arXiv:2101.05610v1 [math.GM] 12 Jan 2021 cuaeagbacfruafrtequintic the for formula algebraic Accurate ouinb trto fradicals of iteration by Solution be is,CrfYoussfi Chrif Missa, Abdel aur 5 2021 15, January & 1 Contents

1 Introduction 3

2 Reduced Forms 4

3 Iteration of radicals algorithm to solve equations in Form 3 5 3.1 Algorithm ...... 5 3.2 Unique root near the positive real axis ...... 7 3.3 Absoluteerrortheorem...... 8 3.3.1 Starting point properties ...... 9 3.3.2 Proofofabsoluteerrortheorem ...... 11 3.4 Relativeerrortheorem ...... 20 3.5 Speedofconvergence ...... 21

4 Solving equations in Form 2 24

5 Solving equations in Form 1 25 5.1 Challenge with iterative radicals ...... 25 5.2 Proposedalgorithm...... 26

6 Trigonometric algorithm 30

7 Examples 31 7.1 Example 1: x5 + x + 0.01 = 0 ...... 32 7.1.1 Trigonometric algorithm ...... 32 7.1.2 Iteration of radicals algorithm ...... 33 7.2 Example 2: x5 + x +(3.08 + 1.68i)= 0 ...... 33 7.2.1 Trigonometric algorithm ...... 33 7.2.2 Iteration of radicals algorithm ...... 34

8 Future direction of work 35

2 Abstract

According to the Abel-Ruffini theorem [1] and Galois theory [2], there is no solution in finite radicals to the general quintic equation. This article takes a different approach and proposes a new method to solve the quintic by it- eration of radicals. But, the most intriguing result is an accurate algebraic formula for absolute and relative root approximation: 3 2 formula- root < 4.32 10− and formula/root -1 < 2.51 10− . We then | | × | | × expand some of the geometric properties discussed to construct a trigono- metric algorithm that derives all roots.

1 Introduction

Solving equations occupies a special place in the history of math- ematics. The first algebraic solution to the quadratic equation is attributed to Al-Khwarizmi [3] in the 9th century. In the 16th century, a solution to the cubic was found by Del Ferro, Tartaglia and Cardano [4] and was arguably the first significant discovery of the European Renaissance. Soon thereafter, Ferrari created a method to solve the general quartic equation [5]. The success achieved in finding formulas to lower degree equations stimulated a search to achieve the same ambition for the quintic. Despite efforts by Euler, Gauss, Lagrange and others, no general solutions were found. Finally, al- most 300 years after solving the cubic, Abel provided a complete proof of the impossibility of solving general quintic by finite combination of arithmetic operations and radicals. Subsequently, Galois went a step further and provided an exact criterion that characterizes solvable equations: their Galois group must be solvable.

Naturally, the quintic equation required more tools. The first significant de- velopment was the discovery of the Bring-Jerrard normal form x5 +ux+v =0 by using respectively quadratic and quartic Tschirnhaus tranformations [6]. In 1858, Hermite [7] used a simplified normal form to provide the first known solution to general quintic equations in terms of elliptic modular functions. Shortly afterwards, Brioschi [8] and Kronecker [9] published equivalent solu- tions. In 1860, Cockle [10] and Harley [11] developed a method for solving the quintic using differential equations, leading to a solution in hypergeo- metric functions. And in 1884, Klein [12] provided an icosahedral solution of

3 the quintic. More recently, Doyle and McMullen [13] solved the quintic using an iterative method, while Glasser [14] provided a derivation to trinomial equations that leads to the same hypergeometric solution for the quintic as Cockle and Harley.

In this article, we explore a new method that solves the quintic by itera- tion of radicals. Perhaps the most astonishing finding is the ability to find an accurate global approximation in radicals to one the five roots from the 3 first iteration: 1st iteration- root < 4.32 10− . We then proceed by proving the speed of convergence| of the proposed| iterative algorithm.

Next, we discuss Bring radicals. Recall that the Bring radical of a com- plex number a is any of the five roots of: x5 + x + a = 0. More importantly, quintic equations can be solved in closed form using radicals and Bring rad- 5 icals. One might think that a simple iterative algorithm xk+1 = √a + xk − can solve the Bring equation. However, this is not the case as we show a counter example (when a is real) where the algorithm does not converge re- gardless of the starting point’s choice outside of the real root itself. On other hand, we prove that our proposed iterative algorithm converges globally for any complex number a. Furthermore, we provide a radical approximation to 2 one of the Bring radicals: BR(z)–approximation < 2.90 10− . | | Finally, we expand some of the properties used in the proposed algorithm to provide the location of the five roots in the . This leads to a trigonometric bisection algorithm that solves general quintic equations.

2 Reduced Forms

The general quintic equation can be transformed to the Bring–Jerrard normal form: 5 v + d1v + d0 =0 From hereon, we define the n-th root of a complex number as the one with its argument in [ π/n, π/n[. −

Excluding the trivial cases of d1 = 0 and d0 = 0, a change of variable

4 4 x = v/√d1 can further simplify the equation to Form 1:

d0 x5 + x + a = 0 where a = (a =0) (1) 4 5 d1 6 p Another change of variable z = a/x leads to Form 2: z5 + z4 a4 = λ where λ = (2) 2 − 2 Let y = uz, (2) becomes: y5 + uy4 = u5λ 2 Choose u to satisfy:

1 λ λ 5 π π u5 = | | or u = | | = eiθ θ [ , [. λ λ ∈ − 5 5   This leads to Form 3: 5 4 y + uy + = ξ ξ = λ R ∗ (3) 2 | | ∈ Think of Form 3 as a simple rotation of Form 2 with the benefit of solving the equation close to the real axis.

Notice that s is solution of (3) if and only if its conjugates ¯ is a solution of y5 +¯uy4 = ξ 2 Therefore we can assume from hereon that 0 θ π/5. ≤ ≤ 3 Iteration of radicals algorithm to solve equa- tions in Form 3

3.1 Algorithm Let us consider Equation (3)

5 4 y + uy + = ξ ξ R ∗ (3) 2 ∈ 5 π u = eiθ θ 0, ∈ 5 It can be exploited in two ways. First, by factorizing:h i

2ξ y = 4 (4) su + y

Second, by completing the quintic:

u 5 2u2 2u3 u4 u5 y + =2ξ + y3 + y2 + y + (5) 5 5 25 125 3125   By substituting y in the right-hand side of (5):

y = G(ξ,y) (6) with: 2 3 4 5 5 2u 2u u u u G (ξ,y)= 2ξ + t3 + t2 + t + (7) r 5 25 125 3125 − 5 2ξ t = 4 (8) su + y This inspires the following fixed point algorithm with the following sequence: (yk)k N by: ∈ Starting point1 2 ξ 9 2+ √2 y0 = with α = (9) α s 4   Iterative process yk+1 = G (ξ,yk) (10)

To prove the convergence and the accuracy of this algorithm, we proceed with the following steps:

• Section 3.2 shows that there exists a unique root y∗ for Equation (3) near the positive real axis.

1The intuition behind α is discussed below.

6 + • Section 3.3 shows that ξ R ∗ and θ [0,π/5]: ∀ ∈ ∀ ∈ 3 y1 y∗ < 4.32 10− | − | In other words, the closed form formula:

2 3 4 5 5 2u 3 2u 2 u u u y1 = 2ξ + t + t + t + r 5 25 125 3125 − 5 with 2ξ 4 t = 2 v ξ 9 uu + α u t  provides an accurate absolute approximation of y∗ • Section 3.4 shows that the relative error of the formula is also small: ξ R+ and θ [0,π/5] ∀ ∈ ∀ ∈

y1 2 1 < 2.58 10− y − ∗

• Finally, section 3.5 shows the sequence (yk)k N converges toward y∗ ∈ and: 1 k N yk+1 y∗ < yk y∗ K 15.44 ∀ ∈ | − | K | − | ≈ 3.2 Unique root near the positive real axis Consider equation (3). When θ = 0, u = 1 and it is obvious that there is a unique real solution that satisfies (3). Now assume θ > 0. We will show in this section that there exists a unique root y of Equation (3) near the real axis: θ y = reiσ with r> 0 and σ , 0 (11) ∈ −4   The real and the imaginary parts of the equation are:

r5 cos(5σ)+ r4 cos(θ +4σ)=2ξ (12) and r5 sin (5σ)+ r4 sin (θ +4σ)=0 (13)

7 From (13): sin (θ +4σ) r = (14) − sin (5σ) Substituting r in (12):

sin4 (θ +4σ) sin (σ θ) f(σ)= − =2ξ (15) sin5 (5σ)

Let’s write: f(σ)= g1(σ)g2(σ) Where sin4 (θ +4σ) sin (σ θ) g1(σ)= and g2(σ)= − sin4 (5σ) sin (5σ) θ g1 is positive and stricly increasing in [ , 0[. Likewise g2 is positive and − 4 stricly increasing since:

sin(4σ + θ) + 4 sin(θ σ) cos(5σ) g′ (σ)= − > 0 2 sin2 (5σ)

Therefore f is continuous and strictly increasing. In addition:

θ f = 0 and lim f (σ)=+ (16) −4 σ 0− ∞   → Consequently, there exists a unique σ [ θ , 0[ satisfying (15) or a unique ∈ − 4 root y∗ such as: θ Arg(y∗) < 0 (17) − 4 ≤ 3.3 Absolute error theorem + In this section, we will prove the absolute error theorem: ξ R ∗ and θ [0,π/5]: ∀ ∈ ∀ ∈ 3 y1 y∗ < 4.32 10− | − | We will start by discussing the unique properties of the selected starting point.

8 3.3.1 Starting point properties • P.1 y∗ y0 (18) | |≤| | • P.2 u + y∗ u + y0 (19) | |≤| | • P.3 Define

4 u + y0 h0 = u + y r ∗ π Arg (h0) (20) | | ≤ 20 Proof:

• P.1: y∗ y0 | |≤| | Using (3) 8 2 2 y∗ u + y∗ =4ξ | | | | Therefore 8 2 2 y∗ 1+ y∗ +2 y∗ cos(θ σ) =4ξ | | | | | | −

2 Since 1 + y∗ 2 y∗ : | | ≥ | | 1 2 9 9 2 θ σ 2 ξ y∗ cos − ξ or y∗ 2 θ σ | | 2 ≤ | | ≤ cos −   2 ! Also:  π θ 0 θ and σ 0 ≤ ≤ 5 − 4 ≤ ≤ Which leads to: θ σ π 0 − ≤ 2 ≤ 8 and θ σ π 2+ √2 cos − cos = = α 2 ≥ 8 s 4     It follows that: 2 ξ 9 y∗ = y0 | | ≤ α | |   9 • P.2: u + y∗ u + y0 | |≤| |

Since u = 1 and the angle between u and y∗ is θ σ: | | − 2 2 u + y∗ =1+ y∗ +2 y∗ cos(θ σ) | | | | | | −

Likewise, since the angle between u and y0 is θ:

2 2 u + y0 =1+ y0 +2 y0 cos(θ) | | | | | |

From P.1: y∗ y0 . Also 0 cos(θ) cos(θ σ). Therefore: | |≤| | ≤ ≤ −

u + y∗ u + y0 | |≤| | • P.3:

4 u + y0 π h0 = Arg (h0) u + y | | ≤ 20 r ∗

Notice u + y0 =(y0 + cos(θ)) + i sin(θ). Since y0 > 0

sin(θ) sin(θ) 0 ≤ y0 + cos(θ) ≤ cos(θ)

Therefore 0 Arg (u + y0) θ (P.3.1) ≤ ≤ 4 + On the other hand, since y∗ (u + y∗)=2ξ R and Arg(y∗)= σ: ∈

Arg (u + y∗)= 4σ − Since: θ σ 0 −4 ≤ ≤

0 Arg (u + y∗) θ (P.3.2) ≤ ≤ From (P.3.1) and (P.3.2)

θ Arg (u + y0) Arg (u + y∗) θ − ≤ − ≤

10 Or u + y0 π Arg θ u + y ≤ ≤ 5  ∗ 

Since 1 u + y0 Arg(h0)= Arg 4 u + y  ∗  π Arg (h0) | | ≤ 20

3.3.2 Proof of absolute error theorem 3 δ1 = y1 y∗ < C0 where C0 = 4.32 10− (21) | − | For k N define: ∈

4 2ξ y∗ 4 u + yk tk = , hk = = , δk = yk y∗ (22) u + y t u + y | − | s k k r ∗ where (yk)k N is defined in (9) and (10) (section 3.1). ∈ Also for t C define: ∈ 2u u2 P (t)=2t3 + t2 + t (23) 5 25

Recall that y∗ = G (ξ,y∗):

2 3 4 5 5 2u 3 2u 2 u u u y∗ = 2ξ + y∗ + y∗ + y∗ + r 5 25 125 3125 − 5 Using (23), this can be simplified to:

5 2 u 5 u u y∗ + = 2ξ + + P (y∗) 5 r 3125 5 Or: u5 u 5 u2 2ξ + = y∗ + P (y∗) (24) 3125 5 − 5   Likewise y1 = G (ξ,y0) leads to:

5 2 5 2 u 5 u u 5 u u y1 + = 2ξ + + P (t0)= y + + (P (t0) P (y )) 5 3125 5 ∗ 5 5 − ∗ r r  11 Therefore: 2 u u 5 u P (t0) P (y∗) y1 + = y∗ + √1+ ǫ1 where ǫ1 = − 5 (25) 5 5 5 y + u   ∗ 5 In the following two lemmas we will show that ǫ1 < 0.049444787  | |

Lemma 1: Upper bound for ǫ1 | | Define:

1 1 2 2 1 y∗ u ω = + r = y∗ d1 = 3.82975138 2α u y | |    ∗  ! and 1 2 4 34 3 21 2 1 1 u 9 ω 9 1 6r + 5 r + 25 r + 25 r A = 2 B = 2− C = u 5 20d1α ω y + ∗ 5

ǫ1 satisfies:

ǫ1 < ABC | | Proof:

2 2 2 2 u P (t0) P (y∗)=(t0 y∗) 2t0 +2t0y∗ +2y∗ + (t0 + y∗)+ (26) − − 5 25   From P.2, u + y∗ u + y0 therefore: | |≤| |

y∗ 4 u + y0 h0 = = 1 | | t0 u + y ≥ r ∗

Or

t0 r = y∗ | | ≤ | | This combined with (26) leads to:

2 4 1 P (t0) P (y∗) t0 y∗ 6r + r + (27) | − |≤| − | 5 25   On the other hand: 4 4 t0 y∗ t0 y∗ = 3 2 − 2 3 (28) − t0 + t0 y∗ + t0y∗ + y∗

12 Since 4 2ξ 4 2ξ t0 = and y∗ = u + y0 u + y∗

4 4 2ξ y∗ y0 4 y∗ y0 t0 y∗ = − = y∗ − − u + y∗ u + y0 u + y0 It follows that:

4 y∗ y0 t0 y∗ = y∗ 3 −2 2 3 − (u + y0)(t0 + t0 y∗ + t0y∗ + y∗ ) Recall from (22):

y∗ 4 t0 = and u + y0 = h0 (u + y∗) h0 Therefore: (y∗ y0) t0 y∗ = y∗ − 2 3 4 (29) − (u + y∗) h0 + h0 + h0 + h0

Since h0 1, Arg(h0) π/20 and the modulus of a complex number is | | ≥ | | ≤ 2 always greater than its real part :

3 2 3 π h0 1+ h0 + h0 + h0 cos k d1 =3.82975138 (30) | || | ≥ 20 ≥ k=0 X   From (29):

2 y∗ y0 y0 y∗ t0 y∗ y∗ | − | = 1 | | | − |≤| |d1 u + y y − d1 u + y ∗ ∗ ∗ | | | | Therefore 2 y0 r t0 y∗ 1 (31) | − | ≤ y − d1 u + y ∗ ∗ | | Let’s now turn our attention to the denominator of ǫ1 in equation (25). Since cos(u,y∗) cos(π/4): ≥

u 2 1 √2 y∗ + r + + r (32) 5 ≥ s 25 5

2 This is a good approximation due to the benefit of working near the real axis.

13 Using (25), (27) and (31):

y0 4 4 2 1 2 1 y∗ 1 6r + 5 r + 25 r ǫ1 | − | 5 (33) | | ≤ 5d1 u + y u | ∗| y∗ + 5

Going back to equation (3):

4 9 y∗ (u + y∗)=2ξ =2αy0 2

By factorizing: 1 1 2 2 1 9 u y∗ 9 u 2 y∗ 2 + =2αy0 2 y u  ∗    !

Since 1 1 1 y 2 u 2 ω = ∗ + 2α u y    ∗  ! 1 1 2 2 9 1 9 y∗ u 1 9 2αy0 2 = u 2 y∗ 2 + =2αu 2 y∗ 2 ω u y    ∗  !

Therefore 9 2 y0 1 = u 2 ω y  ∗  y0 π Since Arg ∗ | y | ≤ 20 y0 1 2   = u 9 ω 9 (34) y∗ From (33) and (34):

1 2 u 9 ω 9 1 4 4 3 1 2 1 − 6r + 5 r + 25 r ǫ1 y∗ 5 | | ≤ 5d1 (1 + ) u u y∗ + 5

Multiplying the numerator and the denominator by 1+ u/y∗ | | 1 2 u 9 ω 9 1 4 4 3 1 2 1 − 6r + 5 r + 25 r u ǫ1 ∗ 5 1+ (35) | | ≤ 5d1 y u u 2 | y | (1 + )(1 + ∗ ) y + ∗ u y ∗ 5

14 Since 1 y u ω2 = 1+ ∗ 1+ 4α2 u y    ∗  1 2 4 4 3 1 2 9 9 1 u ω 1 6r + 5 r + 25 r u ǫ1 − 1+ (36) 2 2 u 5 | | ≤ 20d1α ω y + | y∗ | ∗ 5 u Recall that 1+ ∗ (1+1 /r), then: | y | ≤ 4 1 u 34 21 1 6r4 + r3 + r2 1+ 6r4 + r3 + r2 + r 5 25 y ≤ 5 25 25   ∗

This combined with (36) leads to:

ǫ1 ABC (37) | | ≤ With

1 2 4 34 3 21 2 1 9 9 1 u ω 1 6r + 5 r + 25 r + 25 r A = 2 B = 2− C = 5 20d1α ω y + u ∗ 5

Lemma 2: Upper bounds of A, B, C and ǫ1 • P.4 A 0.0152957 ≤ • P.5 B 0.1635792 ≤ • P.6 C 19.7616704 ≤ • P.7 ǫ1 0.049444787 ≤

15 Proof:

• P.4 A 0.0152957 ≤ Recall: 1 A = 2 20d1α Since: 2+ √2 d1 =3.829751381 α = s 4 A 0.0152957 (38) ≤ • P.5 B 0.1635792 ≤ Define 1 2 v = u 9 ω 9 = y0/y∗ + Using P.1 v 1. Also since y0 R | | ≥ ∈

β = arg(v)= arg(y∗)= σ − − Therefore 0 β θ/4 π . ≤ ≤ ≤ 20 v 1 v 2 +1 2 v cos(β) B = − = | | − | | (39) v9 v 18 s | |

This expression can be rewritten as:

m2 sin2(β) B = + with m = v cos(β) 0 s(m + cos(β))18 v 18 | | − ≥ | | (40) Leading to:

8 1 18 2 B (m 9 + cos(β)m− 9 ) + sin (β) (41) ≤ − q

16 The maximum of the right-hand side is reached when the with respect to m is zero at m = cos(β)/8. This leads to:

16 1 8 B + sin2(β) (42) ≤ 92 9 cos(β) s   As a result: B 0.1635792 (43) ≤ • P.6 C 19.7616704 ≤ Recall that: 6r4 + 34 r3 + 21 r2 + 1 r C = 5 25 25 u 5 y∗ + 5 Using (32) 4 34 3 21 2 1 6r + 5 r + 25 r + 25 r C 5 ≤ 2 1 √2 2 r + 25 + r 5 Therefore:   34 21 1 C 6I4 + I3 + I2 + I1 (44) ≤ 5 25 25 Where rk Ik = 5 k =1, 2, 3, 4 2 1 √2 2 r + 25 + r 5   or 5 − 2 10−2k 1 2k 5−2k √2 5 5 5 Ik = r + r− + r 25 5 !

By deriving with respect to r, the upper bound for Ik is reached at:

(2k 5)√2+ √50 8k2 + 40k r = − − k 100 20k − Which leads to: C 19.7616704 (45) ≤

17 • P.7 ǫ1 0.049444787 ≤ Using (37), (38), (43) and (45):

ǫ1 0.049444787 (46) | | ≤

Lemma 3: Upper bound of y1 y∗ | − | Define: 4 34 3 21 2 1 6r + 5 r + 25 r + 25 r 2 d = 4.801223053 C′ = u 4 y∗ + 5 • P.8 1 y1 y∗ A.B.C′ | − | ≤ d2 • P.9 C′ 8.288300 ≤ • P.10 y1 y∗ 0.004319271 | − | ≤ Proof:

• P.8 1 y1 y∗ A.B.C′ | − | ≤ d2 From (25): u 5 y1 y∗ = y∗ + (√1+ ǫ1 1) (47) | − | 5 −   Using the maximum of the modulus of the derivative:

5 1 (√1+ ǫ1 1) 4 ǫ1 5 | − | ≤ 5(1 ǫ1 ) | | −| | Since ǫ1 0.049444787: | | ≤ 1 1 4 where d2 =4.801223053 5 5 (1 ǫ1 ) ≤ d2 −| | 18 Then: 5 1 (√1+ ǫ1 1) ǫ1 (48) | − | ≤ d2 | | and 1 u y1 y∗ y∗ + ǫ1 (49) | − | ≤ d2 5 | |

Recall from Lemma 1 that ǫ1 ABC , therefore: | | ≤ 1 y1 y∗ A.B.C′ (50) | − | ≤ d2 With 4 34 3 21 2 1 u 6r + 5 r + 25 r + 25 r C′ = y∗ + C = 4 (51) 5 y + u ∗ 5

• P.9

C′ 8.288300 ≤ Using (32) 4 34 3 21 2 1 6r + 5 r + 25 r + 25 r C′ 2 6+ a3J3 + a2J2 + a1J1 + a0J0 ≤ 2 1 √2 ≤ r + 25 + r 5 Where   34 12√2 3 5 12√2 6 a3 = − a2 = a1 = − a0 = 5 −25 125 −625 Where 2 k − r 1 1 3 1 √2 2 2 2 Jk = 2 = r + r− + r− 2 1 √2 25 5 ! r + 25 + r 5   Since a0, a1 and a2 are negative

C′ 6+ a3J3 ≤ Using the derivative, Jk’s maximum is reached at: (k 2)√2+ √8+8k 2k2 rk = − − 40 10k − Therefore when k = 3: J3 2.288300 which leads to: ≤ C′ 8.288300 (52) ≤ 19 • P.10 y1 y∗ 0.004319271 | − | ≤ From (37), (38), (43) and (52)

y1 y∗ 0.004319271 (53) | − | ≤ 3.4 Relative error theorem Theorem 2:

y1 2 1 C1 where C1 = 2.51 10− (54) y − ≤ ∗

Proof: Using (49): u y1 y1 y∗ 1 y∗ + 5 1 = − ǫ1 (55) y∗ − y∗ ≤ d2 y∗ 

Since ǫ1 ABC: ≤ y 1 1 1 ABC′′ (56) y − ≤ d2 ∗ Where u 3 34 2 21 1 y∗ + 5 6r + 5 r + 25 r + 25 C′′ = C = (57) r u 4 y∗ + 5

Using (32) 3 34 2 21 1 6r + 5 r + 25 r + 25 29 21 5√2 1 C′′ 2 =6J3 + J2 + − J1 + ≤ 2 1 √2 5 25 2 1 √2 r + 25 + r 5 r + 25 + r 5     1 Since 25 (decreasing with maximum at 0) 2 1 √2 ≤ r + 25 + r 5 and J1 16.79664960, J2 2.144660941 and J3 0.671865984 (using the ≤ ≤ ≤ maximum values established in P.9).

C′′ 48.06772496 (58) ≤ It follows from (56): y1 1 0.02504947 (59) y − ≤ ∗

20 3.5 Speed of convergence Theorem 3: k 1 ∀ ≥ 1 C0 yk+1 y∗ < yk y∗ and yk+1 y∗ < (60) | − | K| − | | − | Kk with K = 15.44.

Proof: Recall that r = y∗ and from (22): | | y u + y 2ξ h = ∗ = 4 k and t = 4 k t u + y k u + y k r ∗ s k We will breakdown the proof into three lemmas:

Lemma 4: For k N∗ if yk y∗ C0 then for p =1, 2, 3, 4: ∈ | − | ≤ p 1 q hk 1+ q − ≤| | ≤ p p tk (1 + q)r | | ≤ with q =0.004338008 Proof: Since cos(u,y∗) 0 u + y∗ u = 1 and ≥ | |≥| |

yk y∗ − yk y∗ u + y ≤| − | ∗

For k 1 and p = 1, 2, 3, 4: using the maximum of the modulus of the ≥ p derivative for the function: x (1 + x) 4 7−→ p 4 p yk y∗ p yk y∗ yk y∗ hk 1 = 1+ − 1 | − | p | − | 1 4 | − | u + y∗ − ≤ 4 (1 yk y ) − ≤ 1 yk y∗   −| − ∗| −| − |

Which leads to:

p C0 hk 1 q =0.004338008 | − | ≤ 1 C0 ≤ − 21 Therefore p 1 q hk 1+ q (61) − ≤| | ≤ p p p Also, since tk = r hk | | | | p p tk (1 + q)r | | ≤ In particular:

2 3 4 2 3 4 hk + hk + hk + hk = 4+(hk 1)+(hk 1)+(hk 1)+(hk 1) − − − −

Therefore 2 3 4 2 3 4 hk + hk + hk + hk 4 hk 1 hk 1 hk 1 hk 1 ≥ −| − | − − − − − −

Using Lemma 4

2 3 4 hk + hk + hk + hk 4 4q (62) ≥ −

Lemma 5:

yk y∗ k 1 If yk y∗ C0 then yk+1 y∗ | − | ∀ ≥ | − | ≤ | − | ≤ K

Proof: Using exactly the same steps that led to (25):

u 5 yk+1 y∗ = y∗ + 1+ ǫk+1 1 (63) − 5 −   p  With: 2 u P (tk) P (y∗) ǫk+1 = − 5 u 5 y∗ + 5 Similarily, using the steps that led to (26) and (29), we obtain:

2 2 2 2 u2 u (y∗ yk)(2tk +2tky∗ +2y∗ + 5 (tk + y∗)+ 25 ) ǫk+1 = y∗ − (64) 5 2 3 4 u 5 (u + y∗) hk + hk + hk + hk y∗ + 5 Using Lemma 4:   1+ q 4 1 ǫk+1 6I3 + I2 + I1 C0 0.000937238 = q1 | | ≤ 5(4 4q) 5 25 ≤ −   22 Applying the maximum of the modulus of the derivative:

5 1 1 1+ ǫk+1 1 4 ǫk+1 ǫk+1 (65) | − | ≤ 5(1 q1) 5 | | ≤ d3 | | p − Where d3 =4.999128599 Using (63) and (65): 1 u yk+1 y∗ y∗ + ǫk+1 | − | ≤ d3 5  

Or 2 2 2 2 u2 1 u (y∗ yk)(2tk +2tky∗ +2y∗ + 5 (tk + y∗)+ 25 ) yk+1 y∗ y∗ − 2 3 4 u 4 | − | ≤ d3 5 (u + y ) hk + hk + hk + hk y + ∗ ∗ 5

u + y∗ 1 and by using (32), (62) and Lemma 4:   | | ≥ 3 4 2 1 1+ q 6r + 5 r + 25 r yk+1 y∗ yk+1 y∗ 2 | − |≤| − |20d3(1 q) 2 1 √2 − r + 25 + r 5   Therefore 1+ q 4 1 yk+1 y∗ yk+1 y∗ 6J3 + J2 + J1 | − |≤| − |20(1 q)d3 5 25 −   Since 1+ q 4 1 1 6J3 + J2 + J1 where K = 15.44198 20(1 q)d3 5 25 ≤ K −  

yk y∗ yk+1 y∗ | − | | − | ≤ K Lemma 6: k 1 ∀ ≥ 1 C0 yk+1 y∗ < yk y∗ and yk+1 y∗ < | − | K | − | | − | Kk with K = 15.44.

23 Proof: By induction, since y1 y∗ < C0 and K > 1 and using Lemma 5: for all | − | k 1 yk y∗

yk y∗ yk+1 y∗ | − | | − | ≤ K and C0 yk+1 y∗ | − | ≤ Kk 4 Solving equations in Form 2

Consider equations in Form 2:

z5 + z4 = λ 2 Recall from section 2 that a rotation y = uz leads to Form 3:

1 y5 + uy4 λ 5 = ξ with ξ = λ and u = | | 2 | | λ   Therefore the same algorithm would apply to Form 2:

1. Starting point

2 1 λ 9 λ 5 2+ √2 z0 = | | with α = (66) α λ s 4   | | 2. Iteration

5 2 3 2 2 1 1 1 zk+1 = 2λ + mk + mk + mk + (67) r 5 25 125 3125 − 5 Where 4 2λ m = k 1+ z r k

Since u = 1 and zk = yk/u for all k N, the properties discussed in section (3) would| | apply: ∈

24 1. Absolute error theorem:

3 z1 z∗

z1 2 1

3. Speed of convergence theorem: k 1 ∀ ≥ 1 C0 zk+1 z∗ < zk z∗ and zk+1 z∗ < | − | K | − | | − | Kk with K = 15.44.

Also, since Arg(z∗)= Arg(y∗) θ, Arg(y∗) π/20 and θ π/5: − | | ≤ | | ≤ π Arg(z∗) | | ≤ 4 5 Solving equations in Form 1

5.1 Challenge with iterative radicals Consider the equation

x5 + x + a =0 a =0.01

One might think a trivial algorithm to find the real root of this equation is the following: 5 x0 = 0 and xk+1 = √a + x − Which produces the following sequence (approximated to the fifth digit): x0 = 0, x1 = 0.3981, x2 =0.8275, x3 0.9652, x4 =0.9909, x5 = 1.0002, − − − x6 = 0.9980, x7 = 1.0016, x8 = 0.9983, x9 = 1.0017, x10 = 0.9983, − − x11 = 1.0017, x12 =0.9983, x13 = 1.0017, x14 =0.9983,... − −

Notice the algorithm does not converge! But could another starting point x0 lead to a different outcome? Outside the real-root itself, we will show below that it’s not the case.

25 Indeed assume that the sequence converges toward a x∗. Since 4 x∗(x∗ +1)= a, x∗ a. − | | ≤

Use ǫ =0.01a. There exists an integer k0 such that for all k>k0:

xk x∗ < ǫ | − | Note that: 5 x∗ = √a + x − ∗ 5 xk+1 = √a + xk − 5 Using the mean value theorem for the function √a + x between x∗ and xk, − there exist a real number c between x∗ and xk such that:

4 1 5 xk+1 x∗ = a + c − xk x∗ | − | 5 | | | − |

Since a + c a + c a + x∗ + ǫ 2a + ǫ 2.01a | |≤| | | |≤| | | | ≤ ≤

xk+1 x∗ 4.55 xk x∗ | − | ≥ | − | Which proves that the algorithm is not stable.

5.2 Proposed algorithm Consider equation (1):

x5 + x + a = 0 with a =0 6 As discussed in section 2, a change of variable z = a/x leads to:

z5 + z4 = a4 − As proven in section 4, the proposed algorithm provides an accurate first estimate z1 for one of the five roots z∗:

z1 z1 z∗

Consider x∗ = a/z∗ and the sequence xk = a/zk for k N ∈

26 Lemma 7: 1. Absolute error: 2 x1 x∗

3. Speed of convergence: k 1 ∀ ≥ 1 C2 xk+1 x∗ < xk x∗ and xk+1 x∗ < k | − | K′ | − | | − | K′

with K′ = 14.68. Proof:

1. Absolute error: 2 x1 x∗ < C2 C2 = 2.90 10− | − |

Notice: a x1 x∗ = z1 z∗ . (68) | − | z1z | − | ∗ 4 4 z∗ (1 + z∗)= a − Since, φ the argument z∗ is between [ π/4,π/4], − 2 2 1+ z∗ =1+ z∗ +2 z∗ cos(φ) 1 | | | | | | ≥ Which leads to 1+ z∗ 1 4 4 | | ≥ Since z∗ 1+ z∗ = a | | | | | | 4 4 z∗ a | | ≤| | Or r = z∗ a . (69) | |≤| | On the other hand, equation (2) implies: 5 4 4 z∗ + z∗ a . (70) | | | | ≥| | We distinguish two cases:

27 • a 1: Using (69), r 1. From (70): | | ≤ ≤ 4 4 5 4 a 1 a z∗ + z∗ 2 z∗ therefore | | 2 4 | | ≤| | | | ≤ | | z ≤ | ∗| Since z1/z∗ 1

a a z1 x1 x∗ = z1 z∗ = 1 | − | z1z | − | z1 z − ∗ ∗

Therefore: 1 4 2 C1 2 x1 x∗ 2.90 10− (72) | − | ≤ 1 C1 ≤ − 5 4 • a > 1: Using (70), z∗ > C3 = 0.856 (function z∗ + z∗ is | | | |5 4 | | | | increasing in z∗ and C3 + C3 < 1 | | Also from (70):

4 5 1 5 1 a z∗ 1+ z∗ 1+ | | ≤| | z ≤| | C3  | ∗|   Therefore 4 5 C3 z∗ a 5 (73) 1+ C3 | |≥| | r Using (71):

4 5 C3 z1 a 5 (1 C1) (74) 1+ C3 | |≥| | − r From (73) and (74):

2 5 a 3 1 C3 − | | a − 5 (1 C1)− < 1.395955878 z1z ≤| | − 1+ C3 | ∗|   2 x1 x∗ 0.005711884 2.90 10− (75) | − | ≤ ≤ 28 2. Relative error:

x1 2 1 < C1′ C1′ = 2.57 10− x∗ −

Since x∗ = a/z∗ and x1 = a/x1:

x1 z∗ z1 z∗ z1 1 = − = 1 x − z1 z1 z − ∗ ∗

Using the relative error theorem for Form 2 and (71):

z1 z∗ z1 1 1

Therefore x1 C1 2 1 <

3. Speed of convergence: 1 xk+1 x∗ < xk x∗ with K′ = 14.68 | − | K′ | − |

Since x = a/z:

a a a(z∗ zk+1) xk+1 x∗ = = − − zk+1 − z∗ zk+1z∗ and a a a(z∗ zk) xk x∗ = = − − zk − z∗ zkz∗ Which leads to:

xk+1 x∗ zk z∗ zk+1 z∗ − = − (5.1) x x z z +1 z z k ∗ ∗ k k ∗ − − First,

zk+1 z∗ − K z z ≤ k ∗ − Second,

zk zk z∗ zk z∗ z1 z∗ = 1+ − 1+ | − | 1+ | − | 1+ C1 z z ≤ z ≤ z ≤ ∗ ∗ ∗ ∗ | | | |

29 Therefore: zk 1+ C1 (5.2) z∗ ≤

Likewise

zk+1 zk z∗ zk z∗ z1 z∗ = 1+ − 1 | − | 1 | − | 1 C1 z∗ z∗ ≥ − z∗ ≥ − z∗ ≥ − | | | |

Therefore: z 1 ∗ (5.3) z +1 ≤ 1 C1 k − From inequalities (5.1), (5.2) and (5.3)

xk+1 x∗ 1 1+ C1 1 − with K′ = 14.68 x x ≤ K 1 C1 ≤ K k ∗ ′ − −

6 Trigonometric algorithm

The logic used in section 3.2 can be expanded to establish the geometric location of the five roots in the complex plane.

π Case 1: 0 <θ< 5

iσ First, recall that y∗ = re is a root of (3) if and only if two conditions are met: sin4 (θ +4σ) sin (σ θ) f(σ)= − =2ξ (76) sin5 (5σ) sin (θ +4σ) r = > 0 (77) − sin (5σ) Second, since 5 u y∗ 1+ =2ξ y  ∗  u 5σ + Arg 1+ =0 mod2π y   Let σk be the argument of the root y∗k: u 5σk + Arg 1+ =2kπ 2 k 2 y − ≤ ≤  ∗k  30 + When ξ tends to 0 , either y∗ tends to y∗ 2 = u with the argument π + θ − − − or y∗ tends to 0 with the argument for that tends to: 2kπ for k = 1, 0, 1, 2 5 − When ξ tends + , y∗ tends to + with an argument σk that tends to: ∞ | k| ∞ 2kπ θ − for 2 k 2 4 − ≤ ≤ This intuition leads to considering the following intervals: π θ 2π θ I 1 = , I0 = , 0 − − 2 − 4 − 5 −4 −     2π π θ 4π θ I1 = , I2 = , π 5 2 − 4 5 − 4     and 4π I 2 = π + θ, − − − 5   f is continuous in each of these intervals. f is zero at the lower bound of Ik for k = 2, 1, 0 and at the upper bound of Ik for k = 1, 2. f also tends to − + on the upper bound of Ik for k = 2, 1, 0 and at the lower bound Ik for ∞ − k =1, 2. Notice also that if σ Ik then (77) is satisfied (r> 0). ∈

Consequently, equation (76) has at least one solution σk in each of the five disjoint intervals. Since f is strictly monotonic in each interval, we can define a bisection method to approximate σk and the associated modulus from (77), leading to the five roots of equation (3).

π Case 2: θ =0 or θ = 5

The same algorithm described in the previous case would apply to iden- tify four roots, while the fifth can be obtained from Vieta’s formulas. When π θ = 0, we use a bisection in the intervals I 2, I 1, I1 and I2. For θ = 5 , we − − use the bisection in the intervals I 1, I0, I1 and I2. − 7 Examples

All numbers presented in this section are approximated to the 10th decimal digit.

31 7.1 Example 1: x5 + x + 0.01 = 0 7.1.1 Trigonometric algorithm Using the transformation to Form (3) π ξ =0.000000005 and θ = 5

The bisection method in four intervals leads to roots y∗k for k = 1, 0, 1, 2 of − au equation 3 and consequently to roots x∗k using the transformation x∗k = ∗ . y k The fifth root (k = 2) is obtained from Vieta’s formula. −

k y∗k x∗k -2 0.8090170025 0.5877852582i 0.0099999999 − − − -1 0.0015494319 0.0098971415i 0.704595734 + 0.7071179873i − − − 0 0.0098621565 0.0015443338i 0.7095957339 + 0.7071176748i − 1 0.0015788252 + 0.0098566936i 0.7095957339 0.7071176748i − 2 0.0098915418 + 0.0015847876i 0.704595734 0.7071179873i − − −

x∗ 1 x∗0 −

x∗ 2 ℜ −

x∗2 x∗1

Figure 1: Roots of the equation x5 + x +0.01=0

32 7.1.2 Iteration of radicals algorithm Using the algorithm described in section 3 to estimate one root (coincides with k = 0 from trigonometric algorithm):

y∗ =0.0098621565 0.0015443338i 0 − yn Iteration yn yn y∗0 ∗ 1 | − | y 0 − 4 3 1 0.0098512048 0.0015389435i 5.58 10− 2.18 10− − × 6 × 6 2 0.0098621666 0.0015443624i 2.90 10− 1.13 10− − × 8 × 9 3 0.0098621566 0.0015443337i 1.51 10− 5.90 10− − × ×

au Since x = y : x∗0 = 0.7095957339 + 0.7071176748i and the iterative esti- mates are:

xn Iteration xn xn x∗0 ∗ 1 | − | x 0 − 3 3 1 0.7106828395 + 0.707685341i 1.23 10− 1.22 10− × 6 × 6 2 0.7095928286 + 0.7071185567i 3.04 10− 3.03 10− × 9 × 9 3 0.7095957376 + 0.7071176682i 7.53 10− 7.51 10− × ×

7.2 Example 2: x5 + x +(3.08 + 1.68i)= 0 7.2.1 Trigonometric algorithm Using the transformation to Form (3) ξ 75.75327872 and θ 0.228841153 ≈ ≈ The bisection method in each of the intervals leads to roots y∗k of equation au 3 and consequently to roots x∗k using the transformation x∗k = ∗ . y k

k y∗k x∗k -2 2.4358363319 1.6419437613i 1.1834415151 0.1608289168i − − − − -1 0.6487113516 2.6125840601i 0.607389619 + 1.1531182439i − − 0 2.5580193297 0.0347499177i 1.0110954185 + 0.9265109088i − 1 0.6697215821 + 2.5335199517i 1.116784747 0.7383651957i − 2 2.4145458637 + 1.5289087467i 0.3370490315 1.1804350402i − − −

33 ℑ x∗ 1 − x∗0

ℜ x∗ 2 −

x∗1

x∗2

Figure 2: Roots of the equation x5 + x + (3.08+1.68i)=0

7.2.2 Iteration of radicals algorithm Using the algorithm described in section 3 to estimate one root (coincides with k = 0 from trigonometric algorithm):

y∗ =2.5580193297 0.0347499177i 0 −

yn Iteration yn yn y∗0 ∗ 1 | − | y 0 − 4 4 1 2.5575832547 0.0350982734i 6.27 10− 2.13 10− − × 6 × 7 2 2.5580208152 0.0347474236i 2.71 10− 9.23 10− − × 8 × 9 3 2.5580193271 0.0347499325i 1.18 10− 4.00 10− − × ×

au Since x = y , x∗0 = 1.0110954185 + 0.9265109088i and the iterative esti- mates are:

xn Iteration xn xn x∗0 ∗ 1 | − | x 0 − 4 4 1 1.0111375519 + 0.926807176i 2.99 10− 4.07 10− × 6 × 6 2 1.0110957554 + 0.9265093895i 1.56 10− 2.12 10− × 9 × 8 3 1.0110954141 + 0.9265109156i 8.09 10− 1.01 10− × ×

34 8 Future direction of work

It would be interesting to study the feasibility of expanding the proposed iteration of radicals algorithm to identify all roots by selecting appropriate starting points and the right combination of the 5th and 4th unity roots .

References

[1] Abel, Niels Henrik (1881) [1828], ”Sur la resolution alg´ebrique des ´equations”, in Sylow, Ludwig; Lie, Sophus (eds.), Œuvres Compl`etes de Niels Henrik Abel (in French), II (2nd ed.), Grøndahl & Søn, pp. 217–243

[2] Edwards, Harold M. (1984). Galois Theory. Springer-Verlag. ISBN 0-387- 90980-X

[3] Al-Khwarizmi, Muhammad ibn Musa. (circa 825). Al-kitab al-mukhtasar fi hisab al-gabr wa’lmuqabala

[4] Cardano, Gerolamo (1545). Artis Magnæ

[5] O’Connor, John J.; Robertson, Edmund F., ”Lodovico Ferrari”, MacTu- tor History of Mathematics archive, University of St Andrews.

[6] Adamchik, Victor (2003). ”Polynomial Transformations of Tschirnhaus, Bring, and Jerrard”. ACM SIGSAM Bulletin.

[7] Hermite, Charles (1858). ”Sur la r´esolution de l’´equation du cinqu`eme degr´e”. Comptes Rendus de l’Acad´emie des Sciences. XLVI (I): 508–515.

[8] Brioschi, Francesco (1858). ”Sul Metodo di Kronecker per la Risoluzione delle Equazioni di Quinto Grado”. Atti Dell’i. R. Istituto Lombardo di Scienze, Lettere ed Arti. I: 275–282.

[9] Kronecker, Leopold (1858). ”Sur la r´esolution de l’equation du cinqui`eme degr´e, extrait d’une lettre adress´ee `aM. Hermite”. Comptes Rendus de l’Acad´emie des Sciences. XLVI (I): 1150–1152.

[10] Cockle, James (1860). ”Sketch of a Theory of Transcendental Roots”. The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science.

35 [11] Harley, Robert (1862). ”On the Transcendental Solution of Algebraic Equations”. Quart. J. Pure Appl. Math. 5: 337–361

[12] Klein, Felix (1888). Lectures on the Icosahedron and the Solution of Equations of the Fifth Degree. Tr¨ubner & Co. ISBN 978-0-486-49528-6.

[13] Doyle, Peter; Curt McMullen (1989). ”Solving the quintic by iteration”. Acta Math. 163: 151–180.

[14] M.L.Glasser, (2000) “Hypergeometric functions and the trinomial equa- tion”, Journal of Computational and Applied Mathematics, Volume 118, Issues 1-2, Pages 169-173

36