tinytheorems - about groups - rings - modules - models atom/rss

Exploration 3: Polynomials

.

Questions:


Recall that to adjoin an element aa to a ring RR is to create a ring R[a]R[a] generated by the two: something like a,R\<a,R\>.

When you adjoin a symbol xx that commutes with the entire ring RR, the result is a polynomial ring. It turns out polynomial rings are fundamental in the sense that much of ring theory is built upon generalizing the properties of polynomials, such as irreducibility.

The elements ff of a polynomial ring R[x]R[x] are called polynomials with coefficients in RR, and look like this:

f f eq = xn a n x n p1 + ldots ... p2 + x2 a 2 x 2 p3 + x a 1 x indeterminate indeterminate x x->indeterminate p4 + c a 0 leading_coefficient leading coefficient a n leading_coefficient->xn degree degree n degree:s->xn constant_coefficient constant coefficient a 0 constant_coefficient->c

where aiRa_i\in R are known as the coefficients of the polynomial ff, xx is known as an indeterminate over RR, and the exponent of the leading xix^i is known as the degree of the polynomial, denoted deg f\deg f. The degree of a constant polynomial cRc\in R is zero, and the degree of the zero polynomial 00 is undefined.

Let’s show some basic facts that hold when we form a polynomial ring.

Theorem: RR is an integral domain iff R[x]R[x] is an integral domain.
  • WTS the product of two nonzero polynomials f,gf,g in R[x]R[x] is always nonzero.
  • Let ana_n and bmb_m be the leading coefficients of f,gf,g respectively. Since f,gf,g are nonzero, ana_n and bmb_m are nonzero.
  • Then the leading coefficient of fgfg is anbma_nb_m. Since RR is an integral domain, this product is nonzero.
  • Then since the leading coefficient of fgfg is nonzero, fgfg is nonzero, and we are done.

In this section, we define the operations possible on polynomials.

Polynomial addition/subtraction is straightforward – it is pairwise addition/subtraction of corresponding coefficients.

Polynomial multiplication is more complex. In general, it looks like (i=0nrixi)(j=0msjxj)=(k=0n+mrksn+mkxk)(\sum_{i=0}^n r_ix^i)(\sum_{j=0}^m s_jx^j)=(\sum_{k=0}^{n+m} r_ks_{n+m-k}x^k) Note that this means that the resulting coefficients are the convolution of the coefficients rir_i and sjs_j. Also note that this implies that deg fg=deg f+deg g\deg fg=\deg f+\deg g.

What about polynomial division? Dividing a polynomial ff by a polynomial gchar "338=0g\ne 0 with remainder rr means finding unique polynomials q,rq,r where deg r<deg g\deg r<\deg g (or r=0r=0) such that f=gq+rf=gq+r. Let’s explore how this can be done.

Therefore: For any two polynomials f,gR[x]f,g\in R[x] (where gchar "338=0g\ne 0) we can find unique polynomials q,rq,r (where deg r<deg g\deg r<\deg g or r=0r=0) such that f=gq+rf=gq+r, but only if the leading coefficient of gg is a unit. This theorem is known as the division algorithm.
  • Let n=deg fn=\deg f and m=deg gm=\deg g.
  • First of all, if f=0f=0 then the only solution is q=r=0q=r=0. Otherwise if n<mn<m, then the only solution is q=0q=0 and r=fr=f. Therefore we can induct on nn with the assumption that nmn\ge m.
  • If n=0n=0, then nmn\ge m implies m=0m=0 and therefore f,gf,g are both constant polynomials R\in R.
    • Then we have f=qg+rf=qg+r in RR. In RR, the condition that either r=0r=0 or deg r<m\deg r<m collapses to just r=0r=0 since m=0m=0.
    • Since r=0r=0, f=qg+rf=qg+r becomes qg=fqg=f in RR. Thus the solution is q=fg1q=fg^{-1} and r=0r=0. This requires gg be a unit in RR so when gg is constant it must be a unit.
  • Otherwise, assume that for all deg f<n\deg f<n, there is some q,rq,r such that f=gq+rf=gq+r. WTS that’s true for deg f=n\deg f=n as well.
    • With the intention of applying the induction hypothesis, we’d like to obtain a polynomial of degree less than nn. This is easy when ff and gg have the same leading coefficient and degree; then fgf-g has degree less than nn. Otherwise, we must use fkgf-kg where kk is a factor that transforms the leading coefficient and degree of gg to match those of ff.
    • Let f=inaixif=\sum_i^n a_ix^i and g=jmbjxjg=\sum_j^m b_jx^j, where ana_n and bmb_m are nonzero. Then the value of kk that lets us do this is k=anbm1xnmk=a_nb_m^{-1}x^{n-m}.
    • Note that this means that the leading coefficient bmb_m of gg must be a unit. (This matches our earlier finding that when gg is constant, it must be a unit.) Then we have kg=kjmbjxjkg=k(bmxm+jm1bjxj)kg=anbm1xnm(bmxm+jm1bjxj)kg=anxn+jm1anbm1bjxj+nm\begin{aligned} kg&=k\sum_j^m b_jx^j\\ kg&=k\left(b_mx^m+\sum_j^{m-1} b_jx^j\right)\\ kg&=a_nb_m^{-1}x^{n-m}\left(b_mx^m+\sum_j^{m-1} b_jx^j\right)\\ kg&=a_nx^n+\sum_j^{m-1} a_nb_m^{-1}b_jx^{j+n-m} \end{aligned}
    • Since ff and kgkg have the same leading term anxna_nx^n, the difference fkgf-kg cancels out the leading terms and therefore has degree less than nn. Then the induction hypothesis with fkgf-kg and gg gives us unique polynomials q,rq',r' such that fkg=qg+rf-kg=q'g+r (where either r=0r=0 or deg r<deg g\deg r<\deg g). fkg=qg+rf=(k+q)g+r\begin{aligned} f-kg&=q'g+r\\ f&=(k+q')g+r'\\ \end{aligned}
    • Thus we have a solution q=k+qq=k+q' and r=rr=r'.

Corollary: The division algorithm above produces a unique solution q,rq,r.

To see this, towards contradiction assume we have any two distinct solutions q,rq,r and q,rq',r' where f=gq+r=gq+rf=gq+r=gq'+r implies g(qq)=rrg(q-q')=r'-r.

  • We know that deg r<deg g\deg r<\deg g and therefore the RHS rrr'-r has degree <deg g<\deg g.
  • Then the LHS g(qq)g(q-q') has degree deg g+deg (qq)\ge\deg g+\deg(q-q').
  • Since the LHS and RHS should have the same degree, this is a contradiction.

In summary, we always have addition, subtraction, and product defined for R[x]R[x]. But you can only divide by polynomials whose leading coefficient is a unit in RR. This means polynomial rings in general don’t admit a division algorithm, because they have nonzero polynomials that you can’t divide by. Rings that do admit such a division algorithm are known as Euclidean domains.

Theorem: If RR is a field, then R[x]R[x] is a Euclidean domain.

You can only divide by nonzero polynomials in R[x]R[x] whose leading coefficient is a unit in RR, but every leading coefficient is a unit when RR is a field. Thus you can divide by any nonzero polynomial in R[x]R[x], making R[x]R[x] a Euclidean domain.

Corollary: For any two polynomials f,gR[x]f,g\in R[x] where gg is a nonzero monic polynomial (a polynomial whose leading coefficient is 11), we can find unique polynomials q,rq,r (where deg r<deg g\deg r<\deg g or r=0r=0) such that f=gq+rf=gq+r.

Following from the division algorithm, you can only divide by polynomials whose leading coefficient is a unit, and 11 is always a unit. Therefore you can always divide by monic polynomials.

Corollary: Same, but for nonzero polynomials whose leading coefficient is a unit.

Theorem: Every field is a Euclidean domain.

Since every nonzero element in a field is a unit, you can already divide by every element in a field: a=qba=qb. Therefore all fields have a division algorithm and are Euclidean domains.

Note that the division algorithm in question requires a notion of degree for every nonzero element in the ring. To generalize the notion of degree to non-polynomial rings, this means having some function ϕ:R{0}N\phi:R\setminus\{0\}\to\NN mapping each nonzero to a natural number 0\ge 0 called the Euclidean norm, where ϕ(r)=0\phi(r)=0 for all units rRr\in R. The Euclidean norm also needs to satisfy the divisibility condition where for every aRa\in R and nonzero bRb\in R, there is a division a=bq+ra=bq+r where either r=0r=0 or ϕ(r)<ϕ(b)\phi(r)<\phi(b). (We don’t need to mention the norm ϕ\phi in the above proof, since r=0r=0 in that proof.)

Theorem: If a Euclidean norm ϕ\phi exists in an integral domain RR, then a division algorithm exists in RR (making RR a Euclidean domain.)

The Euclidean condition norm requires that for every a,bRa,b\in R, we have either a=qb+ra=qb+r or where either r=0r=0 or ϕ(r)<ϕ(b)\phi(r)<\phi(b). This is exactly the required division.

Then in general, a Euclidean domain is an integral domain for which a Euclidean norm is defined.

Remember that for polynomial rings R[x]R[x] that are not Euclidean domains, you can only divide by polynomials whose leading coefficient is a unit in RR. Then:

Theorem: Let ff and gg be nonzero monic polynomials, each of which divides the other. Then f=gf=g.
  • If ff and gg divide each other, we have f=gqf=gq and g=fqg=fq'.
  • deg f=deg gq\deg f=\deg gq implies deg fdeg g\deg f\ge\deg g.
  • deg g=deg fq\deg g=\deg fq' implies deg gdeg f\deg g\ge\deg f.
  • Therefore deg f=deg g\deg f=\deg g, which implies that q,qq,q' are constant polynomials that don’t affect the degree of the product.
  • The fact that both f,gf,g are monic shows that the leading coefficient is 11 before and after multiplying by q,qq,q’, which implies that q,qq,q' are both 11.
  • Therefore f=gf=g.

This last theorem actually applies in all integral domains. Let’s modify the statement a bit:

Theorem: Let ff and gg be nonzero elements in an integral domain RR. Each of ff and gg generates the other iff they differ by a unit.
  • ff and gg being nonzero in an integral domain implies that neither of them are zero divisors.
  • If ff and gg generate each other, we have f=gqf=gq and g=fqg=fq', and therefore f=fqqf=fqq'.
  • Then 0=f(qq1)0=f(qq'-1) where ff is not a zero divisor, implying qq1=0qq'-1=0.
  • This means qq=1qq'=1, i.e. q,qq,q' are units.
  • Thus f=gqf=gq means ff differs from gg by a unit.
  • The converse is trivial — if f,gf,g differ by a unit uu, then they generate each other via the unit: f=ugf=ug and g=u1fg=u^{-1}f.

In this section, we explore the consequences of evaluating polynomials.

One of the most important things we can do with polynomials is evaluation, in which we substitute xx with aRa\in R.
Therefore: we can evaluate a polynomial to an element of RR by substituting xx with bb.
  • Such a mapping is always a ring homomorphism.
  • The resulting expression iaibi\sum_i a_ib^i is a rational expression, made up of elements of RR connected with addition, subtraction, multiplication, integer scalar multiplication, powers, and division by units.
  • Since ring homomorphisms preserve rational expressions, the resulting expression is an element of RR.

This substitution is formalized as the evaluation map (a ring homomorphism) φb:R[x]R\varphi_b:R[x]\to R. Applying the evaluation map to a polynomial ff can be denoted φb(f)\varphi_b(f), but is more often denoted f(b)f(b), the evaluation of ff at bb.

Lemma: The evaluation map is surjective.

Since there is a constant polynomial rR[x]r\in R[x] for every rRr\in R, and constant polynomials remain unchanged by the evaluation map, every element of RR gets mapped to.

Theorem: R[x]/(x)RR[x]/(x)\iso R.
  • The evaluation map φ0:R[x]R\varphi_0:R[x]\to R, which evaluates a polynomial at 00, has kernel (x)(x). To see this, notice that for every fR[x]f\in R[x], φ0(f)=f(0)=0\varphi_0(f)=f(0)=0 implies the constant coefficient is 00, and therefore f(x)=anxn++a2x2+a1x+0a0f(x)=anxn++a2x2+a1xf(x)=x(anxn1++a2x+a1)f(x)(x)\begin{aligned} f(x)&=a_nx^n+\ldots+a_2x^2+a_1x+0a_0\\ f(x)&=a_nx^n+\ldots+a_2x^2+a_1x\\ f(x)&=x(a_nx^{n-1}+\ldots+a_2x+a_1)\\ f(x)&\in (x) \end{aligned} so the kernel of φ0\varphi_0 is (x)(x).
  • The first ring isomorphism theorem says R[x]/ker φ0im φ0R[x]/\ker\varphi_0\iso\im\varphi_0. We just showed ker φ=(x)\ker\varphi=(x), and since φ0\varphi_0 is surjective, im φ0=R\im\varphi_0=R, therefore the above becomes R[x]/(x)RR[x]/(x)\iso R.

Remember that you can always divide by a monic polynomial in polynomial rings. We’ll see that division relates to evaluation in multiple ways:

Remainder Theorem: When fR[x]f\in R[x] is divided by xax-a, the remainder is f(a)f(a).

Since xax-a is monic, the division algorithm implies some unique q,rq,r such that f=(xa)q+rf=(x-a)q+r. Evaluating at aa gives f(a)=(aa)q+r=rf(a)=(a-a)q+r=r. Therefore, the remainder rr is f(a)f(a).

Factor Theorem: In polynomial rings over a field F[x]F[x], f(a)=0f(a)=0 iff f=(xa)qf=(x-a)q for some unique polynomial qq.

The evaluation f(a)f(a) replaces xx with aa. If f=(xa)qf=(x-a)q, then f(a)=(aa)q=0q=0f(a)=(a-a)q=0q=0. Conversely, if f(a)=0f(a)=0, then since xax-a is monic, by the remainder theorem ff divided by xax-a results in f=(xa)q+f(a)f=(x-a)q+f(a). But f(a)=0f(a)=0, so this becomes f=(xa)qf=(x-a)q.

When f(a)=0f(a)=0, then aa is called a root of ff, and the following are equivalent:

Since the factor theorem lets you factor a polynomial by finding its roots, finding the roots of polynomials is pretty important if you want to factor polynomials.

Theorem: A degree nn polynomial has at most nn roots.

This is easily proved by induction by applying the factor theorem nn times.

The theorems below are useful for factoring a polynomial (i.e. find roots) in a given polynomial ring. Here is an important one for integer polynomials Z[x]\in\ZZ[x]:

Rational Roots Theorem: Let f=a0+a1x+a2x2++anxnZ[x]f=a_0+a_1x+a_2x^2+\ldots+a_nx^n\in\ZZ[x]. If a0char "338=0a_0\ne 0 and anchar "338=0a_n\ne 0, then for every rational root c/dc/d, cc divides the constant a0a_0 and dd divides the leading coefficient ana_n.
  • Let c/dc/d be a rational root of the polynomial f=a0+a1x+a2x2++anxnZ[x]f=a_0+a_1x+a_2x^2+\ldots+a_nx^n\in\ZZ[x].
  • Assume c,dc,d are coprime integers – if they are not, divide both cc and dd by their GCD to make them coprime.
  • Since c/dc/d is a root of ff, we have f(c/d)=0f(c/d)=0: a0+a1(c/d)+a2(c/d)2++an(c/d)n=0a_0+a_1(c/d)+a_2(c/d)^2+\ldots+a_n(c/d)^n=0
  • Multiply both sides by dnd^n: a0dn+a1dn1c+a2dn2c2++ancn=0a_0d^n+a_1d^{n-1}c+a_2d^{n-2}c^2+\ldots+a_nc^n=0
  • From here, we can go in two directions. First, isolate the a0dna_0d^n term and factor out cc from the other terms: c(a1dn1+a2dn2c++ancn1)=a0dnc(a_1d^{n-1}+a_2d^{n-2}c+\ldots+a_nc^{n-1})=-a_0d^n This means cc is a factor of a0dn-a_0d^n. Since c,dc,d are coprime integers, cc must divide a0a_0.
  • Second, isolate the ancna_nc^n term instead and factor out dd from the other terms: d(a0dn1+a1dn2c+a2dn3c2++an1cn1)=ancnd(a_0d^{n-1}+a_1d^{n-2}c+a_2d^{n-3}c^2+\ldots+a_{n-1}c^{n-1})=-a_nc^n Similarly, this means dd is a factor of ancn-a_nc^n. Since c,dc,d are coprime integers, dd must divide ana_n.

Corollary: when ff is monic, the only rational roots are all the integer factors of a0a_0.

To remember this theorem, you can think about the polynomial xcx-c which obviously has a root c/1c/1. So the numerator divides the constant cc, and the denominator divides the leading coefficient 11.

Corollary: m\sqrt{m} is irrational (char "338Q\notin\QQ) unless mm is the square of an integer.
  • Try to interpret m\sqrt{m} as a rational root (Q\in\QQ) of some polynomial Z[x]\in\ZZ[x].
  • If mQ\sqrt{m}\in\QQ, it would be a rational root c/dc/d of the polynomial x2mx^2-m. We know m\sqrt{m} is one such root.
  • By Rational Roots Theorem, cmc\mid m and d1d\mid 1, therefore d=±1d=\pm 1. Then c/d=±cc/d=\pm c so the root m\sqrt{m} must be some integer ±c\pm c. Therefore mm is the square of some integer.

Corollary: mnchar "338Q\sqrt[n]{m}\notin\QQ unless mm is the nnth power of an integer.
  • (Same proof as above)
  • If mnQ\sqrt[n]{m}\in\QQ, it would be a rational root c/dc/d of the polynomial xnmx^n-m. We know mn\sqrt[n]{m} is one such root.
  • By Rational Roots Theorem, cmc\mid m and d1d\mid 1, therefore d=±1d=\pm 1. Then c/d=±cc/d=\pm c so the root mn\sqrt[n]{m} must be some integer ±c\pm c. Therefore mm is the nnth power of some integer.

In this section, we classify the unique factorization of real and complex polynomials.

All the work involved in factoring complex polynomials rests on this one theorem:

Fundamental Theorem of Algebra (FTA): If fC[x]f\in\CC[x] is a nonconstant polynomial, then ff has a root in C\CC.

The proof often requires complex analysis or topology. Here is a proof due to Artin, trying not to use many ideas outside of ring theory:

  • It is a theorem that if you treat evaluation of the polynomial fC[x]f\in\CC[x] as a map CC\CC\to\CC, then ff is continuous.
  • Evaluate ff at each point on a circle CrC_r of radius rr around the origin of the complex plane. Since ff is continuous, the images f(Cr)f(C_r) represent some loop on the complex plane. Each point on the circle CrC_r can be represented in polar coordinates as z=reiθz=re^{i\theta} for some angle θ\theta. Let its corresponding point on the loop f(Cr)f(C_r) be f(z)f(z).
  • There are two loops we want to consider:
    • First, with rr approaching 00, we know the images f(Cr)f(C_r) are all close to the constant coefficient c0c_0 of ff. So f(Cr)f(C_r) is a tiny loop around c0c_0. We assume c0c_0 is nonzero – if it’s zero then f(0)=0f(0)=0 implies 00 is a root and we are done.
    • Second, with rr approaching \infty, we know the images f(Cr)f(C_r) are very large but also represent some loop. Since for large values the leading term cnznc_nz^n of f(z)f(z) dominates the other terms, which we can write as f(z)cnznf(z)-c_nz^n, we know f(z)cnzn<cnzn|f(z)-c_nz^n|<|c_nz^n| and (because cnzn=cnrneinθ=cnrn|c_nz^n|=c_n|r^ne^{in\theta}|=c_nr^n) therefore f(z)cnzn<cnrn|f(z)-c_nz^n|<c_nr^n. This implies the distance between f(z)f(z) and the leading term cnznc_nz^n is always less than cnrnc_nr^n, no matter what θ\theta is. If you imagine θ\theta increasing, then as cnzn=cnrneinθc_nz^n=c_nr^ne^{in\theta} walks a circle of radius cnrnc_nr^n around the origin, while f(z)f(z) (being continuous) is following a distance less than cnrnc_nr^n behind and therefore the loop it traces must also enclose the origin.
  • Importantly, the first loop does not enclose the origin, while the second one does. Since ff is continuous, varying rr within (0,)(0,\infty) will continously vary the corresponding loop between one that doesn’t enclose the origin and one that does. So at some rr between 00 and \infty, the resulting loop crosses the origin, and therefore we get a root f(reiθ)=0f(re^{i\theta})=0 for some θ\theta.

Corollary (FTA in C[x]\CC[x]): For every complex polynomial fC[x]f\in\CC[x], you can write it in the form f=ki(xui)f=k\prod_i(x-u_i) where kk is the leading coefficient and uiu_i are all the roots.

This is repeated application of the FTA. You apply FTA to find a root aa, factor it out with the factor theorem to get f=(xa)gf=(x-a)g, and repeat with gg until gg is a constant polynomial kk. Since we’re always factoring out a monic polynomial xax-a, the leading coefficient never changes, and so the final kk is the leading coefficient of ff.

Conjugate Root Theorem: if a+biCa+bi\in\CC is a root of a real polynomial fR[x]f\in\RR[x], then abia-bi is also a root.
  • Since complex conjugation x\overline{x} fixes the real numbers, f(a+bi)=f(a+bi)f(\overline{a+bi})=\overline{f(a+bi)} when ff is a real polynomial.
  • Therefore, if f(a+bi)=0f(a+bi)=0, then f(a+bi)=0\overline{f(a+bi)}=\overline{0} and therefore f(abi)=f(abi)=0\overline{f(\overline{a-bi})}=f(a-bi)=0 implies abia-bi is also a root of ff.

Corollary (FTA in R[x]\RR[x]): For every real polynomial fR[x]f\in\RR[x], you can write it in the form f=ki(xui)iqif=k\prod_i(x-u_i)\prod_iq_i where kk is the leading coefficient, uiu_i are all the real roots, and qiq_i are all the monic irreducible real quadratics.
  • Do the same thing you did for complex polynomials, factoring ff into a product f=ki(xui)f=k\prod_i(x-u_i) where the roots uiu_i are complex.
  • Since the coefficients of ff are real, by the Conjugate Root Theorem, the complex roots come in conjugate pairs. Then the product of their corresponding factors, (x(a+bi))(x(abi))=x22ax+(a2+b2)(x-(a+bi))(x-(a-bi))=x^2-2ax+(a^2+b^2), is a monic irreducible real quadratic, i.e one of the qiq_i.
  • Therefore, after combining all complex factors into monic real quadratic factors, ff can be written in the form ki(xui)iqik\prod_i(x-u_i)\prod_iq_i where qiq_i are the product of each conjugate pair of complex factors within the original uiu_i.

Corollary: all irreducible polynomials in R[x]\RR[x] are linear or quadratic (have degree 11 or 22).

We see that the Fundamendal Theorem of Algebra implies that (nonconstant) polynomials in C[x]\CC[x] or R[x]\RR[x] factor uniquely into a constant times a product of (monic) irreducible factors.

In this section, we examine the ideals of a polynomial ring.

Let’s explore what the ideals AA of a polynomial ring R[x]R[x] look like.

Therefore: If R[x]R[x] is a Euclidean domain, then every ideal in R[x]R[x] is principal – generated by a single element, in this case a unique monic polynomial.
  • If AA is the zero ideal, it is generated by 00 and we are done.
  • Otherwise, AA contains a nonzero polynomial.
  • We can divide by monic polynomials, but AA doesn’t necessarily contain one, unless R[x]R[x] is a Euclidean domain – then we can make any polynomial monic by dividing the polynomial by the leading coefficient. Since any product with an element in the ideal AA is also in AA, AA always contains a monic polynomial.
  • Take a monic polynomial of minimal degree gg in AA. Every polynomial fAf\in A must have gg as a factor because the division algorithm lets us divide by monic polynomials.
  • Then the division algorithm says f=gq+rf=gq+r where either r=0r=0 or deg r<deg g\deg r<\deg g.
  • Since r=fgqr=f-gq, rr must also be in AA. We can also make rr monic and the result is also in AA.
  • Since deg r<deg g\deg r<\deg g contradicts gg being the monic polynomial of minimal degree in AA. Therefore r=0r=0.
  • This means f=gqf=gq. Since ff was arbitrary, gg generates AA.
  • To show that gg uniquely generates AA, assume that hh is another generator of AA. Then gg and hh generate each other: g=hqg=hq and h=gqh=gq'. But monic polynomials that divide each other are equal, so g=hg=h.
    • deg g=deg hq\deg g=\deg hq implies deg gdeg h\deg g\ge\deg h, and deg h=deg gq\deg h=\deg gq' implies deg hdeg g\deg h\ge\deg g. Therefore deg g=deg h\deg g=\deg h.
    • deg g=deg h\deg g=\deg h implies that q,qq,q' are constant polynomials that don’t affect the degree of the product. The fact that both g,hg,h are monic (leading coefficient is 11 before and after multiplying by q,qq,q’) implies that q,qq,q' are both 11. This implies g=hg=h.

This means we can write all quotient rings of F[x]F[x] (FF a field) as F[x]/(h)F[x]/(h), where hh is some monic polynomial. Since the generator of every ideal is unique, this implies a bijection between monic polynomials in F[x]F[x] and nonzero ideals in F[x]F[x].

When every ideal of a ring is principal, we have a principal ideal domain (PID).

It turns out every Euclidean domain is a PID.

Theorem: All Euclidean domains are PIDs.
  • Let RR be a Euclidean domain with norm ff. Then for every element aRa\in R, we can divide aa by nonzero bRb\in R to get a=bq+ra=bq+r where either r=0r=0 or f(r)<f(b)f(r)<f(b).
  • We can show that for every ideal II in RR, every element aa of II is generated by some element $b$ with the smallest norm f(b)f(b). That is, a=bqa=bq for some qRq\in R.
  • This follows immediately from the division algorithm. We have a=bq+ra=bq+r, and since f(b)f(b) is minimal by definition, there is no f(r)f(r) such that f(r)<f(b)f(r)<f(b), therefore r=0r=0.

In this section, we determine what it means to quotient in polynomial rings defined over a field FF.

In particular, when FF is a field, then F[x]F[x] is a Euclidean domain and therefore a PID. Let’s explore polynomial rings defined over a field FF.

Therefore: The elements of F[x]/(h)F[x]/(h) are every polynomial in F[x]F[x] with degree less than deg h\deg h, under the relation h(x)=0h(x)=0.
  • When you quotient by some ideal h\<h\>, you’re effectively sending the polynomial hh to 00 and therefore enforcing the relation h(x)=0h(x)=0 on the polynomial ring. That means whenever f=hqf=hq, f=0f=0.
  • But in a Euclidean domain, every polynomial ff of degree at least deg h\deg h can factor out hh: f=hqf=hq. This means every polynomial of degree deg h\deg h and above gets sent to 00.
  • The remaining polynomials are necessarily of degree less than deg h\deg h.

Example: F[x]/(x2)F[x]/(x^2)\iso all linear and constant polynomials in F[x]F[x].

We identify x2x^2 with 00, so all the degree 2+2+ polynomials get sent to 00, leaving only the degree 11 and 00 polynomials (and the zero polynomial).

Example: R[x]/(x2+1)C\RR[x]/(x^2+1)\iso\CC.
  • Here we identify x2+1x^2+1 with 00. We can understand this as x=1=ix=\sqrt{-1}=i (in C\CC).
  • Since every resulting polynomial is at most degree 11, they are in the form c0+c1xc_0+c_1x.
  • In other words, the elements of R[x]/(x2+1)\RR[x]/(x^2+1) are c0+c1ic_0+c_1i.
  • This is exactly how we write elements a+bia+bi of C\CC, so the two rings are isomorphic.

Example: Q[x]/(x32)Q(23)\QQ[x]/(x^3-2)\iso\QQ(\sqrt[3]{2}).
  • Here we identify x32x^3-2 with 00. We can understand this as x=23x=\sqrt[3]{2} (in R\RR).
  • Since every resulting polynomial is at most degree 22, they are in the form c0+c1x+c2x2c_0+c_1x+c_2x^2.
  • In other words, the elements of R[x]/(x2+1)\RR[x]/(x^2+1) are c0+c123+c2232c_0+c_1\sqrt[3]{2}+c_2\sqrt[3]{2}^2.
  • But this is the same as adjoining an element 23\sqrt[3]{2} to Q\QQ.

To generalize these last two examples, when FF is a field and hh is irreducible, we can take the quotient F[x]/(h)F[x]/(h) to form a ring isomorphic to F[c]F[c], where cc is a solution to h=0h=0 (i.e. a root of hh). It turns out the result is a field:

Theorem: F[x]/(h)F[x]/(h) is a field iff hh is irreducible.
  • Since F[x]F[x] is a PID, the result follows directly from this theorem.
  • (Note that if hh is reducible, then F[x]/(h)F[x]/(h) sends h=fgh=fg to 00, meaning there are zero divisors f,gf,g so the result fails to be even an integral domain.)

In other words, the quotient rings of polynomial rings can be used to construct field extensions, which we’ll explore in depth later on.

In this section, we introduce how the derivative of a polynomial helps determine roots.

The derivative ff' of the polynomial fF[x]f\in F[x] can be constructed by taking each term of ff (aixia_ix^i), and mapping it to iaixi1ia_ix^{i-1}, where ii is mapped into FF using the canonical homomorphism ZF\ZZ\to F (i.e. taking 1+1+1+1+\ldots but ii times.)

The derivative helps us determine when a polynomial has multiple roots. α\alpha is a multiple root of ff if ff has a factor (xα)n(x-\alpha)^n for some n2n\ge 2.

Theorem: If the derivative of a polynomial ff shares a root α\alpha with ff, then α\alpha is a multiple root of ff.
  • Since ff has α\alpha as a root, factor out f=(xα)gf=(x-\alpha)g. To show that α\alpha is a multiple root, we can show that gg also has α\alpha as a root, i.e. g(α)=0g(\alpha)=0.
  • By the product rule, f=(xα)g+gf'=(x-\alpha)g'+g. Since f(α)=0f'(\alpha)=0, we have 0=(αα)g(α)+g(α)0=(\alpha-\alpha)g'(\alpha)+g(\alpha), implying 0=g(α)0=g(\alpha), implying α\alpha is a root of gg.

Corollary: If the derivative of an irreducible polynomial ff is nonzero, then ff and ff' share no factors.

Irreducible polynomials ff can only share a factor (itself) with either 0 or polynomials of greater or equal degree. Since the derivative ff' always has lesser degree, ff' has to be zero in order for ff and ff' to share a factor.

< Back to category Exploration 3: Polynomials (permalink)
Exploration 2: Ring homomorphisms Exploration 4: Factorization