Skip to main content

Algebra Miscs

· 13 min read
Xinyu Ma
Research Scientist @ Meta

Misc things that are taught in class but not written in the note of MATH 210.

Functors

Functor as a Colimit

For any small category C\mathcal{C} and any functor F:CopSetsF: \mathcal{C}^{op}\to Sets, FF can be written as a colimit of representable functors X^=MorC(,x)\hat{X} = \mathrm{Mor}_{\mathcal{C}}(-,x).

Proof

The trick here is to pick a diagram I\mathcal{I}.

TA said "stop thinking and just write for category problems", so let's just write down what we can do. By Yoneda lemma, there exists a natural transformation α:X^F\alpha: \hat{X}\Rightarrow F if and only if F(X)F(X) is not empty. Therefore, foreach nonempty F(X)F(X) we can take an arbitrary uXF(X)u_X\in F(X) To satisfy the commutativity requirements we need to have uX=Ff(uY)u_X = Ff(u_Y) for f:XYf: X\to Y. There is no guarentee that we can pick such elements for every F(X)F(X), so let's just focus on the part we can pick. Define a natural transformation λX:X^F\lambda^X: \hat{X}\Rightarrow F as αuX\alpha^{u_X} in the Yoneda lemma:

λYX:X^(Y)F(Y)fFf(uX)\begin{darray}{rrcl} \lambda^X_Y:& \hat{X}(Y)&\to& F(Y) \\ & f &\mapsto& Ff(u_X) \end{darray}

FF is a cocone. Because for all g:XYg: X\to Y, λYg=λX\lambda^Y \circ g_* = \lambda^X as:

λZYg(f)=λZY(gf)=FfFg(uY)=Ff(uX)=λZX(f).\lambda^Y_Z\circ g_*(f) = \lambda^Y_Z(g\circ f) = Ff\circ Fg(u_Y) = Ff(u_X) = \lambda_Z^X(f).

Clearly, FF is not a colimit:

  • We did not handle every F(X)F(X).
  • We only picked one special element from each F(X)F(X), which cannot represent the whole property of FF.

So let's fix this. What we do is picking all possible uXu_X, instead of one element. (i.e. Comma Category) Let J\mathcal{J} be the category of all pairs (Y,x)(Y, x) s.t. xF(Y)x\in F(Y), with morphisms (Y,x)(Y,x)(Y, x)\to (Y',x') for f:YYf: Y'\to Y in C\mathcal{C} and Ff(x)=xFf(x) = x' in Set\mathrm{Set}. Note that multiple objects in the diagram can be mapped to one in the category. Therefore, we define this map M:JFun(Cop,Set)M: \mathcal{J}\to\mathrm{Fun}(\mathcal{C}^{op}, \mathrm{Set}) as M(Y,x)=Y^M(Y, x) = \hat{Y}. Then, there are F(X)\|F(X)\| many pairs mapped to the same X^\hat{X}. We need to specify a λX:X^F\lambda_X: \hat{X}\Rightarrow F for each of this copy. Naturally, just assign λX=αx\lambda_X = \alpha^{x} for (X,x)X^F(X, x) \to \hat{X} \to F.

Recall how we defined αX:X^F\alpha^X: \hat{X}\Rightarrow F in the proof of Yoneda lemma:

αYx:X^(Y)=MorC(Y,X)F(Y)fFf(x)\begin{darray}{rcl} \alpha^x_Y: \hat{X}(Y) &=& \mathrm{Mor}_{\mathcal{C}}(Y, X)\to F(Y) \\ f &\mapsto& Ff(x) \end{darray}

There is another way of thinking this. Let Fun(Cop,Set)/F\mathrm{Fun}(\mathcal{C}^{op}, \mathrm{Set}) / F denote this category of objects GFG\Rightarrow F, with morphisms G1G2G_1\to G_2 if G1G2FG_1\Rightarrow G_2\Rightarrow F commutes with G1FG_1\Rightarrow F. Let J\mathcal{J} be the full subcategory of X^F\hat{X}\Rightarrow F for all XCX\in\mathcal{C}. There are F(X)|F(X)| many of them, as a natural transformation α\alpha is uniquely decided by αX(idX)\alpha_X(\mathrm{id}_X), which can be any element in F(X)F(X). Let u:JFun(Cop,Set)u: \mathcal{J}\to\mathrm{Fun}(\mathcal{C}^{op}, \mathrm{Set}) be the forgetful functor:

u(X^F)=X^.u(\hat{X}\Rightarrow F) = \hat{X}.

Then, uu is the map from the diagram to representable functors.

YY is a union of multiple cocones, so it is a cocone, which can be proved exactly as what we did at the beginning. Now, we need to show that YY is a colimit.

Suppose F:CopSetF': \mathcal{C}^{op} \to \mathrm{Set} is another cocone, with natural transformations λX:X^F\lambda'^{X}: \hat{X}\Rightarrow F' for the diagram object (X,x)(X, x) in J\mathcal{J}. By Yoneda lemma, for all XX there exists an element uXF(X)u'_X\in F'(X) (namely uX=λXX(idX)u'_X = \lambda'^X_X(\mathrm{id}_X)) s.t. for all f:YXf: Y \to X, λYX(f)=Ff(uX)\lambda'^X_Y(f) = Ff(u'_X). Now, we can define a natural transformation β:FF\beta: F \Rightarrow F' as: βX(x)=uX\beta_X(x) = u'_X, i.e. mapping the deciding element to the deciding element. Note that uXu'_X depends on the pair (X,x)(X, x) and there is one for each xF(X)x\in F(X). Since for any X,YCX,Y\in\mathcal{C} and f:YXf: Y\to X,

βYFf(x)=uY=Ff(uX)=FfβX(x)\beta_Y\circ Ff(x) = u'_Y = F'f(u'_X) = F'f\circ \beta_X(x)

β\beta is natural. Clearly β\beta is unique. Thus, FF is a colimit.

Adjoint Functor Theorem

Given a small-complete category D\mathcal{D} with small morphism sets, a functor R:DCR:\mathcal{D}\to\mathcal{C} has a left joint if and only if it preserves all small limits and satisfying the following solution set condition.

Solution Set Condition: For each XCX\in\mathcal{C} there is a small set II and gi:XR(Yi)g_i: X\to R(Y_i) s.t. every h:XR(Y)h: X\to R(Y) can be written as h=R(f)gih = R(f)\circ g_i for some f:YiYf: Y_i\to Y.

Proof

If RR has a left adjoint L:CDL:\mathcal{C}\to\mathcal{D}, it must preserve all limits and the natural transformation η:IdCRL\eta: \mathrm{Id}_{\mathcal{C}}\Rightarrow RL satisfies the solution set condition, with II the one-point set. This is because for all h:XR(Y)h: X\to R(Y), we can find a unique αX,Y1(h)=f:L(X)Y\alpha_{X,Y}^{-1}(h) = f: L(X)\to Y s.t. h=R(f)ηXh = R(f)\circ \eta_X.

Conversely, given those conditions, it suffices to construct for each XCX\in\mathcal{C} a universal arrow ηX:XR(YX)\eta_X: X\to R(Y_X), where universality means for each h:XR(Y)h: X\to R(Y), we can find a unique f:YXYf: Y_X\to Y s.t. h=R(f)ηXh = R(f)\circ \eta_X. to RR. Then, RR has a left adjoint LL, by L(X)=YXL(X) = Y_X and L(h)L(h) being the unique ff. In other words, we need to find a initial object ηX,YX\langle\eta_X, Y_X\rangle in the comma category (XR)(X\downarrow R) for all XX.

Here, given XCX\in\mathcal{C} and F:DCF:\mathcal{D}\to\mathcal{C}, the comma category (XF)(X\downarrow F) is defined as a category with objects f,Y\langle f, Y\rangle s.t. f:XF(Y)f: X\to F(Y). Morphisms h:f,Yf,Yh: \langle f, Y\rangle\to\langle f', Y'\rangle are h:YYh: Y\to Y' s.t. f=F(h)ff'= F(h)\circ f.

Lemma: If F:DCF:\mathcal{D}\to\mathcal{C} preserve all small limits, then for each XCX\in\mathcal{C}, the projection:

QX:(XF)D(XF(Y))Y\begin{darray}{rcl} Q_X: (X\downarrow F) &\to& \mathcal{D} \\ (X\to F(Y)) &\mapsto& Y \end{darray}

creates all small limits in the comma category.

Proof of Lemma: Suppose fi:XF(Yi)f_i: X\to F(Y_i) is an II indexed family of objects in the comma category. Let (limIYi,λ)(\lim_I Y_i, \lambda) be a limit in D\mathcal{D}, and θ:YjYk\theta: Y_j\to Y_k s.t. fk=F(θ)fjf_k = F(\theta)\circ f_j. Since FF preserves limits, F(limIYi)F(\lim_I Y_i) is a limit of F(Yj)F(Y_j) and F(λk)=F(λj)F(θ)F(\lambda_k) = F(\lambda_j)\circ F(\theta). Then, there is a unique f:XF(limIYi)f: X\to F(\lim_I Y_i) s.t. fi=F(λi)ff_i = F(\lambda_i)\circ f for all iIi\in I. Thus, f=limIfif = \lim_I f_i.

Now, (XR)(X\downarrow R) is a small-complete category satisfying the solution set condition: there is a small set II and gi(XR)g_i \in (X\downarrow R), s.t. for every h(XR)h\in (X\downarrow R) there exists a morphism gihg_i\to h for some iIi\in I. It suffices to prove (XR)(X\downarrow R) has an initial object. This is true because (XR)(X\downarrow R) is small complete, we can construct a product of all gig_i and then construct an equalizer of all endomorphisms of it. Then, the equalizer is an initial object.

Actually this is what we did in the class to construct the free group functor. Check MacLane p123.

Sylow Theorem and Finite Groups

Notations

  • gh=ghg1{}^{g}h = ghg^{-1}: conjugation — 共軛(やく)
  • orbG(x)=Gx={gx:gG}\mathrm{orb}_G(x) = G\cdot x = \{ g\cdot x: g\in G \}: orbit — 軌道
  • stabG(x)={gG:gx=x}\mathrm{stab}_G(x) = \{g\in G: g\cdot x = x\} — 固定部分群
  • XG={xX:gx=x,gG}X^G = \{ x\in X: g\cdot x = x, \forall g\in G \} — 不変元
    • If GG is a finite pp-group, XXG modulo p\|X\| \equiv \|X^G\| \text{ modulo } p
  • Z(G)={gG:xg=gx,xG}Z(G) = \{ g\in G: xg = gx, \forall x\in G\}: center — 中心
    • If GG is a finite pp-group, pZ(G)p\| Z(G) and thus Z(G)>0Z(G) > 0.
  • n=vp(G)n = v_p(\|G\|): the highest power of pp that divides the order GGGGの位数に於けるppの重複度

Cauchy's Lemma

If pp divides G\|G\|, then GG contains an element of order pp.

Sylow Theorems

  • Existence: pp-Sylow always exists.
  • (a) Every pp-subgroup is contained in a pp-Sylow subgroup.
  • (b) Two Sylow subgroups are conjugate in GG.
  • (c) The number of Sylow subgroups is congruent to 11 modulo pp.
  • Corollary: A pp-Sylow is unique if and only if it is normal in GG.

Proof

Existence:

Inductive Proof: Prove by mathematical induction on the order. If there exists a proper subgroup H<GH < G of index prime to pp, then by induction HH has a pp-Sylow. If all proper subgroups have index divisible by pp, then

Z(G)=GGG0 (mod p)|Z(G)| = |G^G| \equiv |G| \equiv 0\ (\text{mod }p)

Here actions are conjugations. Thus, there exists an element xx of order pp. If vp(G)=1v_p(\|G\|) = 1, then x\langle x\rangle is a Sylow. If vp(G)>1v_p(\|G\|) > 1, then G/xG / \langle x\rangle has a pp-Sylow Pˉ=P/x\bar{P} = P / \langle x\rangle, which gives PP is a Sylow.

Constructive Proof: Consider YY being the set of all subsets of size pnp^{n} of GG, where n=vp(G)n = v_p(\|G\|). Consider GG acts on YY via left multiplication (coset). By Lucas theorem,

Y=(Gpn)(G/pn1)(00)n1=G/pn≢0 (mod p)|Y| = {|G| \choose p^n} \equiv {|G|/p^n \choose 1}{0 \choose 0}^{n-1} = |G|/p^n \not\equiv 0 \ (\text{mod } p)

Since Y=XY[G:stabG(X)]\|Y\| = \sum_{X\in Y}[G : \mathrm{stab}_G(X)], there exists some XX s.t. [G:stabG(X)][G : \mathrm{stab}_G(X)] is not divided by pp. Let H=stabG(X)H = \mathrm{stab}_G(X). Since [G:stabG(X)]=GH[G : \mathrm{stab}_G(X)] = \frac{\|G\|}{\|H\|}, HH is divided by pnp^n. By the definition of stablizer, HxXHx\subseteq X for all xXx\in X. Then, H=HxX=pn\|H\| = \|Hx\| \leq \|X\| = p^n. Thus, H=pn\|H\| = p^n.

Key Remark: If HGH\leq G is a pp-group and PGP\leq G is a pp-Sylow s.t. HH normalizes PP, then HPH\leq P.

Consider HP=HP\left\langle H\cup P \right\rangle = HP. By second isomorphism theorem, [HP:P]=[H:HP][HP: P] = [H: H\cap P] is a power of pp. Thus, HPHP is a pp-subgroup containing PP. Since PP is pp-Sylow, HP=PHP = P, which gives [H:HP]=1[H: H\cap P] = 1 and HPH\leq P.

(a) & (b):

Let P0P_0 be one Sylow subgroup and X={gP0:gG}X = \{ {}^gP_0: g\in G\} be the set of its conjugates. Clearly, XX is a set of pp-Sylow subgroups. Suppose HGH\leq G is a pp group and let HH acts on XX via conjugation.

XHX=[G:stabG(P0)]=[G:NG(P0)] (mod p)|X^H| \equiv |X| = [G : \mathrm{stab}_G(P_0)] = [G: N_G(P_0)] \ (\text{mod } p)

Since P0NG(P0)P_0\leq N_G(P_0), [G:NG(P0)][G: N_G(P_0)] is a factor of [G:P0][G: P_0], and thus prime to pp. Then, XH\|X^H\| is prime to pp. Thus, there exists PXHP\in X^H which is conjugate to P0P_0 and normalized by HH. Note that PP is a pp-Sylow subgroup, so by the key remark, HPH\leq P.

(c):

Consider the action P0P_0 on XX by conjugation. By the key remark, P0P_0 is a subgroup of any fixed point, so the only fixed point is P0P_0 itself. Hence, XXP0=1 (mod p)\|X\| \equiv \|X^{P_0}\| = 1\ (\text{mod } p). By (b), XX contains all Sylow groups.

Corollary:

X={P}X = \{ P \} iff. PP is normal in GG.

Remark:

X\|X\| divides G\|G\|, because it's the index of a stablizer.

Schur–Zassenhaus theorem

Suppose NGHN\to G \to H is a short exact sequence (s.e.s., 短完全系列) s.t. N\|N\| and H\|H\| are coprime. Then, the s.e.s. is a split; i.e. GNHG\cong N\rtimes H. Moreover, any two subgroups in GG of order H\|H\| are conjugate to each other.

A general case is NGN\unlhd G and H=G/NH = G / N s.t. N\|N\| and G/N\|G / N\| are coprime.

The general proof (especially part 2) is difficult and thus omitted here.

For a special case, refer to Prop. 2.7.20 in Math210 Notes, case (ii).

Application: Groups of Order 30

Any group of prime order is cyclic

By the fact that any subgroup generated by a non-trivial element is the whole group.

The only group of order 15 is cyclic

Since 15=3×515 = 3\times 5 and 55, 33 is not congruent to 1 modulo each other, G15G_{15} contains exactly one 55 subgroup and one 33 subgroup, which are all normal in G15G_{15}. Thus, G15C3×C5=C15G_{15}\cong C_3 \times C_5 = C_{15}.

The number of automorphisms of a cyclic group is its order's Euler function

Since Cn={1,c,,cn1}C_n = \{ 1, c, \ldots, c^{n-1} \} is gnerated by cc, an automorphism φ\varphi on it should be decided by φ(c)\varphi(c). Given that every ckc^k where kk is coprime to nn is a generator, there are ϕ(n)\phi(n) automorphisms. Actually, Aut(Cn)Cn×\mathrm{Aut}(C_n) \cong C_n^{\times}, with the latter one denotes the reduced residue system group of nn. (i.e. unit group of ring Z/n\mathbb{Z}/n)

When nn is prime, by the existence of primitive root, Aut(Cn)Cn1\mathrm{Aut}(C_n) \cong C_{n-1}.

There are 4 isomorphic types for 30

Let npn_p denote the number of pp-Sylow subgroups. By Sylow, n2{1,3,5,15}n_2\in \{1, 3, 5, 15\}, n3{1,10}n_3\in \{1, 10\}, n5{1,6}n_5\in \{1, 6\}. It is impossible for n3=10n5=6n_3 = 10 \wedge n_5 = 6, because otherwise there will be (31)×10(3-1)\times 10 elements of order 33 and (51)×6(5-1)\times 6 elements of order 55.

Then, at least one of the Sylow subgroup P3P_3 and P5P_5 will be normal in GG. Thus, by Second Isomorphism Theorem, N=P3P5GN = P_3P_5\leq G. By gcd(3,5)=1\rm{gcd}(3, 5) = 1, P3P5={1}P_3\cap P_5 = \{1\} and thus N=15\|N\| = 15, NC15N\cong C_{15}. [G:N]=2[G: N] = 2, so NN is normal.

Thus, by S-Z theorem, GC15C2G\cong C_{15} \rtimes C_2. Now consider the group action α\alpha of C2C_2 on C15C_{15}. We have Aut(C15)C3××C5×C2×C4\mathrm{Aut}(C_{15}) \cong C_3^{\times}\times C_5^{\times}\cong C_2\times C_4. Since 11 in C2C_2 has order 2, its image's order should divide 2. There are only 4 elements: {(0,0),(1,0),(0,2),(1,2)}\{(0,0),(1,0),(0,2),(1,2)\} or equivalently {[1],[11],[4],[14]}\{[1 ], [11 ], [4 ], [14 ]\} Thus, there are 4 isomorphic types. By the definition of \rtimes, for a,bC15a, b\in C_{15} and hC2h\in C_2, we have

  • (a,0)(b,h)=(a+b,h)(a, 0)\cdot (b, h) = (a+b, h);
  • (a,1)(b,h)=(a+α(b),h)=(a+α(1)b,h)(a, 1)\cdot (b,h) = (a+\alpha(b), h) = (a+\alpha(1)b, h).

Now let's inspect these four types:

  • For α(1)=[1]\alpha(1) = [1 ], we have C15C2C15×C2C30C_{15} \rtimes C_2\cong C_{15} \times C_2 \cong C_{30}.
    • If n,mn, m coprime, then (1,1)Cn×Cm(1,1)\in C_n\times C_m is of order nmnm and thus Cn×CmCnmC_n\times C_m\cong C_{nm}.
  • For α(1)=[14]=[1]\alpha(1) = [14 ] = [-1 ], C15C2D15C_{15} \rtimes C_2\cong D_{15}.
    • DnCn1C2D_n \cong C_n\rtimes_{-1} C_2, with (p,q)(p,q) mapping to rpsqr^ps^q.
  • For α(1)=[4]\alpha(1) = [4 ], the element (1,1)(1,1) has order 66 ((1,1)(5,0)(6,1)(10,0)(11,1)(0,0)(1,1)\to(5,0)\to(6,1)\to(10,0)\to(11,1)\to(0,0)). Thus, we have C15C2C3×D5C_{15} \rtimes C_2\cong C_3\times D_{5}.
    • With (0,0)(0,e)(0,0)\mapsto (0,e), (0,1)(0,s)(0,1)\mapsto (0,s), (1,0)(1,r)(1,0)\mapsto (1,r), (6,0)(0,r),(10,0)(1,e)(6,0)\mapsto (0,r), (10,0)\mapsto (1,e).
  • For α(1)=[11]=[4]\alpha(1) = [11 ] = [-4 ], the element (1,1)(1,1) has order 1010. We have C15C2C5×S3C_{15} \rtimes C_2\cong C_5\times S_3.
    • With (0,0)(0,())(0,0)\mapsto(0,()), (6,0)(1,())(6,0)\mapsto(1,()), (0,1)(0,(12))(0,1)\mapsto(0,(12)), (0,1)(0,(23))(0,1)\mapsto(0,(23)), (5,1)(0,(23))(5,1)\mapsto(0,(23)), (10,0)(0,(123))(10,0)\mapsto(0,(123)), (5,0)(0,(123))(5,0)\mapsto(0,(123)).

Misc on Group

Five Lemma (5項補題)

It's hard to show without figure (>_<) Let G1G2G3G4G5G_1 \to G_2 \to G_3 \to G_4 \to G_5 and H1H2H3H4H5H_1 \to H_2 \to H_3 \to H_4 \to H_5 be exact (完全) at 2,3,4. Let homomorphisms fi:GiHif_i: G_i \to H_i be commutative in the square diagram. Then, if f1,f2,f4,f5f_1, f_2, f_4, f_5 are isomorphisms, so is f3f_3.

Proof

Prove by chasing on two 4-squares. Let αi:GiGi+1\alpha_i: G_i\to G_{i+1} and βi:HiHi+1\beta_i: H_i\to H_{i+1} denote those exact homomorphisms.

Claim: f3f_3 is onto:

  • Arbitrarily pick h3H3h_3\in H_3.
  • Since f4f_4 is surjective, there exists g4G4g_4\in G_4 s.t. f4(g4)=β3(h3)f_4(g_4) = \beta_3(h_3).
  • By exactness at H4H_4, β5β4(h3)=1\beta_5\circ\beta_4(h_3) = 1.
  • By commutativity, f5α4(g4)=β4f4(g4)=β5β4(h3)=1f_5\circ\alpha_4(g_4) = \beta_4\circ f_4(g_4) = \beta_5\circ\beta_4(h_3) = 1.
  • Since f5f_5 is injective, α4(g4)=1\alpha_4(g_4) = 1; thus, g4Im(α3)g_4\in\mathrm{Im}(\alpha_3).
  • Pick g3G3g_3\in G_3 s.t. α3(g3)=g4\alpha_3(g_3) = g_4.
  • Let z=f3(g3)1h3z = f_3(g_3)^{-1}h_3.
  • Then, β3(z)=f4(g4)1β3(h3)=1\beta_3(z) = f_4(g_4)^{-1}\beta_3(h_3) = 1.
  • By the exactness at H3H_3, zIm(β2)z\in \mathrm{Im}(\beta_2).
  • Since f2f_2 is surjective, there exists g2G2g_2\in G_2 s.t. β2f2(g2)=z=f3α3(g2)\beta_2\circ f_2(g_2) = z = f_3\circ\alpha_3(g_2).
  • Thus, f3(g3α3(g2))=h3f_3(g_3\cdot\alpha_3(g_2)) = h_3.

Similarly, f3f_3 is one to one. Thus, f3f_3 is an isomorphism.

References

  • Saunders Mac Lane, Categories for the working mathematician, Springer Science & Business Media, 2013.
  • Paul Balmer, UCLA MATH210A, Fall 2020.