<<

The Limits of a Family; of Asymptotic Solutions to The Equation

James David Nixon [email protected] June 1, 2021

Abstract

In this paper we construct a family of holomorphic functions βλ(s) which are solutions to the asymptotic tetration equation. Each βλ satisfies eβλ(s) the functional relationship β (s + 1) = ; which asymptotically λ e−λs + 1 −λs converges as log βλ(s+1) = βλ(s)+O(e ) as <(λs) → ∞. This family of asymptotic solutions is used to construct a holomorphic tetβ (s): tetβ (s) C/(−∞, −2] → C such that tetβ (s + 1) = e and tetβ :(−2, ∞) → R bijectively. Keywords: Complex Analysis; Infinite Compositions; Complex Dynamics.

2010 Mathematics Subject Classification: 30D05; 30B50; 37F10; 39B12; 39B32

1 Introduction

This paper will start with a general theorem the author has shown a multitude of times, but of which most recently appears in [5, 6]. In [5] it is shown for a specific case, and modified to a real analysis scenario a couple more times; but in [6] the general theorem is given. We’ll use this introduction to introduce the theorem, and talk a little bit about the notation.

arXiv:2104.01990v4 [math.CV] 31 May 2021 1 ∞ Theorem 1.1. Let {Hj(s, z)}j=1 be a sequence of holomorphic functions such that Hj(s, z): S ×G → G where S and G are domains in C. Suppose there exists some A ∈ G, such for all compact sets N ⊂ G, the following sum converges,

∞ ∞ X X ||Hj(s, z) − A||z∈N ,s∈S = sup |Hj(s, z) − A| < ∞ j=1 j=1 z∈N ,s∈S Then the expression,

1We’ve added a proof of this theorem in the appendix 10.

1 n H(s) = lim Ω Hj(s, z) • z = lim H1(s, H2(s, ...Hn(s, z))) n→∞ j=1 n→∞ Converges uniformly for s ∈ S and z ∈ N as n → ∞ to H, a holomorphic function in s ∈ S, constant in z.

Upon which, this theorem provides a manner of proving holomorphy of an infinite composition of holomorphic functions; and it only requires that a certain sum converges. Where here, an infinite composition is denoted,

∞ lim H1(s, H2(s, ...Hn(s, z))) = Ω Hj(s, z) • z n→∞ j=1 Much of the theory of infinite compositions, as the author has written about in [4, 5, 6, 7] depends on the behaviour of a sum which we compare with the infinite composition. In the case of this paper (as in [5, 6]), the infinite compo- sition falls into a degenerate category. This is the case that the value in z of the infinite composition will be constant. And in contrast, in the non-degenerate category (as in [4, 7]), we’d have that our infinite composition is holomorphic in z and non-constant. For this reason, we won’t speak of z at all, except to denote the manner of composition (like binding a variable to an integral, and then tossing it away afterwords). Instead, we’ll be talking about two variables, s, λ ∈ C. And discussing infinite compositions with these two variables. This will birth us a two variable holomorphic function βλ(s); which we’ll call a family of solutions to the asymptotic tetration equation. Where, for us, we’ll call a function l(s) a solution to the asymptotic tetration equation if,

log l(s + 1) − l(s) → 0 as |s| → ∞ Where we’ll mostly be concerned with |s| → ∞ while s is in a half-plane; and these things are holomorphic (or with countable singularities) unless stated otherwise. These asymptotic solutions, essentially look like tetration at infinity, but everywhere else they may not look like tetration. This allows us to talk about of these things at infinity in a nice manner. And if we are able to solve the equation log F (s + 1) = F (s) for large s, repeatedly taking logarithms allows us to extend this definition almost everywhere in C. So the idea is to take our family of asymptotic solutions βλ and construct an error term τλ which solves the tetration equation for large s. Since βλ will be a well behaved solution to the asymptotic tetration equation; this is doable.

2 The family of functions βλ We will start our foray by pulling out of a hat the sequence of functions we want to infinitely compose to get βλ. We’ll denote this sequence of functions qj(s, λ, z); where the z value will disappear in the end. Write,

2 ez q (s, λ, z) = j eλ(j−s) + 1 Where the index j ∈ N and j ≥ 1. Now, we’re going to force <(λ) > 0 and that λ(j − s) 6= (2k + 1)πi for all k ∈ Z. The first restriction will be needed for the summation, and the second restriction ensures we have no poles. We’ll call this domain of holomorphy L, in which each qj(s, λ, z): L × C → C, where (s, λ) ∈ L and z ∈ C. Observe that,

∞ ∞ X X ez |q (s, λ, z)| = | | < ∞ j eλ(j−s) + 1 j=1 j=1 But, even better than this, we have a normally converging sum. Let K ⊂ C and let U ⊂ L both be compact sets. Then,

∞ X ||qj(s, λ, z)||U,K < ∞ j=1

This should tell us that Theorem 1.1 is going to be useful, as qj satisfies all the properties of Hj in the theorem’s statement. Now, a small reminder is that Theorem 1.1 has no restriction on how many variables are involved, although it’s only stated for one variable. For clarification of this, the reader is pointed to [6]; or to the proof of Theorem 1.1 attached in the appendix. Therein, if we take,

∞ βλ(s) = Ω qj(s, λ, z) • z j=1

Then βλ(s) is holomorphic for (s, λ) ∈ L. It’s important to remember what βλ looks like though, and why we’d even want this function. We write,

∞ ez βλ(s) = Ω • z j=1 eλ(j−s) + 1 Then, if we shift the argument in s forward by 1, we get something magical.

∞ ez βλ(s + 1) = Ω • z j=1 eλ(j−s−1) + 1 ∞ ez = Ω • z We’ve shifted the index here j=0 eλ(j−s) + 1 ∞ z Ω e •z e j=1 eλ(j−s)+1 = e−λs + 1 eβλ(s) = e−λs + 1

3 And from this, we’re in a position to state that this is a family of solutions to the asymptotic tetration equation. That is,

−λs log βλ(s + 1) − βλ(s) = − log(1 + e ) Which tends to 0 exponentially as |s| → ∞ while <(λs) > 0. Now this form of the βλ family is difficult to compute, we need to compute a bunch of nested exponentials–at infinity no less, and so therefore it can be easier to make a change of variables s = log(w)/λ. The author would like to thank Sheldon Levenstein for doing this first; as the author rarely numerically evaluates, it was Sheldon’s observation that this form is much less exhausting computationally. Write,

gλ(w) = βλ(s) Then this is holomorphic when w 6= −eλj,

∞ wez gλ(w) = Ω • z j=1 eλj + w

Upon which, calculating Taylor coefficients for gλ(w) at 0 are surprisingly simple; especially by the functional equation, w g (eλw) = egλ(w) λ w + 1 Where now computing the Taylor coefficients at 0 is a relatively simple procedure; it’s inductive. This will construct a Taylor-series valid for |w| < e<λ. And to extend gλ to its maximal domain we just iterate the functional equation. We sketch the process for the curious reader who wants to numerically evaluate these functions. (k) k d gλ(w) Call gλ (0) = ak and bk = k e ; then this is given as, dw w=0

k    k−c   c  (k) X k d w d λk gλ(w) e gλ (0) = e c dwk−c w=0 w + 1 dwc w=0 c=0 k−1   c X k k−c+1 d g (w) = (k − c)!(−1) e λ c dwc w=0 c=0 k−1 X (−1)c eλka = k!(−1)k+1 b k c! c c=0 c−1 X c − 1 b = b a c d d c−d d=0

Where,

4 ∞ X wk g (w) = a for |w| < e<λ λ k k! k=0 Which is the process which generates our coefficients. Below we’ve attached some graphs of these functions. When λ = log(2); and when we’re n composi- tions deep, gλ(x) looks like Figure 1.

Figure 1: The function glog(2)(x) for n = 2, 5, 10 iterations.

Convergence is very fast in the infinite composition manner. And we can begin to see the rapid growth. This will eventually start to grow faster than x g(x) by the law g(2x) = x+1 e ; but it starts out balanced and well behaved. We can also clearly see the essential singularities beginning to form at x = −2j for j ≥ 2 and the pole at x = −2. In Figure 2, we can see this functions behaviour in the complex plane.

For the value λ = 1/2 + 3i we’ve also included a hue plot of gλ(w) in Figure 3. This is solely done for the tenth iteration–and is good enough for local values. This graph includes grid-lines, the unit disk, and a black and white checkered marker for the origin.

In Figure 4 we can see this graph zoomed out further; where we’re bound to have a numerical discrepancy because we’ve only used n = 10 iterations.

As you can see, there is a lot of chaos in these functions. They display essen- tial singularity behaviour; and don’t behave as one would expect a holomorphic

5 Figure 2: The function glog(2)(w) for n = 10 iterations.

6 Figure 3: The function gλ(w) about w = 0 for n = 10 iterations; λ = 1/2 + 3i.

7 Figure 4: The function gλ(w) about w = 0 for n = 10 iterations; λ = 1/2 + 3i; zoomed out further.

8 function to behave as w → ∞. It’s our job to regulate these functions.

Sheldon Levenstein, is again, to thank for this change of variables. This form of many of the equations the author has solved, make the solutions look like a kind of mock Schr¨oderequation. Where in Schr¨oder’s case one would solve,

Ψ(Ls) = eΨ(s) We are solving something similar, but adding a multiplicative factor to the construction. This helps tremendously at manipulating the complex dynamics of these objects. And the convergents can make a very complicated thing less so. Plus, computationally it compares to calculating w versus calculating es–and so avoids overflow errors that much better.

It’s important to also note that βλ(s) has an exponential series–from gλ(w)’s Taylor series. Which is,

∞ X ekλs β (s) = a λ k k! k=1 2πi Which is valid for <(s) < 1. Which implies that βλ(s+ λ ) = βλ(s), so that our function is periodic in the s argument. You can also see this by inspection, plugging in the value in the infinite composition. We compress all this knowledge into the existence of a family of functions which solve the asymptotic tetration equation. Theorem 2.1 (Family Of Asymptotic Tetration Functions). There exists a 2 family of functions βλ which are holomorphic on L = {(s, λ) ∈ C | <(λ) > 0, λ(j − s) 6= (2k + 1)πi, j, k ∈ Z, j ≥ 1}. These functions are expressible as, ∞ ez βλ(s) = Ω • z j=1 eλ(j−s) + 1 Satisfy the functional equation,

eβλ(s) β (s + 1) = λ e−λs + 1 And the asymptotic relationship,

−λs log(βλ(s + 1)) − βλ(s) = O(e ) As |s| → ∞, wherever <(λs) > 0. Proof. For convergence: see Theorem 1.1–see Appendix (or [6]) for proof. The functional equation is given from convergence. The asymptotics, by the func- tional equation.

In Figure 5 is attached a graph of βlog(2)(x) for x ∈ [−10, 4] for n = 100 iterations. We can clearly see the beginning of our super-exponential growth. Trying to go further out in the x variable will cause overflow errors very fast.

9 Figure 5: The function βlog(2)(x) at n = 100 iterations.

3 The exponential convergents

In this section we’ll focus on better approximating tetration using βλ(s) at infinity. This is a difficult idea to intuit, but we’re going to better understand its behaviour at infinity. The first thing we’d like to do is construct a sequence of convergents. Let’s call,

1 −λs log(βλ(s + 1)) − βλ(s) = τλ(s) = − log(1 + e ) 2 log log(βλ(s + 2)) − βλ(s) = τλ(s) 3 log log log(βλ(s + 3)) − βλ(s) = τλ(s) . . ◦n n log βλ(s + n) − βλ(s) = τλ (s)

Upon which, the asymptotic relationship,

n τλ (s + 1) n+1 −λs −λs log(1 + ) − τλ (s) = log(1 + e ) = O(e ) βλ(s + 1) Which is another side effect of being an asymptotic solution to tetration; but it’s required we have exponential convergence of τ as <(s) → ∞. To make sure everything stays well behaved in the iterated .

10 To accomplish this, we need that βλ(s) → ∞ as <(s) → ∞. This paper heavily depends on this result–and the proof is something rather unique to the exponential ez. This is our most heavily cited result from Milnor–alors, one needs to understand it to understand this result.

Theorem 3.1 (The Unbounded Theorem). The function βλ(s) → ∞ as <(s) → ∞.

z Proof. The function βλ(s) looks like the orbits of e as <(s) → ∞. The accu- mulation points of ez for z almost everywhere (under the Lebesgue measure) e are the orbit 1, e, ee, ee , ...; the orbit of the exponential at 0; which diverge to infinity. This fact is cited from Milnor [1]; but was actually shown by Lyubich [3] and Rees [8]. Milnor even makes a joke that overflow errors eventually equal 0 after a small amount of iterations −λs For our case, since β(s+1) = eβ(s)(1+e−λs)−1 = eβ(s)−log(1+e ) = eβ(s)−; e for very large <(s) we eventually converge towards the orbit 1, e, ee, ee , ...,. As these orbits are in -neighborhoods of each other at infinity–with  → 0.

s Take a neighborhood N = {z ∈ C | |z − β(s)| < 2)}; then;

β(s + 1) ∈ exp(N s) ⊃ N s+1  1

For an appropriate 1. Iterating this procedure looks like the iterated orbits exp◦n(N s) ⊃ N s+n. These must be dense in . But additionally; each starting  n C s point z ∈ N (almost everywhere) diverges to infinity as n → ∞ under the ◦n ◦n application exp (z). To see this, each zn = exp (z) gets arbitrarily close to 1–for almost all z. Then, for |z∗ − z| < δ, we must have exp◦n(z∗) = 1. The ∗ s values z cover the neighbourhood N as δ can be found as small as possible by letting n be large enough. Each point β (s + n) ∈ N s+n ⊂ exp◦n(N s). Therefore lim β (s + λ n  n→∞ λ n) → ∞ for almost all s ∈ C–excluding a measure zero set under the Lebesgue measure. The points of non-divergence are points z where exp◦n(z) forms a cycle (or a fixed point). No cycle can occur for βλ(s + n)–by its functional equation. And since all cycles and fixed points are repelling of ez; any neighborhood of a cycle diverges. Thus, the equation lim<(s)→∞ βλ(s) = ∞ is valid for all s.

n We are going to call upon the sequence of functions τλ (s), which are holo- morphic on some subset L (we’ll get to that later), and try to express them. n To begin, we know that each τλ (s) has exponential decay to 0 as <(s) → ∞. To show this, we need only look at the functional equation. Assume it follows for n and go by induction to get n + 1. Take,

n n+1 −λs τλ (s + 1) τλ (s) = − log(1 + e ) + log(1 + ) βλ(s + 1) Since we know that, by The Unbounded Theorem 3.1, 1 → 0 as <(s) → ∞ βλ(s + 1)

11 n+1 −λs −λs We know that τλ (s) must look like − log(1 + e ) + O(e ). Which cer- tainly means each τ has exponential convergence to 0. But we want something slightly stronger. To continue,

n ◦n τλ (s) = log βλ(s + n) − βλ(s) And we want to show by induction that this thing decays like −e−λs(1+o(1)). These functions satisfy the identity,

n+1 n τλ (s) = log(βλ(s + 1) + τλ (s + 1)) − βλ(s) 0 We have begun this iteration with τλ(s) = 0 and,

∞ X (−1)k τ 1(s) = e−kλs λ k k=1 2 Now the goal is, when talking about τλ(s), we can almost express it as an n exponential series; and similarly with τλ . At least, in a neighborhood of ∞. n This lets us say that τλ (s) has a removable singularity at <(s) = ∞–which will suffice to show the limit n → ∞ converges. We skip straight to the theorem.

n Theorem 3.2 (The Removable Singularity Theorem). The functions τλ (s) are holomorphic for (s, λ) ∈ L for large enough <(s) > T depending on |=(s)| ≤ T 0; and satisfy,

n −λs −λs τλ (s) = − log 1 + e + o(e )

j −λs As <(s) → ∞. And the o(e−λs) term looks like Pn−1 q e for some 0 < j=1 Pj (s) e−<λ < q < 1. Where,

j Y Pj(s) = βλ(s + k) k=1 Proof. We go by induction on n. The initial condition is trivally true because, 1 −λs when n = 1, the function τλ(s) = − log(1 + e ). Assume the result for n. Then,

 n  n+1 −λs τλ (s + 1) τλ (s) = − log(1 + e ) + log 1 + βλ(s + 1) This function is holomorphic for λ(j − s) 6= (2k + 1)πi, excluding where; n n Pn−1 j −λs β(s + 1) = −τλ (s + 1). But β(s + 1) → ∞ and τλ (s) = O( j=1 q e ). This confirms holomorphy for <(s) > T and |=(s)| ≤ T 0. So, we can expand the second term in a Taylor series;

∞ k  τ n(s + 1) X (−1)k−1 τ n(s + 1) log 1 + λ = λ βλ(s + 1) k βλ(s + 1) k=1

12 Therefore, using the bound, log(1 + w) ≤ (1 + )|w| for  → 0 as w → 0; writing (1 + )e−<λ < q < 1:

 n  n+1 −λs τλ (s + 1) τλ (s) + log(1 + e ) = log 1 + βλ(s + 1) − log(1 + e−λ(s+1)) + o(Pn−1 qje−λ(s+1))! = log 1 + j=1 βλ(s + 1) n−1 − log(1 + e−λ(s+1)) X = (1 + ) + o( qj+1e−λs)/β (s + 1) β (s + 1) λ λ j=1

And, since βλ(s) → ∞ as <(s) → ∞,

− log(1 + e−λ(s+1)) −e−λ −λs → → 0 βλ(s + 1)e βλ(s + 1)

−λs Since (1 + )e−<λ < q < 1, we can bound the first term q e . Therefore, βλ(s+1) τ n+1(s) + log(1 + e−λs) λ → 0 e−λs

The term Pj(s + 1) · βλ(s + 1) = Pj+1(s)–and so the recursion on the o term is satisfied. Thus the limit is satisfied as <(s) → ∞; and the o-term decays at j −λs least like Pn q e for some e−<λ < q < 1. j=1 Pj (s) The purpose of this o term is pretty straight forward. It allows us to bound n the function τλ (s) independent of the index n. This means, we have a normality n ∞ condition on the family of functions {τλ (s)}n=1. If we call the function,

∞ ∞ X qj X qj P (s, q) = = λ |P (s)| Qj j=1 j j=1 k=1 |βλ(s + k)| Then our sequence of functions,

n −λs −λs ||τλ (s) + log(1 + e )||K ≤ D||e Pλ(s, q)||K For some 0 < e−<λ < q < 1 and D > 1 dependent on a compact set K ⊂ L– so long as the real argument of s in K is sufficiently large. This bound is created on λ and s; and is uniform in both variables. Where as we increase the real argument of the set K we get D → 1 and q → e−<λ. We write a quick theorem, and give a quick proof.

Theorem 3.3 (Normality At Infinity Theorem). For a compact set K ⊂ L ∪ {∞}; in which <(s) > K, and ∞ is interpreted as <(s) → ∞; the sequence of n functions τλ (s) are normal on K (including the point at infinity)–and satisfy the bound:

13 n −λs −λs ||τλ (s) + log(1 + e )||K ≤ D||e Pλ(s, q)||K For D > 1 and 0 < e−<λ < q < 1. Where D → 1 and q → e−<λ as the minimum real argument of K grows.

n Proof. Each τλ (s) looks like,

n−1 X qje−λs τ n(s) = − log(1 + e−λs) + O( ) λ P (s) j=1 j For a value 0 < e−<λ < q < 1; with q → e−<λ as <(s) → ∞. The series,

∞ ∞ X qj X qj P (s, q) = = < ∞ λ |P (s)| Qj j=1 j j=1 k=1 |βλ(s + k)| And converges uniformly for (s, λ) ∈ K ⊂ L; including the point at infinity where Pλ(∞) = 0. We can add these terms up to ∞ and we must have,

n −λs −λs τλ (s) = − log(1 + e ) + O(e Pλ(s, q)) And the constant on this O term is independent of n by The Removable Singularity Theorem 3.2, and the o term we derived. Thus, we can write,

n −λs −λs ||τλ (s) + log(1 + e )||K ≤ D||e Pλ(s, q)||K For some D > 1. We must have D → 1 and q → e−<λ as the min- imum real argument of K grows; because the o-term for each n looks like e−λ(s+j) Pn−1 as <(s) → ∞. j=1 Qj k=1 βλ(s + k) 4 Choosing the proper Riemann mapping

The author would like to take the reader through a brief segue in the history of tetration. The year is 1950, and Kneser has published a treatise on iterating the : Reelle analytische Losungen der Gleichung ϕ(ϕ(x)) = ex und verwandter Funktionalgleichungen [2]. In constructing the iterate of the exponential he had a god knows how revelation. Before the dawn of computers, before the dawn of an efficient way at calculating these things, Kneser saw a solution to tetration. Now the idea isn’t so far out there now. But this idea is still what us “tetrationers” think of when we think of a nice solution to tetration. Hell, as far as most of us are concerned, it’s the nicest tetration. Hell, at this point in history, it’s still the tetration. And what was Kneser’s je ne sais quoi that flipped everything on its head?.. A Riemann mapping. Nothing more, nothing less. It was the key to his tetra- tion; a Riemann mapping.

14 The idea was simple enough; we take the inverse Schr¨oderfunction Ψ of ez about a fixed point L ∈ C. This function Ψ is entire, and satisfies,

eΨ(z) = Ψ(Lz) Where, the standard way to iterate this is to write,

F (z) = Ψ(eL(z−z0)) Upon which,

F (z + 1) = eF (z) Where, this may betray the simplicity, it looks right, it should be our solu- tion. The problem is, by construction, this function F : R 6→ R. This tetration is not real-valued. We definitely want the iteration of exponentiation to be real-valued. So, this form is pretty useless. Now, Kneser’s idea was simple, but breathtaking. Instead of talking just about a fixed point L, we also talk about its conjugate pair L. And we want to create a Riemann mapping which glues the two from above into a single entity. And this will be real-valued, so long as we remember to keep them conjugate similar. It’s difficult to find analyses of Kneser’s method, as his paper is in German, and it’s yet to be translated. However many people have re-explained his work. The most acute, I find, is in Sheldon Levenstein’s interpretation. He simpli- fies looking for the Riemann mapping by framing it as a search for a periodic function. In many ways, he refers to Kneser’s construction in a much more hands-on, crunch the , approach. To explain, if we call H = {z ∈ C | =(z) > 0}, there is a 1-periodic function θ(z): H → C such that,

∞ X 2πikz θ(z) = pke k=0

Lz tetK (z) = Ψ(e θ(z)) Where,

Lz tetK (z) = Ψ(e θ(z)) Is precisely,

tetK (z) Where the conjugate of Ψ is the inverse Schr¨oderfunction about the fixed point L, and the conjugate θ function works in the same manner here. Now, finding this function θ is very difficult, and no matter how you slice it, you need to compute a Riemann mapping. That Riemann mapping being a miraculous thing. Where firstly, Kneser finds a nice simply connected domain

15 E–where on the boundary Ψ is real-valued, and then produces the Riemann mapping to H, all while respecting the Abel functional equation, and all while making sure this will be real-valued. And so, at this point in our paper, we have to pay respect that in construct- ing a holomorphic tetration–Kneser had to pull out a very complicated mapping theorem. And so, any solution to tetration avoiding some kind of complexity similar to this, would either be miraculous or just wrong.

In our approximate solutions to tetration, in the asymptotic, they have sin- gularities. They have singularities everywhere. When we talk about,

n n Fλ (s) = βλ(s) + τλ (s) We have to account for a plethora of singularities which occur when λ(j−s) = (2k + 1)πi. These are very ugly singularities too, where if,

βλ(s) = ∞ ∞ βλ(s + 1) = e

And, if we go by inspection, we must have,

(2k + 1)πi β (1 + ) are simple poles λ λ But,

(2k + 1)πi β (j + ) are essential singularities for j ≥ 2 λ λ

So, if we have any hope in constructing a tetration function tetβ(s) which is holomorphic on s ∈ H, which is real-valued; satisfies tetβ(s) = tetβ(s); we’ll need to get rid of these singularities somehow. And, we’ll need a Riemann mapping on L to do this. Now, given, we don’t need as miraculous a Riemann√ mapping as Kneser. We just have to avoid the poles somehow. Using the · function will suffice. The way we’re going to do this is actually pretty simple. We’re going to use a function λ(s): {s ∈ C | | arg(s)| < θ, 0 < θ < π/2} → L2. Where L2 is a projection into the second component of L. This can be better explained, such that (s, λ(s)) ∈ L is satisfied for | arg(s)| < θ. We’ll require that λ : R+ → R+. And of course that λ is holomorphic. From here, we’ll define a sequence of functions,

◦n Fn(s) = log βλ(s+n)(s + n) Where we use all of the above expressions from the previous sections, to n prove that the error τλ(s+n)(s) converges as n → ∞. And then we use this to prove that Fn → F . And here, F will be holomorphic for | arg(s)| < θ.

16 But as soon as F is holomorphic for =(s) = A it is necessarily holomorphic for =(s − j) = A by the functional relationship and the non-zero nature of F . And so, this will construct a tetration function F (s): C/(−∞,X] → C which is real-valued. Hereupon we take the value x0 ∈ (X, ∞) in which F (x0) = 1 as a transfer value so that F (s + x0) = tetβ(s); which will be our desired tetration. Which is quite the mouthful; but summarizes pretty clearly our line of attack. So we write the desired theorem, 1 Theorem 4.1 (The Desired Mapping Theorem). The function λ(s) = √ , 1 + s is holomorphic for | arg(s)| < θ for some 0 < θ < π/2, and λ : R+ → R+, where the pair (s, λ(s)) ∈ L for | arg(s)| < θ. We’ll prove this in a bit. First we need to better understand the family of functions βλ(s).

5 Normality theorem at infinity; removing the singularity

When discussing the convergence of βλ(s) so far; we’ve only referred to conver- gence on compact sets. When discussing infinite compositions at infinity, the author points the reader to [4], where asymptotics were used on infinite compo- sitions. Where the general idea was, we had uniform convergence wherever the sum converges; even at infinity. This can be seen in the statement of Theorem 1.1. The set S is a set where P∞ the sum j=1 ||Hj(s, z) − A||s∈S < ∞–and the infinite composition converges uniformly for all s ∈ S. The point at infinity can be included in S if the sum converges uniformly on S. Ipso, if we call a U ⊂ L ∪ ∞ in which (s, λ) ∈ U satisfy,

∞ X 1 < ∞ ||eλ(j−s) + 1|| j=1 U

Then, βλ(s) is continuous on U, including the point at infinity (it becomes a removable singularity). In this case, conveniently, we can think of the point at infinity as <(s) → −∞ or <(s) → ∞. And so, in a half plane of L we have the identification that,

β(−∞) = 0 Because when <(s) = −∞ we have that,

∞ ∞ X ez X ez lim | | = lim | | = 0 <(s)→−∞ eλ(j−s) + 1 <(s)→−∞ eλ(j−s) + 1 j=1 j=1

17 And this can be done compactly, where if we include this point at infinity, we −λ(s−1) know that the sum converges uniformly. And additionally that βλ(s)e = 1 as <(s) → −∞. This was explored deeply in [4]; where we derived asymptotic relationships for infinite compositions. This transformation can be thought similarly to how we’ve constructed gλ(w). The value of gλ(0) = 0 is the value of βλ(s) at (s, λ) = −∞; and the value of gλ(∞) = ∞ is the values of βλ(s) at (s, λ) = ∞. This means, λs exactly like e our function βλ(s) 6→ 0 as (s, λ) → ∞, but tends to zero as (s, λ) → −∞. Since L is in two complex dimensions, we can assign two points at infinity. In such a sense, we are considering Lb = L ∪ {∞, −∞} Now, if we take the gλ(w) approach at understanding Lb, the real trouble is when we look at the real ∞ in question (when <(s) = ∞, w = ∞). This is when gλ(w) is in a neighborhood of infinity; which is when our asymptotic kicks in. In this space, we consider gλ(1/w), which we give by, yet again, an infinite composition. We’ll begin now to denote this,

∞ ez fλ(w) = gλ(1/w) = Ω • z j=1 eλjw + 1 And we’ll attach the commutative diagram,

s7→s+1 eβλ(s) βλ(s) e−λs+1

s=− log(w)/λ s=− log(w)/λ

−λ w7→e w efλ(w) fλ(w) w+1

This function is holomorphic on C× = {w ∈ C | w 6= 0, w 6= −e−λj}; but we don’t need such an expansive domain. Instead we’ll focus on the unit disk D subtracting our bad points. The set in question is,

× −λj D = {w ∈ D : 0 < |w| < 1, w 6= −e } And on this punctured disk (sort of, forgive the abuse),

efλ(w) f (e−λw) = λ w + 1 Where in this space,

−λ log fλ(e w) = fλ(w) − log(1 + w) This is again understood as a commutative diagram,

18 Figure 6: The function flog(2)(w) = glog(2)(1/w) about w = 0 for n = 10 iterations. For w ∈ Rx≤0 the function exhibits its most extreme behaviour–we can see the essential singularities pretty clearly. To the right of these (on the right side of the graph) you can see a nice convergence to infinity as w → 0.

−λ w7→e w efλ(w) fλ(w) w+1

z7→log z z7→log z

w7→e−λw log fλ(w) fλ(w) − log(1 + w)

Where here, all our discussions of convergence are linearized. We have the contraction T : p(w) 7→ p(e−λw) and a sum z 7→ z + log(1 + w); and our ex- ponential e−λs 7→ w. What we want to say, which is incredibly simple in this space, is that this asymptotic relationship is satisfied for D× ∪{0}. It isn’t much, but the holomorphy of tetration follows from this.

In Figure 6, we see a regular kind of structure. Everything looks fairly uniform in a neighborhood of zero. This function gets closer and closer to w f w ef(w) satisfying the equation f( 2 ) = e –while satisfying f( 2 ) = 1+w . Nonetheless, it looks fairly regular.

19 Figure 7: The function flog(2)(w) = glog(2)(1/w) about w = 0 for n = 10 iterations. We can see its behaviour near zero is fairly regular to the right of the singularities. As it grows super-exponentially to infinity. As we increase the depth of the iteration, this area grows to nearly encircle 0, excepting Rx≤0.

20 Figure 8: The function fλ(w) = gλ(1/w) about w = 0 for n = 10 iterations; λ = 1/2 + 3i. As chaotic as this graph looks; it’s precisely as well behaved as the previous figures.

21 In Figure 7, we get a closer look at it diverging to infinity. Where, we can expect iterated log’s to converge uniformly–because flog(2) stays very very large. In Figure 8 we can see the absolute chaos that happens near 0. Where the essential singularities begin to converge towards 0. However, since we’re only 10 iterations deep, there are only 9 essential singularities. We should expect worse in the final result–but also better because at n = ∞ (the depth of iteration) the functional equation is satisfied. Nonetheless, do not be put off by Figure 8; it is exactly as regular as Figure 6 and Figure 7. It just looks insanely more complex, because we’ve used a com- plex multiplier, rather than a real multiplier. The mapping argument we make works the exact same way in both scenarios.

Despite the chaos; the following identity is satisfied for all D× and since log(1 + w) is holomorphic on D we know that,

−λ log fλ(e w) − fλ(w) = − log(1 + w) ∼ −w |w| → 0 And when we pull back into the space L, it means that,

log βλ(s + 1) − βλ(s) = 0 when (s, λ) = +∞ And it means this uniformly. In such a sense we can assign a point at ±∞ in L in which β(−∞) = 0 and β(∞) = ∞; if we keep this with the correlation that ±∞ =

n+1 −λ n −λ uλ (w) = log(fλ(e w) + uλ(e w)) − fλ(w) n × −λj Each uλ(w) is holomorphic on δD = {0 < |w| < δ, w 6= −e } for an appropriately small δ and has a removable singularity at 0; which is a fixed point n n0 uλ(0) = 0 with multiplier uλ (0) = −1. We can relate this to The Removable Singularity Theorem 3.2; which is,

n ∗ uλ(w) = − log(1 + w) + o(w) as w → 0

22 For all w ∈ δD×–*if we take this limit appropriately. The manner to take the limit is as w → 0 7→ e−λjw as j → ∞. Which is to say, the limit must be taken as a spiral into zero, controlled by our multiplier e−λ. This translates into right translations of s → s + 1, when we make a change of variables. This observation is very important because fλ(w) has an essential singularity at 0. We want to spiral into this essential singularity in a certain way, such that our asymptotics can take over. There’s nothing we can say about alternative paths. In fact, there will always be a path where this construction fails; it’s an essential singularity; Picard’s Theorem is king. The o term, by The Removable Singularity Theorem 3.2, can be expressed as,

n−1 X qj o(w) = w B (w) j=1 j For a e−<λ < q < 1; and for the sequence of functions,

j Y −λk Bj(w) = fλ(e w) k=1 n This implies we have a normality condition on uλ(w). Which is to say, that × n + on a compact set B ⊂ D , the function ||uλ(w)||B ≤ M for some M ∈ R . So n we can expect the sequence of functions uλ to be bounded. But we have something better than that. By The Normality At Infinity Theorem 3.3 we know that,

∞ j n X wq ||u (w) + log(1 + w)||B ≤ D λ Qj −λk j=1 k=1 |fλ(e w)| B For B a compact set for w ∈ D× and <λ > 0. This can be done uniformly for both w and λ–and the value q is as it was before. This gives us a clear n n normality condition on the sequence uλ, just as we have for τλ .

Again, thinking of this as a commutative diagram is perfectly possible. We’re trying to develop the limit of this commutative diagram.

−λ n+1 n w7→e w fλ+u fλ + uλ e λ

z7→log(z) z7→log(z)

−λ n w7→e w n+1 n→∞ log fλ + uλ fλ + uλ

n→∞ z7→log(z)

w7→e−λw ϕλ

23 n Where, the limit of this commutative diagram is a map ϕλ = fλ+limn→∞ uλ = fλ + uλ; which satisfies the diagram,

−λ w7→e w ϕ ϕλ e λ

z7→log(z) z7→log(z)

w7→e−λw log ϕλ ϕλ

Now, if we consider the contraction map T on functions y holomorphic on δD× and of the form y(w) = − log(1 + w) + o(w) (where the o term is bounded in n); then,

−λ −λ −λ T y1 − T y2 = y1(e w) − y2(e w) = o(e w) Thus, looks like e−λwp(w) where p(w) → 0 as w → 0. Which implies that, on a compact set J ⊂ δD× that,

||T y1 − T y2||w∈J ≤ q||y1 − y2||w∈J For |e−λ| < q < 1; which shrinks to |e−λ| as δ → 0. So, on this space we can call T a contraction mapping. We can better understand why we need this, by looking at the formula,

 f (e−λw) + un(e−λw)  un+1(w) − un(w) = log λ λ λ λ −λ n−1 −λ fλ(e w) + uλ (e w) −λ −λ −λj Then as, u(e w)/fλ(e w) → 0 under the spiral e ; we can expect a Lipschitz constant A → 1.

n+1 n n −λ n−1 −λ |uλ (w) − uλ(w)| ≤ A|uλ(e w) − uλ (e w)|

Which is better understood,

 0  fλ(w) + w 0 log ≤ A|w − w | fλ(w) + w B With a fixed point at w, w0 = 0. The constant 1 < A < 1 +  is eventual for n 0 uλ (w) 1 |w|, |w | < δ; where  → 0 as δ → 0; because w → −1 and f(e−λkw) → 0 as k → ∞. This implies the Lipschitz constant satisfies at least A → 1. A better 1 estimate would be A = −λj = o(1)–but we have such slow decay, A → 1 fλ(e w) needs to be enough. Now if we take a compact set,

24 2 −λj 0 −<λj Bδ,δ0 = {(w, λ) ∈ C | |w| ≤ δ, |w + e | ≥ δ e , j ≥ 1, 0 < a ≤ <λ ≤ b, c ≤ =λ ≤ d} ∪ {0}

n −λ n−1 −λ n n−1 −<λ Then ||uλ(e w) − uλ (e w)||B ≤ q||uλ(w) − uλ (w)||B for e ≤ −a n e < q < 1. And therefore uλ converges uniformly as n → ∞ on the com- pact set B because 0 < Aq < 1. And this is precisely our tetration exis- tence theorem. Where upon uλ(w): B → C, in which the functional equation λ λ λ fλ(e w) + uλ(e w) = ϕλ(e w) = log ϕλ = log(fλ + uλ) extends this function almost everywhere. Now, the compact set B depends on δ and δ0, and we can shrink δ or δ0. The value δ controls our Lipschitz constant A and our contracting constant q. And the value δ0 controls how close we allow ourselves to the essential singularities at w = −e−λj. Our contracting constant q can get as close as we want to e−a; similarly with A → 1. But additionally, so long as e−<λ ≤ e−a, we can alter δ and δ0 uniformly in λ in which qA < 1. And as such, the convergence is uniform in λ as well as w, because (δ, δ0) can be chosen uniformly in λ.

Theorem 5.1 (Tetration Existence Theorem). For (s, λ) ∈ L ⊂ L there exists a holomorphic tetration function Fλ(s) such that,

Fλ(s) Fλ(s + 1) = e Where L/L is a measure zero set in C2. Proof. The correct way to observe this is a bit more tacit. This is the quick run through of everything we’ve done above. We start with the result,

f (w) + w λ  0 log 0 ≤ A|w − w | fλ(w) + w Where the smaller |w − w0| < δ is, the closer 1 < A < 1 +  shrinks to  = 0. ∗ Because 1/fλ(w) = o(1) as w → 0 –*in a spiral–this is guaranteed. And so the recursive process we’ve derived,

n uλ(w) = − log(1 + w) + o(w) 1 Because the initial convergent is uλ(w) = − log(1+w), and since the further 1 ∗ n ∗ iterates will converge to uλ(w) as w → 0 we know uλ(w)/w → −1 . Hence, A → 1 at least. For the second half of the proof; parce que,

n −λ −λ −λ uλ(e w) = − log(1 + e w) + o(e w) + n And we can bound, for some D ∈ R , |uλ(w)| ≤ w(1 + D). So, we know that the operator,

25 −λ T uλ(w) = uλ(e w) Is a contraction, subject to Banach’s theorem; with a contraction constant e−<λ < q < 1 for q dependent on a compact set. This theorem can be applied uniformly. So choose a compact set B = {(w, λ) ∈ C2 | 0 < |w| ≤ δ, |w +e−λj| ≥ δ0e−<λj, j ≥ 1, 0 < a ≤ <λ ≤ b, c ≤ =λ ≤ d} ∪ {0}, mais soi,

n −λ n−1 −λ n n−1 ||uλ(e w) − uλ (e w)||B ≤ q||uλ − uλ ||B And from this, we can choose e−a < q < 1 for δ > 0. Choose δ, δ0 such that 0 < qA = q(1 + ) < 1, and then,

n+1 n n −λ n−1 −λ ||uλ − uλ||B ≤ A||uλ(e w) − uλ (e w)||B n n−1 < qA||uλ − uλ ||B

n n Which concludes the convergence of uλ(w) = τλ (s) by Banach’s Fixed Point Theorem. The function τλ(s) = log(βλ(s + 1) + τλ(s + 1)) − βλ(s); upon which we can iterate this process to s ∈ C for all τλ(s), except for branching points. Additionally, this proof was done in a uniform manner so that we have local holomorphy in λ; giving us the theorem. To better explain the situation with λ, we suggest the reader look at John Milnor’s [1] treatment of the holomorphy of the Schr¨oderfunction in its mul- tiplier. Which is a limit process described by a multiplier, where the result is holomorphic in the multiplier; which is what we have here. Except the multiplier is written e−λ. Milnor uses a similar summation argument–attributed as Koenig’s Lineariza- tion Theorem; where the majorant for the sum is locally independent of λ. This is precisely the case here; a sum is bounded uniformly in λ–such e−<λ ≤ e−a < q < 1; and a acts as a limiter for λ. And a > 0 can be chosen with δ, δ0. Now this theorem isn’t exactly what we want. We want to take the limit n limn→∞ uλ(w) as λ depends on w. This is perfectly doable by the above analy- n sis; since convergence is uniform in λ and w. But we need uλ(w) to converge in a manner where λ → 0 as n → ∞; this is not covered by this theorem. We want a mapping; a nice enough function λ 7→ λ(s) to wash away all the problems. It has to tend to 0 like n− for  > 0 at least, and  < 1 at most.

We’d like to spend some time visualizing this theorem. It may be difficult to see what’s going on exactly. For that reason, it’s helpful to use a geometric interpretation. But sadly, as we are dealing with iterated exponentials at infinity, overflow errors abound. The author still has no way of graphing this function without overflow errors. Which in many ways, is expected to happen–as we’re pulling back a super-exponential function at ∞.

26 Figure 9: The function log(log(log(log(βlog(2)(x + 4))))); where we can see it converging to a tetration function over [−2, −1] minus a shifting argument. It’s difficult to do this better without hitting overflow arguments.

Figure 10: This is a better graph of what Flog(2) looks like using a sequential approach at defining the error term τ.

27 Figure 11: A spiral in question, for w = 1/2 + 3/4i and λ = (1 + i) log(2).

−λj The function fλ(w) has essential singularities at the points w = −e accumulating at the essential singularity at w = 0. However, if we spiral into this singularity like e−λnw; we have good control over it. Such that,

n ◦n −λn uλ(w) = log fλ(e w) − fλ(w) Will have very good control, and we have convergence. If we choose to do this in another spiral, or in another arbitrary path to 0; there’s no guarantee as to what will happen. It may explode to infinity, or shrink astronomically to 0. But, along this spiral all is good.

The attached spirals in Figure 11 and 12 essentially describe what the general shape will look like. It’s really nothing more than an odd shaped logarithmic spiral. And we want to trace along these spirals into 0. Where in such manners ◦n −λn log fλ(e w) − fλ(w) → uλ(w) as n → ∞. This works for two reasons. The first being that the operator T p(w) = −λn −λ u(e w) p(e w) is a contraction mapping. The second being, e−λnw → −1 as n → ∞–which is covered by The Removable Singularity Theorem 3.2. We can think of the spiral as a right-translation in our change of variables. Additionally, along this spiral, our function fλ(w) satisfies a natural func- tional equation,

efλ(w) f (e−λw) = λ 1 + w

Where, for other spirals–we are clueless as to how our function fλ behaves. All of this surgery means–along this spiral, we can artfully remove the singularity at 0. And at 0; our tetration functional equation gets closer and closer to being

28 Figure 12: A spiral in question, for w = −1/3 + 1/5i and λ = 1/2 + 2i. satisfied. So long as we pay attention to the other singularities which happen along the spiral w = −e−λn; and we make sure to stay somewhat away from these. Which is precisely what our compact set B does. If we write,

2 −λj 0 −<λj Bδ,δ0 = {(w, λ) ∈ C | |w| ≤ δ, |w + e | ≥ δ e , j ≥ 1, 0 < a ≤ <λ ≤ b, c ≤ =λ ≤ d} ∪ {0}

Then B is chosen so that (e−λw, λ) ∈ B if (w, λ) ∈ B. And so our spiral sits well in this set. We can visualize this in w as a compact disk about 0 of radius δ–and each point w = −e−λj has a small disk about it (with a radius δ0e−<λj); in which these disks are excluded from the compact disk about 0. So we can think of this as a disk with a bunch of smaller and smaller holes spiraling into 0. Where the hole at 0 is a removable singularity. Figure 13 displays the general shape.

29 0 Figure 13: The set Bδ,δ0 for δ = 0.3, δ = 0.2, λ = log(2) − i. The shaded area is the domain B. The singularities of fλ(w) occur in the center of each disk.

30 6 Finding an appropriate mapping; and gener- alizing Theorem 5.1

It’s helpful to look at this problem from two different angles. The first being a × mapping λ(w): D → <(λ) > 0 and the second being a mapping λ : L1 → L2. × λj Recall that D = {w ∈ C | 0 < |w| < 1, w 6= −e , j ≥ 1} and L = L1 × L2 = {(s, λ) ∈ C2 | <(λ) > 0, λ(j − s) 6= (2k + 1)πi, j ≥ 1, k ∈ Z}. In the second form, we want a mapping λ : {s ∈ C | | arg(s)| < θ < π/2} → + + L2 in which λ : R → R . We want to convert this into a restriction on λ(w). The answer to this riddle isn’t too difficult. To find an appropriate mapping, we just want a mapping that expands the lines where,

λ(j − s) = (2k + 1)πi √ This can√ be done with the function s, and many functions like this will work, but s is simple enough. To visualize this, when λ is constant we have a lattice of points in the right half plane. And if we were to multiply that lattice by √ 1 we’d be able to place a sector | arg(s)| < θ within it. Or rather, 1+s j − s √ = (2k + 1)πi 1 + s Then, √ s = j + 1 + s2(k + 1)πi And there’s a sector | arg(s)| < θ < π/2 in which this can’t be true for j ≥ 1. So with this we’re going to consider the function,

∞ ez β(s) = Ω j−s • z j=1 √ e 1+s + 1 And it’s alternate form on D×, ∞ ez f(w) = Ω j • z j=1 √ e 1+s w + 1 But,

w = e−λs s = − log(w)/λ log(w) s = −√ 1 + s

So that h(w) = √ 1 for some holomorphic function h, 1+s

31 ∞ ez f(w) = Ω • z j=1 eh(w)jw + 1 n Now, this expression looks very cryptic. But we know that uλ(w) as n → ∞ converges uniformly for (w, λ) ∈ Bδ,δ0 –where this is a compact set. When we translate this back into β (s); where λ = √ 1 we get that, first of all β(s): λ 1+s {| arg(s)| < θ < π/2} → C; and secondly that,

τ n(s) = log◦n β(s + n) − β(s) → 0 Should converge uniformly for | arg(s)| < θ < π/2 as |s| → ∞ at least like O(|s|−1/2). As such we still have our identity log β(∞) − β(∞) = 0. Where this again equates to a compact set in Lb = L ∪ {−∞, ∞}. −λn If we think of our spirals wn = e w; we are modifying the spirals with the equivalence λ = h(w); and letting λ → 0 as wn → 0. This must be done very carefully, but can be done using nothing more than Banach’s Fixed Point Theorem. As we have control over the multiplier, it’s very possible.

This will imply we have a tetration function F which satisfies,

F (s + 1) = eF (s) For all | arg(s)| < θ. But then, for every s ∈ C there exists some n such that s + n is in this sector, therefore we can undo this by taking n logarithms. This will define our tetration function almost everywhere–excluding logarithmic branch-cuts/singularities. Since it will be real valued, we know that there is a real value x0 ∈ R such that,

tetβ(s) = F (s + x0):(−2, ∞) → R bijectively tetβ(0) = 1

This gives us a tetration function which definitely has singularities at the negative integers and isn’t holomorphic on the line (−∞, −2]. But there may be other singularities which appear elsewhere, it is our job to show this doesn’t happen; which we’ll do in the next section.

n Before the dramatic conclusion, we have to take a closer look at τλ . As n you’ll note, for a varying λ, τλ = limn→∞ τλ is holomorphic; but in the case of interest, we have a sequence of λn → 0 as n → ∞ and 0 is on the boundary −1/2 of our domain L. So we have to use that λn = O(n ) somewhere in our construction, and show that the limit is still holomorphic in s. This is a tricky job. And of it we’ll show in the following theorem.

Theorem 6.1 (The Pasted Together Theorem). The function Fn → F is holo- morphic on | arg(s)| < θ; and convergence is uniform on compact sets of this.

32 Proof. The goal of this proof is to control the Lipschitz constant A and the contraction constant q from The Tetration Existence Theorem 5.1. Now, when we shrink δ from our compact set we know that q → e−<λ and that A → 1, which ensures convergence. In the current situation, we know that <(λ) → 0 −1/2 Qn −<λ Pn −1/2 like O(n ). This implies that j=1 e → 0 because j=1 O(j ) → ∞ as n → ∞, and this will be −∞ because <λ > 0. This implies the product of 1/2 our contractions looks like e−Bn for some B > 0 which is enough to ensure convergence of the final result if we can keep q → e−<λ well enough, which we can because δ, δ0 can be chosen uniformly for λ. But in order for this work, we need A → 1 in a fast enough manner to not −1/2 affect this convergence. We need that A → 1 like eµn for 0 < µ < 1. Now we know we can choose A such that qA < 1, and this choice can be done uniformly. This implies that A ≤ eµ<λ for 0 ≤ µ < 1. Choose an appropriate δ and δ0 such in,

n 2 −λj 0 −<λj Bδ,δ0 = {(w, λ) ∈ C | |w| ≤ δ, |w + e | ≥ δ e , j ≥ 1, 0 < an−1/2 ≤ <λ ≤ b, c ≤ =λ ≤ d}

Our constants qnAn < 1 and qnAn → 1 as n → ∞. Note, that the compact set depends on n. The set,

n 2 −λj 0 −<λj B → U = {(w, λ) ∈ C | |w| ≤ δ, |w + e | ≥ δ e , j ≥ 1, 0 ≤ <λ ≤ b, c ≤ =λ ≤ d}

But by the time we hit U, we’ll be at u1 − u0 = log(1 + w) in the inverse iteration; which has a finite value. Now, the value λ depends on w; and so we can project this into a set,

−λj 0 −<λj B = {w ∈ C | |w| ≤ δ, |w + e | ≥ δ e , j ≥ 1} And then, to work through this, we perform the iteration,

n+1 n n n−1 ||uλ − uλ||B ≤ Anqn||uλ − uλ ||B n−1 n−2 ≤ AnqnAn−1qn−1||uλ − uλ ||B . . n Y 1 0 ≤ Ajqj||uλ − uλ||B j=1

And here,

33 −(1−µ)O(n−1/2) Anqn = e And we are finding such sequences, that,

n Y −(1−µ) Pn O(j−1/2) −O(n1/2) −Dn1/2 Ajqj = e j=1 = e = e j=1 For a constant D = (1 − µ)B > 0. Therefore by induction,

n+1 n −Dn1/2 −Dn1/2 ||uλ − uλ||B ≤ e || log(1 + w)||B = Me 1 0 Wherefore, − log(1+w) = uλ −uλ. For some constant M = || log(1+w)||B = log(1 − δ). Now, for all  > 0 there exists N such for n, m > N it’s known,

m−1 X 1/2 e−Dj < /M j=n And therefore,

m−1 m n X k+1 k ||uλ − uλ||B ≤ ||uλ − uλ||B <  k=n Which concludes the proof.

We’ve written this theorem very quickly because we’ve built The Tetration Existence Theorem 5.1 to allow for a quick generalization. All that’s needed to go from that theorem to this theorem is control over the Lipschitz constant A and the contraction constant q–which we can. This argument works exactly the same if λ = O(n−) for 1 >  > 0. Further, Banach’s fixed point theorem will ensure that these two processes will both converge to the same F . In many ways, if we take our function,

◦n Fλ(s) = lim log βλ(s + n) n→∞ From The Tetration Existence Theorem 5.1; then the function F we’ve con- structed in The Pasted Together Theorem 6.1, is essentially Fλ(s) . This is λ=0 to mean, it’s the boundary value of our function Fλ. However, we have to take this limit in a specific manner–not just plugging in λ = 0. This is to imply, our constructed function is,

◦n − F (s) = lim lim log βλ(s + n) where λ = O(n ) for 0 <  < 1 n→∞ λ→0 Where, this is the gist of the construction, but obviously requires much more depth to be valid.

34 7 tetβ is non-zero in the upper half-plane

This section is devoted to showing that tetβ(s) is non-zero for s ∈ H = {s ∈ C | =(s) > 0}. This can be equivalently said, that the only zero of tetβ(s) is at s = −1. This equates to there being no singularities for =(s) > 0. The only way a singularity arises is if log tetβ(s0+1) is singular; which implies tetβ(s0 +1) = ∞, 0. Where we know for large enough N for all n > N the values tetβ(s0 + n) are non-singular. Therein, the only thing that can start this chain of singularities is if tetβ(s0 +k) = 0 for some k. Upon which tetβ(s0 +k−j) = ∞ for all j ≥ 1. So we need a theorem that tetβ is non-zero in the upper-half plane, and we’ve simultaneously showed that tetβ is holomorphic in the upper half-plane. By conjugation, we’ll know that tetβ is holomorphic in the lower half-plane. And then tetβ : C/(−∞, −2] → C will be our maximal domain of holomorphy for tetβ. To do this, we need to understand what happens when log(tetβ(s0)) = 0. It can only happen if tetβ(s0) = 1, but it doesn’t necessarily happen if tetβ(s0) = 1. There exists a curve C in a neighborhood of s0 in which tetβ(C) ∈ R. Now, supposing that C = s0 + t for t ∈ (−δ, δ) then when we continue to iterate this procedure, necessarily tetβ(s0 + t) will be real-valued as t grows. This will force tetβ(s0 − 1) = 0. Assuming that C is not a line, then the line tetβ(s0 + t) is not real-valued for t ∈ (−δ, δ) excepting at t = 0. This means the logarithm log(tetβ(s0 +t)) must be complex valued, and this means that tetβ(s0 −1) must be in a neighborhood of 2πik for k 6= 0, and it is non-zero. So all we have to do is focus our attention on tetβ(s0 + t) for t ∈ (−δ, δ) and show that it cannot be real-valued.

Lemma 7.1 (The Non-real Lemma). For all s0 ∈ H such that tetβ(s0) ∈ R there exists δ > 0 such for t ∈ (−δ, δ) and t 6= 0 the values tetβ(s0 + t) 6∈ R. Proof. Take our asymptotic solution to tetration,

∞ ez β(s) = Ω j−s • z j=1 √ e 1+s + 1

And note there are no lines s0 + t for t ∈ R in which this is real-valued, excepting when s0 ∈ R. Therefore, the limiting process,

◦n Fn(s) = log β(s + n)

Cannot be real-valued on a line s0 + t, unless s0 ∈ R; because this looks like β(s0 + t) for large t. Therefore the result. To justify the following theorem further we need only to add a small point. Let y(z) be a holomorphic function such that y(1) = 1 and y(0) = 0. If we take the principal branch of the logarithm log : C/Rx≤0 → C, and we know log y(z) + has a branch cut along z ∈ Rx≤0, then necessarily y(x) ∈ R for x ∈ (0, 1). Which is nothing more than a mapping theorem on the log function.

35 Lemma 7.2. Suppose y(z): D → C is a holomorphic function with y(0) = 0, and y(1−) = 1. Suppose, using the principal branch of the logarithm,

log(y(z)) : D/(−1, 0] → C Then,

y : [0, 1] → [0, 1]

Proof. Immediately, we must see that y :(−1, 0) → −R+; as this is where the branch-cut is located. Therefore y : (0, 1) → (0, 1); because once a holomorphic function is real-valued it will still be real-valued. With this, we state the following theorem, which will only require a quick justification.

Theorem 7.3 (The Non-zero Theorem). The tetration function tetβ(s) 6= 0 for =(s) > 0.

Proof. By The Non-real Lemma 7.1, we know for any point =(s0) > 0 and tetβ(s0) ∈ R, that tetβ(s0 + t) is not real-valued for t ∈ (−δ, δ). Choose an s0 in which tetβ(s0) = 1. The goal is to show that tetβ(s0 − 1) = 2πik for some k 6= 0. We go by contradiction. Assume that tetβ(s0 − 1) = 0, then tetβ(s0 − 2) is a singularity with a branch-point. This branching process can be done along the + line s0 − 2 − t for t ∈ R . Therefore tetβ(s + s0 − 2) is holomorphic for |s| < δ and s 6∈ (−δ, 0]. And further y(s) = tetβ(s + s0 − 1) : D → C; where y(1) = 1 and y(0) = 0. And, we’ve used the principal branch of the logarithm to define log y(s) = tetβ(s + s0 − 2). Therefore y(s) : [0, 1] → [0, 1]. Therefore the function tetβ(s + s0 − 1) must be real-valued for s ∈ (0, δ), contradicting The Non-real Lemma 7.1. And with this we have constructed a tetration function holomorphic in the upper-half plane and real-valued analytic on the real-line. Therefore,

Theorem 7.4 (The Tetration Theorem). The function tetβ is holomorphic on C/(−∞, −2], satisfies tetβ(0) = 1 and,

tetβ (s) tetβ(s + 1) = e

And is given by the equation, for some x0 ∈ R,

◦n tetβ(s) = lim log β(s + x0 + n) n→∞ Where,

∞ ez β(s) = Ω j−s • z j=1 √ e 1+s + 1

36 8 Coding and Graphing

We’re going to spend a fair amount of time discussing manners of programming these various tetration functions; and attaching graphs based on the aforemen- tioned coding method. The author has no efficient way of constructing the Taylor series of any of the tetrations. And so, the coding is largely done by bruteforce. We’re going to write the first code dump in Matlab, but the theory extends to similar languages. The second code dump will be more difficult, and is a hard-coded method for Pari-GP. This will produce much more accurate results, and will graph much more efficiently. Both code dumps are still in beta stages, but produce enough evidence to concur with the results of this paper

To begin, we call our function beta(s,l,n)=

n z n e βλ (s) = Ω • z j=1 eλ(j−s) + 1 z=0 Where n can be thought of as the depth of iteration, and l is the variable n λ. All this is, is a for-loop. We can, somewhat, think of βλ → βλ, as a for-loop where the depth of iteration is infinite. In that we’ll write,

function f = beta(s,l,n) f=0; for i = 0:n-1 f = exp(f)./(1+exp(l*(n-i-s))); end end

This is a very naive way of constructing the function βλ, but it works well enough. This is the equivalent of an alternative method; which is to use the Taylor series of gλ(w) = βλ(s) where s 7→ log(w)/λ. Where one would compute the Taylor coefficients of gλ(w) about w = 0 and undo the substitution s 7→ log(w)/λ. However, the naive way, is much simpler and much more generalizable. It’s what we’ve used for all the graphs we’ve calculated here. It also more clearly displays the simple recursive nature of βλ. Which, in many ways, is a self- referential for-loop iterated to infinity; n → ∞. Now, when defining τ as code; the naive way is to fix an index n in βλ, and create an index k in τ; for its iteration. Mathematically speaking, this would be,

n,k  n n,k−1  n τλ (s) = log βλ (s + 1) − τλ (s + 1) − βλ (s) This can be written as the code:

function f = tau(s,l,n,k) if k == 1

37 Figure 14: y=beta(s,log(2),10) + tau(s,log(2),10,5); and we graphed abs(y) over the box |

f = log(beta(s+1,l,n)) - beta(s,l,n); return end

f = log(beta(s+1,l,n) + tau(s+1,l,n,k-1)) - beta(s,n); end

But, it equates, no less, to the function,

n,k ◦k n n τλ = log βλ (s + k) − βλ (s) Which, as you may suspect; the function,

n βλ (s + k) → ∞ as k → ∞ However, this tends to infinity too fast. If you try and graph this for large iterations, too many overflow errors happen in the circuit that everything over flows. Figure 14 and 15 display clear anomalies where we’ve overflowed some- where in the process. But also, they display the uniformity away from the singularities; where all is good. The points of divergence in the program do not equate to divergence in the math. The points of divergence in the program are just overflow errors. We get a short circuit.

The answer of which is to iterate while you iterate. We want to talk about the sequence of functions:

tau_k(s,l,k) = tau(s,l,k,k)

This can be written as the code,

function f = tau_k(s,l,k) if k == 1 f = log(beta(s+1,l,2)) - beta(s,l,2);

38 Figure 15: y=beta(s,0.5+1i,10) + tau(s,0.5+1i,10,5); and we graphed abs(y) over the box |

Figure 16: A faster converging view of tetration for λ = 1/2 + i. This is done with the function beta(z,0.5+1i,10) + tau(z,0.5+1i,6), over the area |<(s)|, |=(s)| ≤ 1.

return end

f = log(beta(s+1,l,k) + tau(s+1,l,k,k-1)) - beta(s,l,k); end

Of which we can see a larger more rapid area of convergence in Figure 16. In the same breath, a more rapid divergence as well. We can continue to massage these things, by alternating method of constructing the iteration. These fairly accurately construct Fλ(s); but it says nothing of the actual tetration we want.

For this, we’ll need to introduce the β function from The Pasted Together Theorem 6.1. With that, call beta2(s,n)=

39 Figure 17: The same function, with an increased real argument. Again, we can observe a short circuit. This time it occurs because our function is beginning to grow super-exponentially.

n ez n β (s) = Ω j−s • z j=1 √ e 1+s + 1 z=0 This can be programmed as,

function f = beta2(s,n) f=0; for i = 0:n-1 f = exp(f)./(1+exp((n-i-s)./sqrt(1+s))); end end

Now coding a tau2(s,n,k) function is pretty straight forward.

function f = tau2(s,n,k) if k == 1 f = log(beta2(s+1,n)) - beta2(s,n); return end

f = log(beta2(s+1,n) + tau2(s+1,n,k-1)) - beta2(s,n); end

Trying to graph this will immediately produce overflow errors on the real- line, as we go out; as this grows far too large. But it looks leveled in the complex plane. Now, convergence on the real-line is rather trivial, and it is more important that this object converges in the complex plane. It looks somewhat like Figure 19.

40 Figure 18: Here is a graph of the function β(s)–initialized as beta2(s,100) for 0 ≤ |<(s)| ≤ 2, |=(s)| ≤ 1. We can see its very leveled growth to infinity pretty clearly.

Figure 19: Here is almost our tetration function over |=(s)|, |<(s)| ≤ 1; where we haven’t yet shifted to the right domain. It begins to over flow as you increase the real argument, as the iteration takes too many large values of β; and we short-circuit. The ridge is the branch cut at (−∞, −2); and the edge of the graph dips to 0 before shooting off to infinity–causing all the short-circuits.

41 Figure 20: If we clean up the code of our previous iterations; and use the function TET; then this is what you get. This is precisely what our tetration should look like.

Now, these views are of the iterative procedure. We can clean this code up a lot, by using a step function approach. With that, we attach,

function f = TET(z) if (-1

f = exp(TET(z-1)); end

And the graph in Figure 20 looks much more accurate as to what our tetra- tion looks like,

Now onto the more difficult problem of programming in Pari-GP. The author will simply reference his GitHub repository for Tetration using Pari-GP. The functions are fairly similar, but slightly more difficult to use. The code dump can be found here https://github.com/JmsNxn92/Recursive_Tetration_PARI. Instead, we’ll begin by attaching a bunch of graphs which are constructed us- ing these functions. These graphs are created using Graphing.gp in the GitHub Repository; which was not coded by me, but by the user mike3 on The Tetra- tion Forum [9]. The following Figures 21 22 23 24 are all various graphs of Fλ(s).

And attached further are various graphs for the function tetβ.

42 Figure 21: The function Flog(2)(s) for −1 ≤ <(s) ≤ 3.5 and −2.5 ≤ =(s) ≤ 2.5.

Figure 22: The function F1+0.1i(s) for −1 ≤ <(s) ≤ 6 and 3.2 ≤ =(s) ≤ 9.

43 Figure 23: The function Flog(2)−i(s) for 0 ≤ <(s) ≤ 2 and −1 ≤ =(s) ≤ 1.

Figure 24: The function F1+i(s) for 0 ≤ <(s) ≤ 3 and −1.5 ≤ =(s) ≤ 1.5.

44 Figure 25: This is the Taylor series of tetβ(z) about z = 1 for 100 terms and 100 precision.

Figure 26: The function tetβ for −1 ≤ <(z) ≤ 4 and |=(z)| ≤ 0.8.

45 9 Additional properties of tetβ

In this section we’ll list some properties of tetβ which are extra to the general theory. These are nice things we can say about our solution. These are largely properties which are inherited from the family βλ. Underlining how they are inherited is the important part. We start with our function β(s) which we write as,

∞ ez β(s) = Ω j−s • z j=1 √ e 1+s + 1 In which,

log β(s + 1) − β(s) = O(|s|−1/2)

To intuitively see this, observe if we have a sequence of functions hj(s) = 1 + O(s−1/2), and we compose them to get the function,

∞ z H(s) = Ω hj(s)e • z j=1 Then,

H(s + 1) = h (s + 1)e... eH(s) 1

In our particular case, we know the exponent disappears, and h1(s + 1) = 1+O(|s|−1/2). Therefore the logarithm will behave as expected. This is a quick asymptotic, that the author hasn’t used, and doesn’t really plan on using; but it describes the shape of these things. Continuing this thread of discourse,

−1/2 tetβ(s) − β(s + x0) = O(|s| ) as |s| → ∞ while | arg(s)| < θ And so we can expect this with higher-order too, as the conver- gence is uniform for |s| → ∞ while | arg(s)| < θ. Now differentiating β is fairly easy; especially if we view this as a manner of computing Taylor coefficients of g(w). We’ll skip a few steps here, but the algebraic relationship as <(s) → ∞,

√−s β0(s)eβ(s) eβ(s)e 1+s  d −s  β0(s + 1) ∼ + √ √−s −s 2 1+s  √  ds 1 + s e + 1 e 1+s + 1

Means that,

j−s ∞ √ zeβ(s−j) eβ(s−j)e 1+s  d j − s  β0(s) ∼ + √ • z Ω j−s j−s 2 j=1 √  √  ds 1 + s e 1+s + 1 e 1+s + 1

Which is eventually non-zero for large enough s. As such, we can expect that,

46 0 tetβ(s) 6= 0 for |s| > R 0 0 And hereupon, since tetβ(s) 6= 0 we can derive that tetβ(s − 1) 6= 0. Where,

0 0 tetβ(s) tetβ(s − 1) = 6= 0 tetβ(s) 0 Which implies that tetβ(s) 6= 0 everywhere tetβ(s) 6= 0 which excludes the point −1 but at −1 we know that tetβ(−2) is a singularity, so this isn’t a problem. Which is something really advantangeous to know, but to most studies 0 of tetration, is rather apparent. This implies, yet again, that tetβ(x) > 0 for all x ∈ (−2, ∞). This means that tetβ is a bijection of (−2, ∞) → R. We can have a similar discussion with higher-order derivatives as well. It doesn’t work out as nice, but is still worth while. Now, again we know that higher order derivatives of β are also non-zero. As such, we get that,

(n) tetβ (s) 6= 0 for |s| > Rn This matters particularly for the real-line. Where it says that for large (n) enough Xn, then for all x > Xn we know that tetβ (x) > 0. This tells us, eventually, each of our derivatives will be monotone. This is a slightly weaker criterion than all of its derivatives being monotone.

We’d like to take a quick moment to discuss the inverse function slogβ of tetβ. We specifically refer to the slogβ function which takes R → (−2, ∞) bijectively. This function will be analytic by the implicit function theorem. Take N a neighborhood of zero in which slogβ is holomorphic. Then,

z slogβ(e ) = slogβ(z) + 1

This allows us to analytically continue slogβ to the set S in which,

∞ [ S = exp◦n(N ) n=0 Now, the orbits of the exponential map on an arbitrary neighborhood are dense in the complex plane. Which is the equivalent statement that the Julia set of exp is all of C (Again, we cite [1, 3, 8]). This amounts to S = C. As such, we know that slogβ is holmorphic almost everywhere in C; upto a measure zero set in C. This gives us a clear language that,

◦s exp (z) = tetβ(s + slogβ(z)) Is holomorphic on a domain P in which (s, z) ∈ P and C2/P is a measure-zero set in C2. This function satisfies the functional equation,

0 0 exp◦s(exp◦s (z)) = exp◦s+s (z)

47 For appropriately chosen s and s0. This constructs what we’d think of as an appropriate fractional iteration of exponentiation; which satisfies the exponent law and takes real-values to real-values. Upon which the identity value z 7→ z is given at s = 0 and tetβ(s) given at z = 1. If we fix z; this produces a holomorphic function in s excepting branch cuts; and vice versa. Where the restriction s, z ∈ R+ implies exp◦s(z) ∈ R+. This produces a family of functions ripe to construct ...

10 In Conclusion

To conclude this paper we broach the idea of doing this for more exotic functions. We ask if for other transcendental functions h(z): C → C, the asymptotic approach works to construct a super-function H(z) such that h(H(z)) = H(z + 1); so H satisfies the inverse Abel equation. Constructing an arbitrary function,

∞ h(z) ρλ(s) = Ω • z j=1 eλ(j−s) + 1 Which satisfies,

h(ρ (s)) ρ (s + 1) = λ λ e−λs + 1 is not a difficult task, if the domains of h are well behaved. But pulling back with iterates h◦−n is a very careful procedure. Upon which, we were lucky with ez because log is a well behaved inverse. And despite the rapid growth of ez, we were able to do this; where rapid growth is actually very beneficial. In essence, this method is more effective for rapid growing functions than it is for slowly growing functions. Least of all, with these functions ρλ we can describe asymptotically what H should look like. Wherein, the equation,

−1 −λs h (ρλ(s + 1)) − ρλ(s) = O(e ) as |s| → ∞ <(λs) > 0 Is certainly viable (so long as we have a decently well behaved function h−1 at ∞). But without a decently behaved inverse h−1, the most we’d be able to say is that ρλ is a solution to the asymptotic inverse Abel equation–expressing the same thing but in a more implicit manner. The equation above being the frank way. The author foresees no problem in utilizing this asymptotic method for h(z) = bz for b > e1/e, where the iterates of h are unbounded here. He imagines this would follow little differently than the case for b = e; subtracting minor de- tails. Specifically, we would need a proof that ρλ(s) → ∞ as <(s) → ∞; in the ωz same way we had βλ(s) → ∞. The complex plane e for ω ∈ C is a different story though–it may be tractable, as long as its iterates are unbounded; though the complexity of the logarithms sounds like a serious headache. The author also knows no way of understanding the dynamics of tetβ(s − n) for =(s) > 0. This equates to the repeated application of the logarithm; for

48 varying branches of log. The author is somewhat convinced this tetration tetβ 6= tetK , Kneser’s tetration. Where in this regard, he expects the iterated log’s on tetβ(s) may converge to varying fixed points, or diverge like the Julia set of the log map; and lim|s|→∞ tetK (s) = L for π/2 ≤ arg(s) < π. This is to say, Kneser’s tetration is normal in the upper left half plane; the function tetβ is not. With this, I conjecture that limn→∞ tetβ(s − n) → Ls, ∞; where Ls is a L fixed point e = L. And ∞ means that tetβ(s) ∈ J for J the Julia set of log; upon which repeated applications don’t converge. Infer, we interpret ∞ as non-normality, and Ls as normality, and convergence towards a fixed point. I, further, do not expect this solution to be Kneser’s tetration because the behaviour as =(s) = t → ∞ of tetβ should be ∞; as it should look like − √ it β(it) + O(e 1+it ), which β(it) should tend to infinity (again the author isn’t certain here, it just looks like it might work this way).

We thank the reader for their time, and their willingness to get to the bottom of this paper.

Appendix

We’ve attached here a proof of Theorem 1.1.

∞ Theorem 10.1. Let {Hj(s, z)}j=1 be a sequence of holomorphic functions such that Hj(s, z): S ×G → G where S and G are domains in C. Suppose there exists some A ∈ G, such for all compact sets N ⊂ G, the following sum converges,

∞ ∞ X X ||Hj(s, z) − A||z∈N ,s∈S = sup |Hj(s, z) − A| < ∞ j=1 j=1 z∈N ,s∈S Then the expression,

n H(s) = lim Ω Hj(s, z) • z = lim H1(s, H2(s, ...Hn(s, z))) n→∞ j=1 n→∞ Converges uniformly for s ∈ S and z ∈ N as n → ∞ to H, a holomorphic function in s ∈ S, constant in z.

Proof. The first thing we show is for all  > 0, there exists some N, such when m ≥ n > N,

m | Ω Hj(s, z) • z − A| <  j=n For z in N ⊂ G (where A is in the open component of N ), and s ∈ S. This then implies as we let m → ∞, the tail of the infinite composition stays bounded. Forthwith, the infinite composition becomes a normal family, and

49 proving convergence becomes simpler. We provide a quick proof of this inequal- ity.

Set ||Hj(s, z)−A||S,N = ρj. Pick  > 0, and choose N large enough so when n > N,

ρn <  m Denote: φnm(s, z) = Ωj=n Hj(s, z) • z = Hn(s, Hn+1(s, ...Hm(s, z))). We go by induction on the difference m − n = k. When k = 0 then,

||φnn(s, z) − A||S,N = ||Hn(s, z) − A||S,N = ρn <  Assume the result holds for m − n < k, we show it holds for m − n = k. Observe,

||φnm(s, z) − A||S,N = ||Hn(s, φ(n+1)m(s, z)) − A||S,N

≤ ||Hn(s, z) − A||S,N

= ρn < 

Which follows by the induction hypothesis because φ(n+1)m(s, z) ⊂ N –it’s in a neighborhood of A which is in N . That is m − n − 1 < k. m The next step is to observe that Ωj=1 Hj(s, z) is a normal family as m → ∞, for z ∈ N and s ∈ S. This follows because the tail of this composition is m bounded. We can say || Ωj=1 Hj(s, z)||S,N < M for all m. m Since φm(s, z) = Ωj=1 Hj(s, z) • z are a normal family for all compact sets N ⊂ G; there is some constant M ∈ R+ and L ∈ R+ such, dk || φ (s, z)|| ≤ M · k! · Lk dzk m S,N

To see this, take |z − A| < 2δ and observe,

k Z d k! φm(s, ξ) k φm(s, z) = k+1 dξ dz 2πi |ξ−A|=2δ (ξ − z) So that, taking the supremum norm across |z − A| ≤ δ

50 dk k! Z ||φ (s, ξ)|| || φ (s, z)|| ≤ m S dξ dzk m S,|z−A|≤δ 2π k+1 |ξ−A|=2δ |ξ − z||z−A|≤δ k! Z M ≤ k+1 dξ 2π |ξ−A|=2δ δ 2Mk! ≤ δk

Where we’ve used the bound |ξ − z| ≥ δ when |ξ − A| = 2δ and |z − A| ≤ δ. This bound can be derived regardless of N for varying M and L.

Secondly, using Taylor’s theorem,

φm+1(s, z) − φm(s, z) = φm(s, Hm+1(s, z)) − φm(s, z) ∞ k k X d (Hm+1(s, z) − z) = φ (s, z) dzk m k! k=1 ∞ k k−1 X d (Hm+1(s, z) − z) = (H (s, z) − z) φ (s, z) m+1 dzk m k! k=1

So that, setting z = A,

∞ X k k−1 ||φm+1(s, A) − φm(s, A)||s∈S ≤ ||Hm+1(s, A) − A||s∈S ML ||Hm+1(s, A) − A|| k=1 ML ≤ ||H (s, A) − A|| m+1 S 1 − q

For L||Hm+1(s, A) − A||S ≤ q < 1, which is true for large enough m > N. ML Setting C = 1−q . Applying from here,

||φm+1(s, A) − φm(s, A)||s∈S ≤ C||Hm+1(s, A) − A||s∈S This is a convergent series per our assumption. Choose N large enough, so that when m, n > N,

m−1 X  ||H (s, A) − A|| < j+1 s∈S C j=n Then,

51 m−1 X ||φm(s, A) − φn(s, A)||s∈S ≤ ||φj+1(s, A) − φj(s, A)||s∈S j=n m−1 X ≤ C ||Hj+1(s, A) − A||s∈S j=n < 

So we can see φm(s) must be uniformly convergent for s ∈ S, and therefore defines a holomorphic function H(s) as m → ∞. This tells us,

∞ H(s) = Ω Hj(s, z) • z j=1 z=A Converges and is holomorphic. To show this function equals,

∞ Ω Hj(s, z) • z j=1 For all z ∈ G; simply notice that,

∞ Ω Hj(s, z) • z j=m Is arbitrarily close to A as we let m grow (which was shown at the beginning of this proof). Then,

∞ m−1 ∞ Ω Hj(s, z) • z = Ω Hj(s, z) • Ω Hj(s, z) • z j=1 j=1 j=m m−1 ∞ = lim Ω Hj(s, z) • lim Ω Hj(s, z) • z m→∞ j=1 m→∞ j=m ∞ = Ω Hj(s, z) • z j=1 z=A

This concludes our proof.

Acknowledgements

I’d like to thank the community at The Tetration Forum [9]. Without this community, I do not think this paper would’ve come into fruition the manner that it has. In this regard, I would not have been able to spot the error in my

52 initial construction using a function φ which satisfied a different criterion than the family βλ. To this, I owe many thanks to Sheldon Levenstein particularly. I also owe a great deal of thanks to Tom Marcel Raes–who was very per- sistent in the belief that there may be some function in which the original φ method would work–which was pushed to its extreme in [5], but only for real- values. I had significant doubt after the failure of holomorphy with φ, that no function could properly work–produce a holomorphic tetration. But upon the many suggested functions he had, an idea formed–make it solve the tetration equation at ∞! Which led me to consider the family of functions βλ–which are asymptotically tetration. I’d also like to thank the user MphLee, who by happenstance asked me a 1 bunch of questions, which led myself to consider the as e−λs + 1 our multiplier rather than es (as it was with the φ method). Whereupon, this was the function which worked most universally for the construction we needed to answer his questions–or at least attempt to do so. I’d also like to thank him for all of the commutative diagrams he made–which encouraged me to use my own. He aided me greatly with many of his funny drawings. I’d also like to thank Henryk Trappman, for of course curating and creat- ing the forum; and for the disappointment that there were no pictures in the first paper [5], encouraging me to include graphics. And I’d like to thank the many members of the forum who have helped me understand these problems throughout the course of my mathematical development. Where, there exists posts from many years gone by when I was no more than a teenager. Regards, James

References

[1] Milnor, John. (2006). Dynamics In One Complex Variable. Princeton Uni- versity Press. [2] Kneser, Hellmuth. (1950). Reelle analytische Losungen der Gleichung ϕ(ϕ(x)) = ex und verwandter Funktionalgleichungen. Journal fur die reine und angewandte Mathematik. 187: 56-67 [3] M. Lyubich. (1987). The measurable dynamics of the exponential map. J. Math. 28, 111-127. [4] Nixon, James D., (2019). ∆y = esy, Or: How I Learned To Stop Worrying And Love The Γ-Function. arXiv [5] Nixon, James D., (2021). Hyper-operations By Unconventional Means. arXiv [6] Nixon, James D., (2021). Second Order Transfer Equations; and General- izations to Arbitrary Orders. arXiv [7] Nixon, James D., (2020). The Compositional Integral: The Narrow And The Complex Looking Glass. arXiv

53 [8] M. Rees. (1986). The exponential map is not recurrent. Math. Zeit. 191, 593-598. [9] The Tetration Forum, https://math.eretrandre.org/tetrationforum/index.php

54