, the Thom space,Thom complex, or Pontryagin–Thom construction of algebraic topology and differential topology is a topological space associated to a vector bundle, over any paracompact space.

<span class="mw-page-title-main">Immersion (mathematics)</span> Differentiable function whose derivative is everywhere injective

In mathematics, an immersion is a differentiable function between differentiable manifolds whose differential pushforward is everywhere injective. Explicitly, f : MN is an immersion if

In 4-dimensional topology, a branch of mathematics, Rokhlin's theorem states that if a smooth, orientable, closed 4-manifold M has a spin structure, then the signature of its intersection form, a quadratic form on the second cohomology group , is divisible by 16. The theorem is named for Vladimir Rokhlin, who proved it in 1952.

In mathematics, and especially gauge theory, Seiberg–Witten invariants are invariants of compact smooth oriented 4-manifolds introduced by Edward Witten, using the Seiberg–Witten theory studied by Nathan Seiberg and Witten during their investigations of Seiberg–Witten gauge theory.

In differential geometry, a metaplectic structure is the symplectic analog of spin structure on orientable Riemannian manifolds. A metaplectic structure on a symplectic manifold allows one to define the symplectic spinor bundle, which is the Hilbert space bundle associated to the metaplectic structure via the metaplectic representation, giving rise to the notion of a symplectic spinor field in differential geometry.

In algebraic geometry and differential geometry, the nonabelian Hodge correspondence or Corlette–Simpson correspondence is a correspondence between Higgs bundles and representations of the fundamental group of a smooth, projective complex algebraic variety, or a compact Kähler manifold.

In differential geometry, a quaternionic manifold is a quaternionic analog of a complex manifold. The definition is more complicated and technical than the one for complex manifolds due in part to the noncommutativity of the quaternions and in part to the lack of a suitable calculus of holomorphic functions for quaternions. The most succinct definition uses the language of G-structures on a manifold. Specifically, a quaternionic n-manifold can be defined as a smooth manifold of real dimension 4n equipped with a torsion-free -structure. More naïve, but straightforward, definitions lead to a dearth of examples, and exclude spaces like quaternionic projective space which should clearly be considered as quaternionic manifolds.

In mathematics, and especially differential geometry and mathematical physics, gauge theory is the general study of connections on vector bundles, principal bundles, and fibre bundles. Gauge theory in mathematics should not be confused with the closely related concept of a gauge theory in physics, which is a field theory which admits gauge symmetry. In mathematics theory means a mathematical theory, encapsulating the general study of a collection of concepts or phenomena, whereas in the physical sense a gauge theory is a mathematical model of some natural phenomenon.

References

  1. 1 2 Haefliger, A. (1956). "Sur l'extension du groupe structural d'un espace fibré". C. R. Acad. Sci. Paris. 243: 558–560.
  2. J. Milnor (1963). "Spin structures on manifolds". L'Enseignement Mathématique. 9: 198–203.
  3. Lichnerowicz, A. (1964). "Champs spinoriels et propagateurs en rélativité générale". Bull. Soc. Math. Fr. 92: 11–100. doi: 10.24033/bsmf.1604 .
  4. Karoubi, M. (1968). "Algèbres de Clifford et K-théorie". Ann. Sci. Éc. Norm. Supér. 1 (2): 161–270. doi: 10.24033/asens.1163 .
  5. Alagia, H. R.; Sánchez, C. U. (1985), "Spin structures on pseudo-Riemannian manifolds" (PDF), Revista de la Unión Matemática Argentina, 32: 64–78
  6. Borel, A.; Hirzebruch, F. (1958). "Characteristic classes and homogeneous spaces I". American Journal of Mathematics . 80 (2): 97–136. doi:10.2307/2372795. JSTOR   2372795.
  7. Pati, Vishwambhar. "Elliptic complexes and index theory" (PDF). Archived (PDF) from the original on 20 Aug 2018.
  8. "Spin manifold and the second Stiefel-Whitney class". Math.Stachexchange.
  9. Lawson, H. Blaine; Michelsohn, Marie-Louise (1989). Spin Geometry . Princeton University Press. p.  391. ISBN   978-0-691-08542-5.
  10. R. Gompf (1997). "Spinc–structures and homotopy equivalences". Geometry & Topology . 1: 41–50. arXiv: math/9705218 . Bibcode:1997math......5218G. doi:10.2140/gt.1997.1.41. S2CID   6906852.
  11. Friedrich, Thomas (2000). Dirac Operators in Riemannian Geometry . American Mathematical Society. p.  26. ISBN   978-0-8218-2055-1.
  12. Gompf, Robert E.; Stipsicz, Andras I. (1999). 4-Manifolds and Kirby Calculus . American Mathematical Society. pp.  55–58, 186–187. ISBN   0-8218-0994-6.
  13. 1 2 Lazaroiu, C.; Shahbazi, C.S. (2019). "Real pinor bundles and real Lipschitz structures". Asian Journal of Mathematics. 23 (5): 749–836. arXiv: 1606.07894 . doi:10.4310/AJM.2019.v23.n5.a3. S2CID   119598006..
  14. Lazaroiu, C.; Shahbazi, C.S. (2019). "On the spin geometry of supergravity and string theory". Geometric Methods in Physics XXXVI. Trends in Mathematics. pp. 229–235. arXiv: 1607.02103 . doi:10.1007/978-3-030-01156-7_25. ISBN   978-3-030-01155-0. S2CID   104292702.
  15. Friedrich, Thomas; Trautman, Andrzej (2000). "Spin spaces, Lipschitz groups, and spinor bundles". Annals of Global Analysis and Geometry . 18 (3): 221–240. arXiv: math/9901137 . doi:10.1023/A:1006713405277. S2CID   118698159.

Further reading