Normal subgroup

Subgroup invariant under conjugation
Algebraic structure → Group theory
Group theory
Basic notions
  • Subgroup
  • Normal subgroup
Group homomorphisms
  • wreath product
  • simple
  • finite
  • infinite
  • continuous
  • multiplicative
  • additive
  • cyclic
  • abelian
  • dihedral
Modular groups
  • PSL(2, Z {\displaystyle \mathbb {Z} } )
  • SL(2, Z {\displaystyle \mathbb {Z} } )
  • v
  • t
  • e

In abstract algebra, a normal subgroup (also known as an invariant subgroup or self-conjugate subgroup)[1] is a subgroup that is invariant under conjugation by members of the group of which it is a part. In other words, a subgroup N {\displaystyle N} of the group G {\displaystyle G} is normal in G {\displaystyle G} if and only if g n g 1 N {\displaystyle gng^{-1}\in N} for all g G {\displaystyle g\in G} and n N . {\displaystyle n\in N.} The usual notation for this relation is N G . {\displaystyle N\triangleleft G.}

Normal subgroups are important because they (and only they) can be used to construct quotient groups of the given group. Furthermore, the normal subgroups of G {\displaystyle G} are precisely the kernels of group homomorphisms with domain G , {\displaystyle G,} which means that they can be used to internally classify those homomorphisms.

Évariste Galois was the first to realize the importance of the existence of normal subgroups.[2]

Definitions

A subgroup N {\displaystyle N} of a group G {\displaystyle G} is called a normal subgroup of G {\displaystyle G} if it is invariant under conjugation; that is, the conjugation of an element of N {\displaystyle N} by an element of G {\displaystyle G} is always in N . {\displaystyle N.} [3] The usual notation for this relation is N G . {\displaystyle N\triangleleft G.}

Equivalent conditions

For any subgroup N {\displaystyle N} of G , {\displaystyle G,} the following conditions are equivalent to N {\displaystyle N} being a normal subgroup of G . {\displaystyle G.} Therefore, any one of them may be taken as the definition.

  • The image of conjugation of N {\displaystyle N} by any element of G {\displaystyle G} is a subset of N , {\displaystyle N,} [4] i.e., g N g 1 N {\displaystyle gNg^{-1}\subseteq N} for all g G {\displaystyle g\in G} .
  • The image of conjugation of N {\displaystyle N} by any element of G {\displaystyle G} is equal to N , {\displaystyle N,} [4] i.e., g N g 1 = N {\displaystyle gNg^{-1}=N} for all g G {\displaystyle g\in G} .
  • For all g G , {\displaystyle g\in G,} the left and right cosets g N {\displaystyle gN} and N g {\displaystyle Ng} are equal.[4]
  • The sets of left and right cosets of N {\displaystyle N} in G {\displaystyle G} coincide.[4]
  • Multiplication in G {\displaystyle G} preserves the equivalence relation "is in the same left coset as". That is, for every g , g , h , h G {\displaystyle g,g',h,h'\in G} satisfying g N = g N {\displaystyle gN=g'N} and h N = h N {\displaystyle hN=h'N} , we have ( g h ) N = ( g h ) N . {\displaystyle (gh)N=(g'h')N.}
  • There exists a group on the set of left cosets of N {\displaystyle N} where multiplication of any two left cosets g N {\displaystyle gN} and h N {\displaystyle hN} yields the left coset ( g h ) N {\displaystyle (gh)N} . (This group is called the quotient group of G {\displaystyle G} modulo N {\displaystyle N} , denoted G / N {\displaystyle G/N} .)
  • N {\displaystyle N} is a union of conjugacy classes of G . {\displaystyle G.} [2]
  • N {\displaystyle N} is preserved by the inner automorphisms of G . {\displaystyle G.} [5]
  • There is some group homomorphism G H {\displaystyle G\to H} whose kernel is N . {\displaystyle N.} [2]
  • There exists a group homomorphism ϕ : G H {\displaystyle \phi :G\to H} whose fibers form a group where the identity element is N {\displaystyle N} and multiplication of any two fibers ϕ 1 ( h 1 ) {\displaystyle \phi ^{-1}(h_{1})} and ϕ 1 ( h 2 ) {\displaystyle \phi ^{-1}(h_{2})} yields the fiber ϕ 1 ( h 1 h 2 ) {\displaystyle \phi ^{-1}(h_{1}h_{2})} . (This group is the same group G / N {\displaystyle G/N} mentioned above.)
  • There is some congruence relation on G {\displaystyle G} for which the equivalence class of the identity element is N {\displaystyle N} .
  • For all n N {\displaystyle n\in N} and g G , {\displaystyle g\in G,} the commutator [ n , g ] = n 1 g 1 n g {\displaystyle [n,g]=n^{-1}g^{-1}ng} is in N . {\displaystyle N.} [citation needed]
  • Any two elements commute modulo the normal subgroup membership relation. That is, for all g , h G , {\displaystyle g,h\in G,} g h N {\displaystyle gh\in N} if and only if h g N . {\displaystyle hg\in N.} [citation needed]

Examples

For any group G , {\displaystyle G,} the trivial subgroup { e } {\displaystyle \{e\}} consisting of just the identity element of G {\displaystyle G} is always a normal subgroup of G . {\displaystyle G.} Likewise, G {\displaystyle G} itself is always a normal subgroup of G . {\displaystyle G.} (If these are the only normal subgroups, then G {\displaystyle G} is said to be simple.)[6] Other named normal subgroups of an arbitrary group include the center of the group (the set of elements that commute with all other elements) and the commutator subgroup [ G , G ] . {\displaystyle [G,G].} [7][8] More generally, since conjugation is an isomorphism, any characteristic subgroup is a normal subgroup.[9]

If G {\displaystyle G} is an abelian group then every subgroup N {\displaystyle N} of G {\displaystyle G} is normal, because g N = { g n } n N = { n g } n N = N g . {\displaystyle gN=\{gn\}_{n\in N}=\{ng\}_{n\in N}=Ng.} More generally, for any group G {\displaystyle G} , every subgroup of the center Z ( G ) {\displaystyle Z(G)} of G {\displaystyle G} is normal in G {\displaystyle G} . (In the special case that G {\displaystyle G} is abelian, the center is all of G {\displaystyle G} , hence the fact that all subgroups of an abelian group are normal.) A group that is not abelian but for which every subgroup is normal is called a Hamiltonian group.[10]

A concrete example of a normal subgroup is the subgroup N = { ( 1 ) , ( 123 ) , ( 132 ) } {\displaystyle N=\{(1),(123),(132)\}} of the symmetric group S 3 , {\displaystyle S_{3},} consisting of the identity and both three-cycles. In particular, one can check that every coset of N {\displaystyle N} is either equal to N {\displaystyle N} itself or is equal to ( 12 ) N = { ( 12 ) , ( 23 ) , ( 13 ) } . {\displaystyle (12)N=\{(12),(23),(13)\}.} On the other hand, the subgroup H = { ( 1 ) , ( 12 ) } {\displaystyle H=\{(1),(12)\}} is not normal in S 3 {\displaystyle S_{3}} since ( 123 ) H = { ( 123 ) , ( 13 ) } { ( 123 ) , ( 23 ) } = H ( 123 ) . {\displaystyle (123)H=\{(123),(13)\}\neq \{(123),(23)\}=H(123).} [11] This illustrates the general fact that any subgroup H G {\displaystyle H\leq G} of index two is normal.

As an example of a normal subgroup within a matrix group, consider the general linear group G L n ( R ) {\displaystyle \mathrm {GL} _{n}(\mathbf {R} )} of all invertible n × n {\displaystyle n\times n} matrices with real entries under the operation of matrix multiplication and its subgroup S L n ( R ) {\displaystyle \mathrm {SL} _{n}(\mathbf {R} )} of all n × n {\displaystyle n\times n} matrices of determinant 1 (the special linear group). To see why the subgroup S L n ( R ) {\displaystyle \mathrm {SL} _{n}(\mathbf {R} )} is normal in G L n ( R ) {\displaystyle \mathrm {GL} _{n}(\mathbf {R} )} , consider any matrix X {\displaystyle X} in S L n ( R ) {\displaystyle \mathrm {SL} _{n}(\mathbf {R} )} and any invertible matrix A {\displaystyle A} . Then using the two important identities det ( A B ) = det ( A ) det ( B ) {\displaystyle \det(AB)=\det(A)\det(B)} and det ( A 1 ) = det ( A ) 1 {\displaystyle \det(A^{-1})=\det(A)^{-1}} , one has that det ( A X A 1 ) = det ( A ) det ( X ) det ( A ) 1 = det ( X ) = 1 {\displaystyle \det(AXA^{-1})=\det(A)\det(X)\det(A)^{-1}=\det(X)=1} , and so A X A 1 S L n ( R ) {\displaystyle AXA^{-1}\in \mathrm {SL} _{n}(\mathbf {R} )} as well. This means S L n ( R ) {\displaystyle \mathrm {SL} _{n}(\mathbf {R} )} is closed under conjugation in G L n ( R ) {\displaystyle \mathrm {GL} _{n}(\mathbf {R} )} , so it is a normal subgroup.[a]

In the Rubik's Cube group, the subgroups consisting of operations which only affect the orientations of either the corner pieces or the edge pieces are normal.[12]

The translation group is a normal subgroup of the Euclidean group in any dimension.[13] This means: applying a rigid transformation, followed by a translation and then the inverse rigid transformation, has the same effect as a single translation. By contrast, the subgroup of all rotations about the origin is not a normal subgroup of the Euclidean group, as long as the dimension is at least 2: first translating, then rotating about the origin, and then translating back will typically not fix the origin and will therefore not have the same effect as a single rotation about the origin.

Properties

  • If H {\displaystyle H} is a normal subgroup of G , {\displaystyle G,} and K {\displaystyle K} is a subgroup of G {\displaystyle G} containing H , {\displaystyle H,} then H {\displaystyle H} is a normal subgroup of K . {\displaystyle K.} [14]
  • A normal subgroup of a normal subgroup of a group need not be normal in the group. That is, normality is not a transitive relation. The smallest group exhibiting this phenomenon is the dihedral group of order 8.[15] However, a characteristic subgroup of a normal subgroup is normal.[16] A group in which normality is transitive is called a T-group.[17]
  • The two groups G {\displaystyle G} and H {\displaystyle H} are normal subgroups of their direct product G × H . {\displaystyle G\times H.}
  • If the group G {\displaystyle G} is a semidirect product G = N H , {\displaystyle G=N\rtimes H,} then N {\displaystyle N} is normal in G , {\displaystyle G,} though H {\displaystyle H} need not be normal in G . {\displaystyle G.}
  • If M {\displaystyle M} and N {\displaystyle N} are normal subgroups of an additive group G {\displaystyle G} such that G = M + N {\displaystyle G=M+N} and M N = { 0 } {\displaystyle M\cap N=\{0\}} , then G = M N . {\displaystyle G=M\oplus N.} [18]
  • Normality is preserved under surjective homomorphisms;[19] that is, if G H {\displaystyle G\to H} is a surjective group homomorphism and N {\displaystyle N} is normal in G , {\displaystyle G,} then the image f ( N ) {\displaystyle f(N)} is normal in H . {\displaystyle H.}
  • Normality is preserved by taking inverse images;[19] that is, if G H {\displaystyle G\to H} is a group homomorphism and N {\displaystyle N} is normal in H , {\displaystyle H,} then the inverse image f 1 ( N ) {\displaystyle f^{-1}(N)} is normal in G . {\displaystyle G.}
  • Normality is preserved on taking direct products;[20] that is, if N 1 G 1 {\displaystyle N_{1}\triangleleft G_{1}} and N 2 G 2 , {\displaystyle N_{2}\triangleleft G_{2},} then N 1 × N 2 G 1 × G 2 . {\displaystyle N_{1}\times N_{2}\;\triangleleft \;G_{1}\times G_{2}.}
  • Every subgroup of index 2 is normal. More generally, a subgroup, H , {\displaystyle H,} of finite index, n , {\displaystyle n,} in G {\displaystyle G} contains a subgroup, K , {\displaystyle K,} normal in G {\displaystyle G} and of index dividing n ! {\displaystyle n!} called the normal core. In particular, if p {\displaystyle p} is the smallest prime dividing the order of G , {\displaystyle G,} then every subgroup of index p {\displaystyle p} is normal.[21]
  • The fact that normal subgroups of G {\displaystyle G} are precisely the kernels of group homomorphisms defined on G {\displaystyle G} accounts for some of the importance of normal subgroups; they are a way to internally classify all homomorphisms defined on a group. For example, a non-identity finite group is simple if and only if it is isomorphic to all of its non-identity homomorphic images,[22] a finite group is perfect if and only if it has no normal subgroups of prime index, and a group is imperfect if and only if the derived subgroup is not supplemented by any proper normal subgroup.

Lattice of normal subgroups

Given two normal subgroups, N {\displaystyle N} and M , {\displaystyle M,} of G , {\displaystyle G,} their intersection N M {\displaystyle N\cap M} and their product N M = { n m : n N  and  m M } {\displaystyle NM=\{nm:n\in N\;{\text{ and }}\;m\in M\}} are also normal subgroups of G . {\displaystyle G.}

The normal subgroups of G {\displaystyle G} form a lattice under subset inclusion with least element, { e } , {\displaystyle \{e\},} and greatest element, G . {\displaystyle G.} The meet of two normal subgroups, N {\displaystyle N} and M , {\displaystyle M,} in this lattice is their intersection and the join is their product.

The lattice is complete and modular.[20]

Normal subgroups, quotient groups and homomorphisms

If N {\displaystyle N} is a normal subgroup, we can define a multiplication on cosets as follows:

( a 1 N ) ( a 2 N ) := ( a 1 a 2 ) N . {\displaystyle \left(a_{1}N\right)\left(a_{2}N\right):=\left(a_{1}a_{2}\right)N.}
This relation defines a mapping G / N × G / N G / N . {\displaystyle G/N\times G/N\to G/N.} To show that this mapping is well-defined, one needs to prove that the choice of representative elements a 1 , a 2 {\displaystyle a_{1},a_{2}} does not affect the result. To this end, consider some other representative elements a 1 a 1 N , a 2 a 2 N . {\displaystyle a_{1}'\in a_{1}N,a_{2}'\in a_{2}N.} Then there are n 1 , n 2 N {\displaystyle n_{1},n_{2}\in N} such that a 1 = a 1 n 1 , a 2 = a 2 n 2 . {\displaystyle a_{1}'=a_{1}n_{1},a_{2}'=a_{2}n_{2}.} It follows that
a 1 a 2 N = a 1 n 1 a 2 n 2 N = a 1 a 2 n 1 n 2 N = a 1 a 2 N , {\displaystyle a_{1}'a_{2}'N=a_{1}n_{1}a_{2}n_{2}N=a_{1}a_{2}n_{1}'n_{2}N=a_{1}a_{2}N,}
where we also used the fact that N {\displaystyle N} is a normal subgroup, and therefore there is n 1 N {\displaystyle n_{1}'\in N} such that n 1 a 2 = a 2 n 1 . {\displaystyle n_{1}a_{2}=a_{2}n_{1}'.} This proves that this product is a well-defined mapping between cosets.

With this operation, the set of cosets is itself a group, called the quotient group and denoted with G / N . {\displaystyle G/N.} There is a natural homomorphism, f : G G / N , {\displaystyle f:G\to G/N,} given by f ( a ) = a N . {\displaystyle f(a)=aN.} This homomorphism maps N {\displaystyle N} into the identity element of G / N , {\displaystyle G/N,} which is the coset e N = N , {\displaystyle eN=N,} [23] that is, ker ( f ) = N . {\displaystyle \ker(f)=N.}

In general, a group homomorphism, f : G H {\displaystyle f:G\to H} sends subgroups of G {\displaystyle G} to subgroups of H . {\displaystyle H.} Also, the preimage of any subgroup of H {\displaystyle H} is a subgroup of G . {\displaystyle G.} We call the preimage of the trivial group { e } {\displaystyle \{e\}} in H {\displaystyle H} the kernel of the homomorphism and denote it by ker f . {\displaystyle \ker f.} As it turns out, the kernel is always normal and the image of G , f ( G ) , {\displaystyle G,f(G),} is always isomorphic to G / ker f {\displaystyle G/\ker f} (the first isomorphism theorem).[24] In fact, this correspondence is a bijection between the set of all quotient groups of G , G / N , {\displaystyle G,G/N,} and the set of all homomorphic images of G {\displaystyle G} (up to isomorphism).[25] It is also easy to see that the kernel of the quotient map, f : G G / N , {\displaystyle f:G\to G/N,} is N {\displaystyle N} itself, so the normal subgroups are precisely the kernels of homomorphisms with domain G . {\displaystyle G.} [26]

See also

Notes

  1. ^ In other language: det {\displaystyle \det } is a homomorphism from G L n ( R ) {\displaystyle \mathrm {GL} _{n}(\mathbf {R} )} to the multiplicative subgroup R × {\displaystyle \mathbf {R} ^{\times }} , and S L n ( R ) {\displaystyle \mathrm {SL} _{n}(\mathbf {R} )} is the kernel. Both arguments also work over the complex numbers, or indeed over an arbitrary field.

References

  1. ^ Bradley 2010, p. 12.
  2. ^ a b c Cantrell 2000, p. 160.
  3. ^ Dummit & Foote 2004.
  4. ^ a b c d Hungerford 2003, p. 41.
  5. ^ Fraleigh 2003, p. 141.
  6. ^ Robinson 1996, p. 16.
  7. ^ Hungerford 2003, p. 45.
  8. ^ Hall 1999, p. 138.
  9. ^ Hall 1999, p. 32.
  10. ^ Hall 1999, p. 190.
  11. ^ Judson 2020, Section 10.1.
  12. ^ Bergvall et al. 2010, p. 96.
  13. ^ Thurston 1997, p. 218.
  14. ^ Hungerford 2003, p. 42.
  15. ^ Robinson 1996, p. 17.
  16. ^ Robinson 1996, p. 28.
  17. ^ Robinson 1996, p. 402.
  18. ^ Hungerford 2013, p. 290.
  19. ^ a b Hall 1999, p. 29.
  20. ^ a b Hungerford 2003, p. 46.
  21. ^ Robinson 1996, p. 36.
  22. ^ Dõmõsi & Nehaniv 2004, p. 7.
  23. ^ Hungerford 2003, pp. 42–43.
  24. ^ Hungerford 2003, p. 44.
  25. ^ Robinson 1996, p. 20.
  26. ^ Hall 1999, p. 27.

Bibliography

  • Bergvall, Olof; Hynning, Elin; Hedberg, Mikael; Mickelin, Joel; Masawe, Patrick (16 May 2010). "On Rubik's Cube" (PDF). KTH. {{cite journal}}: Cite journal requires |journal= (help)
  • Cantrell, C.D. (2000). Modern Mathematical Methods for Physicists and Engineers. Cambridge University Press. ISBN 978-0-521-59180-5.
  • Dõmõsi, Pál; Nehaniv, Chrystopher L. (2004). Algebraic Theory of Automata Networks. SIAM Monographs on Discrete Mathematics and Applications. SIAM.
  • Dummit, David S.; Foote, Richard M. (2004). Abstract Algebra (3rd ed.). John Wiley & Sons. ISBN 0-471-43334-9.
  • Fraleigh, John B. (2003). A First Course in Abstract Algebra (7th ed.). Addison-Wesley. ISBN 978-0-321-15608-2.
  • Hall, Marshall (1999). The Theory of Groups. Providence: Chelsea Publishing. ISBN 978-0-8218-1967-8.
  • Hungerford, Thomas (2003). Algebra. Graduate Texts in Mathematics. Springer.
  • Hungerford, Thomas (2013). Abstract Algebra: An Introduction. Brooks/Cole Cengage Learning.
  • Judson, Thomas W. (2020). Abstract Algebra: Theory and Applications.
  • Robinson, Derek J. S. (1996). A Course in the Theory of Groups. Graduate Texts in Mathematics. Vol. 80 (2nd ed.). Springer-Verlag. ISBN 978-1-4612-6443-9. Zbl 0836.20001.
  • Thurston, William (1997). Levy, Silvio (ed.). Three-dimensional geometry and topology, Vol. 1. Princeton Mathematical Series. Princeton University Press. ISBN 978-0-691-08304-9.
  • Bradley, C. J. (2010). The mathematical theory of symmetry in solids : representation theory for point groups and space groups. Oxford New York: Clarendon Press. ISBN 978-0-19-958258-7. OCLC 859155300.

Further reading

  • I. N. Herstein, Topics in algebra. Second edition. Xerox College Publishing, Lexington, Mass.-Toronto, Ont., 1975. xi+388 pp.

External links

  • Weisstein, Eric W. "normal subgroup". MathWorld.
  • Normal subgroup in Springer's Encyclopedia of Mathematics
  • Robert Ash: Group Fundamentals in Abstract Algebra. The Basic Graduate Year
  • Timothy Gowers, Normal subgroups and quotient groups
  • John Baez, What's a Normal Subgroup?