Fact-checked by Grok 2 weeks ago

Subgroups of cyclic groups

In group theory, a is generated by a single element, and its possess a rigid and well-understood structure that reflects the group's abelian nature and simplicity. Every of a is itself cyclic, generated by a power of the original . This property holds for both finite and , making them a cornerstone for studying more complex group structures. For infinite cyclic groups, such as the integers under addition \mathbb{Z} = \langle 1 \rangle, the subgroups are exactly the sets k\mathbb{Z} for each integer k \geq 0, where k=0 yields the trivial subgroup \{0\} and k=1 recovers the full group. These subgroups form a total order under inclusion: m\mathbb{Z} \subseteq n\mathbb{Z} if and only if n divides m. Each nontrivial subgroup k\mathbb{Z} is isomorphic to \mathbb{Z} itself and has index k in the parent group. In finite cyclic groups, the situation is even more discrete and countable. If G = \langle g \rangle has n, then for every positive d of n, there exists exactly one of d, generated by g^{n/d}. This has n/d and is unique up to among all subgroups of that in G. Consequently, the number of subgroups equals the number of positive of n, and the mirrors the of n, with corresponding to divisibility relations. This classification extends to applications in and , such as understanding the structure of abelian groups via the fundamental theorem, where cyclic components determine subgroup behaviors. The uniqueness of subgroups per order in finite cyclic groups also implies that cyclic groups are subgroup-rigid, distinguishing them from non-cyclic groups of the same order, which may have multiple or non-cyclic s of a given size.

General Properties

Cyclicity of Subgroups

A G is generated by a single element g \in G, denoted G = \langle g \rangle = \{ g^k \mid k \in \mathbb{Z} \}. A fundamental property of is that every is also cyclic. To see this, suppose G = \langle g \rangle and let H be a of G. Consider the set S = \{ k \in \mathbb{Z} \mid g^k \in H \}. This set S forms an in the ring \mathbb{Z}, and since \mathbb{Z} is a , S = m\mathbb{Z} for some m \in \mathbb{Z}_{\geq 0}. Thus, H = \{ g^{mk} \mid k \in \mathbb{Z} \} = \langle g^m \rangle, so H is cyclic, generated by g^m. This proof holds regardless of whether G is finite or infinite, as it relies solely on the structure of \mathbb{Z}. The cyclicity of subgroups implies that they inherit the single-generator structure of the parent group, which greatly simplifies their classification and study within cyclic groups. For example, consider the finite cyclic group \mathbb{Z}/6\mathbb{Z} = \langle 1 \rangle = \{0, 1, 2, 3, 4, 5\} under addition modulo 6. The subgroup H = \{0, 3\} equals \langle 3 \rangle, which is cyclic of order 2.

Uniqueness of Cyclic Subgroups

In a finite cyclic group G of order n, there is exactly one subgroup for each divisor d of n, and this subgroup has order d. To see this, suppose H and K are subgroups of G each of order d. Then the product HK is a subgroup of G whose order is |H||K|/|H \cap K| = d^2 / |H \cap K|, which must divide n by . Since G is cyclic, generated by some element g, both H and K are generated by powers of g; specifically, if H = \langle g^k \rangle and K = \langle g^l \rangle, the orders imply that the minimal positive exponents align uniquely via the with n, forcing H = K. Explicitly, if G = \langle g \rangle with |g| = n, the unique subgroup of order d (where d \mid n) is \langle g^{n/d} \rangle, which has order d because the order of g^{n/d} is n / \gcd(n, n/d) = d. This uniqueness implies that cyclic groups have no distinct isomorphic subgroups of the same order, in stark contrast to non-cyclic groups like the , which has three distinct subgroups of order 2. The result generalizes to infinite cyclic groups: if G is infinite cyclic, isomorphic to \mathbb{Z}, then for each positive integer m, there is a unique subgroup of index m, namely m\mathbb{Z}. This rigid structure underscores the linear ordering of subgroups in infinite cyclic groups, where each is principal and distinguished by its index.

Finite Cyclic Groups

Subgroup-Divisor Correspondence

In a finite cyclic group G of order n, there exists a between the of G and the positive of n. For each positive d of n, there is a unique H_d of G having order d. Let G = \langle g \rangle, where g is a of n. The corresponding to the d is explicitly given by H_d = \langle g^{n/d} \rangle. This has d, since (g^{n/d})^d = g^n = e and no smaller positive exponent works: the of g^{n/d} is n / \gcd(n/d, n) = n / (n/d) = d. Every of G arises in this manner, as all subgroups are cyclic (by the general of cyclic groups) and thus generated by some of g, with the uniqueness of cyclic subgroups of each ensuring the bijection. For example, consider G = \langle g \rangle of order 12. The positive divisors of 12 are 1, 2, 3, 4, 6, and 12, yielding the subgroups H_1 = \langle g^{12} \rangle = \{e\}, H_2 = \langle g^6 \rangle, H_3 = \langle g^4 \rangle, H_4 = \langle g^3 \rangle, H_6 = \langle g^2 \rangle, and H_{12} = \langle g \rangle = G. This subgroup-divisor correspondence builds on that subgroup orders divide the group order and was first established for by Gauss in , later systematized in foundational texts around the early to highlight the structure of abelian groups.

Counting and Generating Subgroups

In a finite G of order n, generated by an element g, there is a unique for each positive d of n, namely the subgroup H_d = \langle g^{n/d} \rangle of order d. Consequently, the total number of subgroups of G equals the number of positive divisors of n, denoted \tau(n) or d(n). For example, if n = p^k where p is prime and k \geq 1, then the divisors are $1, p, p^2, \dots, p^k, so \tau(p^k) = k+1 and G has exactly k+1 subgroups. Similarly, for n=6=2 \cdot 3, the divisors are $1, 2, 3, 6, yielding \tau(6)=4 subgroups: the trivial subgroup, subgroups of orders 2 and 3, and G itself. Within the subgroup H_d = \langle g^{n/d} \rangle of order d, the generators are precisely the elements h \in H_d of order d, and there are \phi(d) such elements, where \phi is . More broadly, in the full group G, the number of elements of order d is \phi(d) if d \mid n, and 0 otherwise. This distribution of generators across subgroups underscores the structured nature of cyclic groups. The explicit counting of subgroups and their generators via divisors and the totient function enables efficient enumeration, which is particularly valuable in cryptographic applications involving s, such as the Diffie-Hellman key exchange, where identifying and mitigating risks from small subgroups prevents confinement attacks.

Infinite Cyclic Groups

Characterization of Subgroups

The infinite cyclic group G is isomorphic to the additive group \mathbb{Z} of integers under addition, generated by the element 1. Every subgroup H of \mathbb{Z} is of the form m\mathbb{Z} for some nonnegative integer m \geq 0, where m is the smallest positive element in H if H is nontrivial, and m = 0 corresponds to the trivial subgroup \{0\}. This characterization arises because H, as a of \mathbb{Z}, is an additive in the \mathbb{Z}, and all ideals in \mathbb{Z} are principal, generated by the of the elements in H. To see this explicitly, if H is nontrivial, let m be the smallest positive in H; then, for any h \in H, the division algorithm yields h = mq + r with $0 \leq r < m, and since H is a , r = h - mq \in H, implying r = 0 (as m is minimal), so h \in m\mathbb{Z}; thus, H = m\mathbb{Z}. The trivial subgroup is \{0\} = 0\mathbb{Z}, which consists solely of the identity element. For example, the subgroup generated by 3 is $3\mathbb{Z} = \{\dots, -6, -3, 0, 3, 6, \dots \}, the set of all integer multiples of 3. All nontrivial subgroups of \mathbb{Z} are themselves infinite cyclic groups, each isomorphic to \mathbb{Z}.

Index and Generation

In the infinite cyclic group \mathbb{Z}, every nontrivial subgroup H takes the form H = m\mathbb{Z} for some m \geq 1. The [ \mathbb{Z} : H ] is defined as the number of distinct cosets of H in \mathbb{Z}, which equals m. Specifically, the cosets are \{0 + m\mathbb{Z}, 1 + m\mathbb{Z}, \dots, (m-1) + m\mathbb{Z}\}, partitioning \mathbb{Z} into m residue classes modulo m. For the trivial subgroup H = \{0\}, the is infinite, as there are infinitely many cosets k + \{0\} for each k \in \mathbb{Z}. The subgroup generated by an integer k \in \mathbb{Z} is \langle k \rangle = |k| \mathbb{Z}, consisting of all integer multiples of |k|. This follows from the fact that any multiple of k is a multiple of \gcd(k, \mathbb{Z}) = |k|, and conversely, multiples of |k| can be expressed using positive and negative coefficients to reach all elements in \langle k \rangle. For k = 0, \langle 0 \rangle = \{0\}, the trivial subgroup. For example, \langle 2 \rangle = 2\mathbb{Z}, the even integers, which has index 2 in \mathbb{Z} with cosets $2\mathbb{Z} (evens) and $1 + 2\mathbb{Z} (odds). Similarly, \langle 0 \rangle = \{0\} has infinite index. This index m corresponds to the order of the quotient group \mathbb{Z} / m\mathbb{Z}, which is isomorphic to the cyclic group \mathbb{Z}_m of order m. By the first isomorphism theorem, such quotients arise naturally from homomorphisms whose kernels are m\mathbb{Z}. An important implication is that any group homomorphism \phi: \mathbb{Z} \to F, where F is a finite group, must have finite kernel m\mathbb{Z} for some m, and thus factors through the finite quotient \mathbb{Z} / m\mathbb{Z} \to F. This reflects the universal property of \mathbb{Z} as a free group on one generator, where the image of 1 in F has finite order dividing |F|.

Lattice of Subgroups

Divisor Lattice in the Finite Case

In a finite G of n, the set of all , ordered by , forms a . The meet operation is the of , while the join is the subgroup generated by their , which coincides with the product since G is abelian. This is isomorphic to the of positive of n, denoted (D_n, \mid), where the partial is divisibility. Each d of n corresponds to a unique cyclic H_d of d, and H_d \leq H_e d \mid e. Under this , the of H_d \cap H_e corresponds to H_{\gcd(d,e)}, and the join \langle H_d \cup H_e \rangle corresponds to H_{\mathrm{lcm}(d,e)}. The divisor lattice (D_n, \mid) is distributive, meaning it satisfies the identities x \wedge (y \vee z) = (x \wedge y) \vee (x \wedge z) and x \vee (y \wedge z) = (x \vee y) \wedge (x \vee z) for all elements, and hence modular. The height of the , or the length of the longest from the bottom element (corresponding to the trivial ) to the top element (corresponding to G), equals \Omega(n), the total number of prime factors of n counted with multiplicity. For example, consider n = 12 = 2^2 \cdot 3, with divisors $1, 2, 3, 4, 6, 12. The consists of two maximal chains: $1 \mid 2 \mid 4 \mid 12 and $1 \mid 3 \mid 6 \mid 12, which merge at the endpoints, reflecting the prime factorization and yielding height $3. In contrast, the lattices of non-cyclic finite groups of n are generally more complex and may fail to be distributive; for instance, the lattice of the (non-cyclic of $4) is modular but not distributive.

Divisibility Lattice in the Infinite Case

The of the infinite cyclic group \mathbb{Z} under addition consists of all of the form n\mathbb{Z} for nonnegative integers n \geq 0, where $0\mathbb{Z} = \{0\} is the trivial and n\mathbb{Z} = \{nk \mid k \in \mathbb{Z}\} for n \geq 1. These are ordered by , with m\mathbb{Z} \subseteq k\mathbb{Z} k divides m. This partial order corresponds to the reverse of the usual divisibility order on the nonnegative integers: larger multiples yield smaller under . This structure forms a distributive . The meet of two subgroups m\mathbb{Z} and k\mathbb{Z} (their ) is \operatorname{lcm}(m,k)\mathbb{Z}, as the common multiples are precisely the multiples of the . The join of m\mathbb{Z} and k\mathbb{Z} (the subgroup they generate) is \operatorname{gcd}(m,k)\mathbb{Z}, since the subgroup generated by multiples of m and k consists of all integer linear combinations, which are multiples of their . For instance, consider the even integers $2\mathbb{Z} and the multiples of three $3\mathbb{Z}. Their meet is $6\mathbb{Z}, the multiples of six, while their join is \mathbb{Z} itself, as \operatorname{gcd}(2,3) = 1. This example illustrates how subgroups—neither $2\mathbb{Z} \subseteq 3\mathbb{Z} nor $3\mathbb{Z} \subseteq 2\mathbb{Z}—combine to generate the full group. The is complete, admitting arbitrary meets and joins: for any family \{n_i \mathbb{Z} \mid i \in I\}, the join is d\mathbb{Z} where d = \gcd\{n_i \mid i \in I\} (conventionally d=0 if the family includes \{0\} or has gcd 0, yielding the trivial subgroup), and the meet is e\mathbb{Z} where e = \operatorname{lcm}\{n_i \mid i \in I\} (with infinite lcm yielding \{0\} if unbounded). This completeness arises from the well-defined gcd and lcm operations extended to sets in the nonnegative integers. In contrast to the finite cyclic case, the infinite features unbounded chains, such as the descending chain \mathbb{Z} \supseteq 2\mathbb{Z} \supseteq 4\mathbb{Z} \supseteq \cdots, reflecting its non-finite height. The entire poset is isomorphic to the dual of the divisibility on the nonnegative integers, providing a precise without branches in the sense of modular branches but with elements like prime-powered subgroups.

References

  1. [1]
    [PDF] Subgroups of cyclic groups - Keith Conrad
    Theorem 2.1 tells us how to find all the subgroups of a finite cyclic group: compute the subgroup generated by each element and then just check for redundancies ...
  2. [2]
    Group Theory - Cyclic Groups
    Theorem: All subgroups of a cyclic group are cyclic. If is cyclic, then for every divisor of there exists exactly one subgroup of order which may be generated ...
  3. [3]
    [PDF] Math 403 Chapter 4: Cyclic Groups 1. Introduction
    Classification of Subgroups of Cyclic Groups: (a) Theorem (Fundamental Theorem of Cyclic Groups):. Suppose G = <g> is cyclic. (i) Every subgroup of G is cyclic.
  4. [4]
    [PDF] properties of cyclic groups
    PROPERTIES OF CYCLIC GROUPS. 1. Every subgroup of a cyclic group is cyclic. 2. Suppose G is an infinite cyclic group. Then, for every m ≥ 1, there exists a ...<|separator|>
  5. [5]
    [PDF] topics in algebra - Mathematics Area
    Page 1. Page 2. i. n. herstein. University of Chicago. TOPICS IN. ALGEBRA. 2nd ... subgroups in a large set of symmetric groups. For some mysterious reason ...
  6. [6]
    Group theory - MacTutor History of Mathematics
    By 1832 Galois had discovered that special subgroups (now called normal subgroups) are fundamental. He calls the decomposition of a group into cosets of a ...
  7. [7]
    [PDF] MAT301H1S Lec5101 Burbulla
    Theorem: if G is a cyclic group of order n and d is a positive divisor of n, then the number of elements of order d in G is φ(d). Proof: suppose d | n. Then the ...
  8. [8]
    RFC 2785 - Methods for Avoiding the "Small-Subgroup" Attacks on ...
    This document will describe those situations in which protection from "small-subgroup" type attacks is necessary when using Diffie-Hellman key agreement.Missing: cyclic | Show results with:cyclic
  9. [9]
    [PDF] §3.5 Cyclic Groups - University of South Carolina
    Proof: mZ is an infinite cyclic group and so mZ. ∼. = Z by 2nd Theorem i). In the case of infinite groups, it is possible to have a proper subgroup that is ...
  10. [10]
    [PDF] cyclic-groups.pdf
    Theorem. Subgroups of cyclic groups are cyclic. Proof. Let G = hgi be a cyclic group, where g ∈ G. Let H<G.Missing: textbook | Show results with:textbook
  11. [11]
    [PDF] Math 403 Chapter 10: Homomorphisms 1. Introduction: If quotient ...
    Example: The mapping φ : Z → Z5 given by φ(x) = x mod 5 is a homomorphism and has Kerφ = 5Z and consequently Z/5Z ≈ Z5. Intuition: In the picture earlier ...
  12. [12]
    [PDF] Counting in Finite Cyclic Groups and Finite Fields - OU Math
    It follows that the lattice of subgroups of C is just the lattice of divisors of n. In particular, for any subgroups D1 and D2 of C,. |D1 ∩ D2| = (|D1|, |D2|).Missing: history | Show results with:history
  13. [13]
    [PDF] THE LATTICE OF SUBGROUPS OF FINITE GROUPS
    Every subgroup of a cyclic group is cyclic. 4. If Hand Kare normal subgroups of the finite group G, then HK= KH is a normal subgroup. 5.