Next Article in Journal
Dynamic Behavior of a Predator–Prey Model with Double Delays and Beddington–DeAngelis Functional Response
Previous Article in Journal
The Influence of Hedge, Arbitrage, and After-Hours Trading on the Holding Returns of TAIEX Futures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Functional Epistemology “Nullifies” Dyson’s Rebuttal of Perturbation Theory

Institute for Theoretical Physics, TU Wien, Wiedner Hauptstrasse 8-10/136, 1040 Vienna, Austria
Axioms 2023, 12(1), 72; https://doi.org/10.3390/axioms12010072
Submission received: 5 December 2022 / Revised: 6 January 2023 / Accepted: 10 January 2023 / Published: 11 January 2023
(This article belongs to the Section Hilbert’s Sixth Problem)

Abstract

:
Functional epistemology is about ways to access functional objects by using varieties of methods and procedures. Not all such means are equally capable of reproducing these functions in the desired consistency and resolution. Dyson’s argument against the perturbative expansion of quantum field theoretic terms, in a radical form (never pursued by Dyson), is an example of epistemology taken as ontology.
MSC:
00A30; 03F50; 03F03; 41A21; 41A58; 41A60; 41A30

1. Functional Epistemology

Notwithstanding metaphysical preferences about ontological realism, also known as Platonism—asserting that “some [[mathematical or physical objects or]] entities sometimes exist without being experienced by any finite mind” [1,2]—versus mathematical nominalism—claiming that mathematical entities such as numbers and functions do not exist, quasi “a subject with no object” [3]—every application of mathematical formalism requires some operational access to these objects and entities. In a broader perspective this can also be seen as semantics in need for a syntactical formalization.
One important aspect of access is a representation of functional objects and entities that in some formal form correspond to the important aspects of those objects and entities. Nevertheless, although representations vary—spanning a wide range of efficacies and deficiencies—they should not be confused with the respective mathematical objects or entities.
The original informal conception of function y = f ( x ) was that of a unique association of an output “value” y given an input “argument” x. More recent conceptions consider ordered pairs x , y , where again x stands for argument(s) and y for unique value(s); in particular, there must not be two pairs x , y and x , y with y y .
This naive functional conception was challenged by Gödel’s, Kleenee’s, and Touring’s formalization of what functional “access” means; for instance, in the form of paper and pencil operations on a “paper machine” [4]. These developments closely followed the paradigm change from Cantor’s naive set theory [5] to axiomatic set theories [6]; for instance, Zermelo–Fraenkel set theory. Indeed, from a foundational perspective, it might not get worse: Rice’s theorem, usually proven by reduction to the halting problem, states that any nontrivial semantic property of a computable function (evaluated by a Turing machine) is undecidable.
Therefore, due to incompleteness and related theorems, for the sake of formalization, these earlier intuitive and heuristic perceptions of functional performance had to be modified and restricted. The current formalization of functions is in terms of “desirable” properties, such as, effective computability. This has resulted in the abandonment of functional totality—the pretension that any (every) arbitrary argument x can be associated with a (unique) value y—in favor of partiality: certain functions, such as predictors of the large-scale performance of deterministic systems, need not have a value accessible by some algorithm—in short, there is a difference between determinism and predictability [7]. In such a regime, it can no longer be maintained that the value y exists —that is, can be obtained or accessed by some algorithm or computation.
Typical examples of such “critical” functions are Turing’s halting function, or Specker’s theorems of recursive analysis [8,9,10], or Chaitin’s “halting probability” Ω in terms of its bitwise expansion [11]. Physical realizations have been, for instance, by reduction to the halting problem [12], suggested in terms of undecidable classical dynamics [13,14], N-body problems [15], or spectral gaps [16].
It may happen that a program implemented on a computer that “is supposed to compute a limit”—and, with finite resources, even “accesses a few approximations or bounds thereof”—and yet this limit is uncomputable and algorithmically inaccessible because some resources, such as computing time or space, that are necessary to compute this limit with, say, precision up to its nth bit, grow faster asymptotically than any computable function of n [17]. In intuitive algorithmic terms, the difference between a total versus a partial function may be imagined as the distinction between a DO–loop (with fixed finite beginning, ending, and increments) and a WHILE-loop. The latter WHILE-loop may or may not “take forever”, depending on the respective termination condition.
Another area of partial value assignments is quantum mechanics. Extensions of the Kochen–Specker theorem suggest that, relative to the assumption of noncontextuality, only a “star-shaped” [18] [Figure 5] (in terms of hypergraphs [19] representing individual contexts by smooth curves [20]) context can have definite value assignments. Observables in all other contexts must be value indefinite [21,22,23].
Still another issue of functional epistemology is the means relativity of functional representation. The same function can have very different representations and encodings; some exhibit more or less problematic issues. For the sake of an example, we shall later represent one and the same function in five different forms.
The selection of particular means is often not a matter of choice but one of pragmatism or even desperation. Especially theoretical physicists are often criticized for their “relaxed” stance on formal rigor. Dirac’s introduction of the needle-shaped delta function is often quoted as an example. Heaviside, in another instance, responded to criticism for his use of the “highly nonsmooth” unit step function [24] [p. 9, § 225]: “But then the rigorous logic of the matter is not plain! Well, what of that? Shall I refuse my dinner because I do not fully understand the process of digestion? No, not if I am satisfied with the result.”
This, in a nutshell, seems to be the attitude of field theorists regarding the use of perturbation series: It is well documented [25,26,27] that the commonly used power series expansion which can be rewritten as inverse power series expansion
f ( α ) = α a 0 + a 1 α + a 2 α 2 + = n = 0 a n α n = 1 α n = 0 a n 1 α n + 1 .
in terms of the fine structure constant α —that is, the square of the coupling constant—is divergent.
For this power series to converge, there has to be a finite radius of convergence centered at the origin at (fictitious) value α = 0 , thereby including (fictitious) nonvanishing negative values α < 0 within which F has to be analytic. However, because if the (fictitious) coupling between like charges becomes negative, and because by tunneling this cannot be “contained”, the vacuum becomes unstable due to pair creation, and (fictitiously) disintegrates explosively. Hence, Dyson concludes, the power series f ( α ) cannot converge and thus cannot be analytic—a complete contradiction to the assumption.
An immediate reaction would be to perceive these coincidences as “bordering on the mysterious” [28]. This spirit is corroborated [29] by statements like Carrier’s Rule, pointing out that “divergent series converge faster than convergent series because they don’t have to converge.” However, as already surmised by Dyson, quantitative considerations from partial summations show [30] [p. 4] that convergent series “initially”—that is, with only “a few orders” added—may “largely deviate” from the true value of the function it encodes (eg, consider the straightforward Taylor expansion of sin e e e e e ), whereas some associated asymptotic divergent series “initially” converges toward this value: a “reasonable” approximation can be obtained by taking relatively few terms of this divergent series, whereas “many more” terms of the convergent series are needed to achieve that same degree of accuracy.
Current experiences in quantum field theory corroborate this view: although the asymptotic perturbation series have a zero radius of convergence, it is effectively possible to obtain good agreement between theoretical calculations based on asymptotic series and experimental results. As it turns out, the terms in these asymptotic series become increasingly accurate as the series is extended, and hence the error in the truncated series decreases as more terms are included.
This is true, in particular, for the QED contribution to the electron anomalous magnetic moment g 2 up to the tenth order [31], as compared with the experimental value [32,33]. The same applies for the muon anomalous magnetic moment [34,35]. Likewise, the theoretical predictions [36] of the Lamb shift show similar good agreement with experiments [37,38].
However, it is important to keep in mind that asymptotic series eventually diverge as more orders are taken into account. One way of coping with the apparent asymptotic divergence is the resummation of the respective series, in particular, Borel (re)summations [39,40,41,42,43,44], which are in some instances capable to reconstruct an analytic function from its asymptotic expansion [45].

2. Euler’s Series of 1760 and Its Multiple Representations

For the sake of an example that exhibits a wide spread of varied (asymptotic) behaviors “catching” the same “ontologic” function (or, from a nominal point of view, the same “subject without object”), consider a series
s ( x ) = x x 2 + 2 x 3 6 x 4 +
mentioned by Euler in a 1760 publication [46] [§ 6, p. 220]. As already observed by Euler this series can, in a nominal way, be considered a “solution” of
d d x + 1 x 2 s ( x ) = 1 x ;
associated with the differential operator L x = d d x + 1 x 2 . This first-order ordinary differential equation has an irregular (essential) singularity at x = 0 because the coefficient of the zeroth derivative 1 / x 2 has a pole of order 2. Therefore, (3) is not of the Fuchsian type, and cannot be subjected to the Frobenius method of creating convergent power series solutions.
Nevertheless, s ( x ) can be represented in at least five ways, differing substantially with respect to convergence and utility for (physical) computation and prediction. In what follows, these cases will be enumerated: s ( x ) can be represented by
(i)
A convergent Maclaurin series (Ramanujan found a series which converges even more rapidly) solution (4) based on the Stieltjes function;
(ii)
A proper (Borel) summation of Euler’s divergent series (5) [40] [Equation (3.3)];
(iii)
Quadrature, that is, by direct integration of (6) [40] [Equation (3.3)];
(iv)
Evaluating Euler’s (asymptotic) divergent series (7) to “optimal order” [40] [Equations (2.12)–(2.14)]; and
(v)
Evaluating the respective inverse factorial series (8) [47] [Equation (5.7)].
Let S ( x ) = 0 e t / ( 1 + t x ) d t stand for the Stieltjes function (cf. [48] [formula 5.1.28, p. 230] but with x 1 x ), γ = lim n j = 1 n 1 j log n 0.5772 , be the Euler–Mascheroni constant [49], Γ ( z , x ) represent the upper incomplete gamma function [48] [formula 6.5.3, p. 260], B s ( x ) be the Borel transform of s ( x ) , x n = Γ x + n / Γ x = x x + 1 x + j 1 and ( x ) 0 = 1 be Pochhammer symbols [48] [Section 24.1.3, p. 824], also known as the rising factorial power, and S j ( k ) be Sterling numbers of the first kind that are the polynomial coefficients of the Pochhammer symbol ( z j + 1 ) j [48] [Section 24.1.3, p. 824]; that is, k = 0 j S j ( k ) z k = ( z j + 1 ) j = ( 1 ) j ( z ) j for j N { 0 } . Then
s ( x ) = x S ( x ) = e 1 x Γ 0 , 1 x
= e 1 x γ log x + n = 1 ( 1 ) n n ! n x n = 0 B S ( y ) e y x d y
= 0 e y x 1 + y d y = 0 x e t 1 + x t d t
= 0 x e 1 x 1 t t d t
= j = 0 ( 1 ) j j ! x j + 1
= x j = 0 ( 1 ) j 1 x j + 1 k = 0 j S j ( k ) k ! .
What we can learn from this prototypic example is the wide variety of mathematical representations associated with one and the same function. Not everybody might agree with all the equality signs in (4)–(8) and the legality of the respective methods though, thereby reflecting a variety of metamathematical stances.

3. Quantum Field Theoretical Perturbation Series Need Not Diverge

Let us discuss two critical aspects in the derivation of the power series expansion of (1). One critical step in the derivation of f amounts to interchanging a sum with an integral in the case of nonuniform convergence of the former [50] [Sect. II.A]. One may perceive asymptotic divergence as a “penalty” for such manipulations. It may come as a surprise that those calculations performed well for empirical predictions.
Ritt’s theorem inspires one strategy to cope with such issues [51,52] by stating that any (not necessarily asymptotic) divergent power series with arbitrary coefficients can be converted into nonunique analytic functions. Thereby, every summand is multiplied with a suitable nonunique “convergence factor” (conversely, every analytic function can be approximated by a unique asymptotic series).
A general regularisation of divergent series using such convergence factors, also called cutoff functions, has been recently introduced by Tao [53] [Section 3.7]. The resulting smoothed sums may become uniformly convergent, thereby allowing interchanging a sum with an integral and avoiding the aforementioned issues while preserving inherent properties of the original divergent series. This is not dissimilar to the use of test functions in the theory of distributions.
A second critical aspect is the expansion of (1) in terms of a power series, and the resulting vanishing of the radius of convergence. Dyson already mentioned a possible remedy, his “Alternative A: There may be discovered a new method of carrying through the renormalization program, not making use of power series expansions.” One such candidate expansion that does not necessarily share the catastrophic fate of the power series caused by the “explosive disintegration” of the vacuum state for negative arguments, is an expansion of f ( α ) in terms of inverse factorial series [54,55] and recently investigated by Weniger [47] as well as O. Costin, R. D. Costin, and Dunne [44,56]:
f ( α ) = b 0 0 ! α + b 1 1 ! α ( α + 1 ) + = n = 0 b n n ! ( α ) n + 1 ,
where again α n + 1 are Pochhammer symbols.
Stirling numbers of the first kind S j ( k ) mentioned earlier serve as “translations”—that is, as expansions from an inverse power 1 / α n + 1 in terms of inverse factorial series ( α ) n + j + 1 : for k N { 0 } [57] [Equation (6), § 30, p. 78],
1 α n + 1 = j = 0 ( 1 ) j ( α ) n + j + 1 S n + j ( n ) .
The respective “reverse” expansion of a Pochhammer symbol ( α ) k + 1 in terms of an inverse power series 1 / α n + j + 1 for k N { 0 } and | α | > 0 [57] [Equation (9), § 26, p. 68] is given by:
1 ( α ) n + 1 = j = 0 ( 1 ) j α n + j + 1 S n + j ( n ) .
Insertion of (10) into (1), rearranging the order of the summations through an index shift m = n + j with n 0 and j 0 , hence m 0 and j = m n 0 and n m yields:
f α = 1 α n = 0 a n j = 0 ( 1 ) j 1 α n + j + 1 S n + j ( n ) = 1 α j = 0 n = 0 a n ( 1 ) j 1 α n + j + 1 S n + j ( n ) = 1 α m = 0 ( 1 ) m 1 α m + 1 n = 0 m ( 1 ) ± n S m ( n ) a n .
Therefore, if we define the inverse power series f ( β ) = ( 1 / β ) f ( 1 / β ) = m = 0 a m / β m + 1 = m = 0 b m m ! / ( β ) m + 1 with β = 1 / α , then, by comparison,
b m = 1 m ! n = 0 m ( 1 ) m ± n S m ( n ) a n .
For the sake of studying the fascinating convergence [29,47,55,57] of the inverse factorial series, note that terms of the form n ! / ( z ) n + 1 can, for large z , be estimated with the help of [48] [Formula 6.1.47] Γ ( z + a ) / Γ ( z + b ) = z a b 1 + O 1 z and for large n as follows:
n ! ( α ) n + 1 = Γ ( n + 1 ) Γ ( α + n + 1 ) / Γ ( α ) = Γ ( n + 1 ) Γ ( n + 1 + α ) Γ ( α ) = ( n + 1 ) α 1 + O 1 n + 1 ( α 1 ) ! = O n α .
Therefore, the inverse factorial series (9) converges with the possible exception of the points α = m with m N { 0 } (where the Pochhammer symbols in the denominator might vanish) if and only if the associated Dirichlet series n = 1 c n n α converges.
Unlike a power series that has a radius of convergence, a Dirichlet series has an abscissa of convergence ( α ) > λ , that is, it converges on this half-plane [58] [§ 58, 255, p. 456]. Therefore, the inverse factorial series may converge for all positive coupling constants α > 0 although it may diverge for negative values α < 0 . The physically relevant region lies within the abscissa of convergence. Even if the inverse power series diverges factorially, the respective inverse factorial series may converge, but this has to be checked explicitly.
Thus, as already suggested by Dyson [25], representing quantum field theoretical entities in terms of the Tomonaga–Schwinger–Feynman–Dyson power series expansion in the coupling constant [59] may suffice for all practical purposes [60] so far, although their divergencies may cause uneasiness for a variety of (pragmatic and formal) reasons. One should not confuse these field-theoretic entities with their actual representations, that is, functional ontology with epistemology. Such considerations might present a positive outlook for an improved theory of convergent perturbation series. A candidate for such a theory might be inverse factorial expansions exhibiting an abscissa rather than a radius of convergence.
However, a convergence issue encountered in inverse factorial series is the Stokes phenomenon [44,56]: the asymptotic behavior of functions need not be uniform in different regions of the complex plane, bounded by (anti-)Stokes lines. In particular, inverse factorial series may not be suitable for the study of Stokes phenomena if Stokes lines are present in the right complex half-plane ( α ) > λ because of the singularities on these Stokes lines. One may conjecture that inverse factorials might converge in regions where the associated power series are Borel summable; yet convergence fails in the presence of Stokes lines. This would mean that quantum field theories have convergent inverse factorial expansions only in less than four dimensions; and that this expansion might fail for four dimensions. Nevertheless, Berry, O. Costin, R. D. Costin, and Howls have pointed out [61] [p. 10] that, although “typically, classical factorial series have two major limitations: slow convergence, at best power-like, and a limited domain of convergence (a half plane which cannot be centered on the asymptotically important Stokes line) … for resurgent functions these limitations can be overcome. Ecallé–Borel summable series can be summed by rapidly convergent factorial series.”
Transseries from Borel–Ecallé summations of divergent (power) series offer a method to “recover” nonperturbative information from such power series [62,63], thereby indicating that the divergent perturbative power series expansion contains information of the nonperturbative kind. The situation is not totally dissimilar from tempered distributions: using test functions with unbounded (noncompact) support allows the representation and reconstruction of generalized functions by Fourier transforms.
A further method for alternative representations of functions are Padé approximations by rational functions (of given order) near a specific point. Padé approximantions offer practical methods of defining and computating the value of a power series even if such series diverge [64].

4. Summary

In this brief exposé I have bundled together two ideas: first, that mathematical objects or entitities such as functions or proofs [65] may have very varied representations and realizations. Not all of them might require comparable means to access them—think of convergence or (asymptotic) divergence, or of partial functions. Different means might not be equally appropriate or sufficient and necessary for different purposes.
Second, in particular and more specifically, as conjectured already by Dyson, arguments against the existence or convergence of power expansions of Tomonaga–Schwinger–Feynman–Dyson perturbative quantum field theory might be “liftable” by using other expansion techniques.
For the sake of illustration, suppose for a moment that Gödel’s “unadulterated” Platonism [2,66] is acceptable. (An analogous argument can be made within nominalism.) Then mathematical objects or entities such as functions can be conceptualized by their ontological existence.
However, on second thought, it is an entirely different, highly nontrivial, issue to “touch” or to epistemically access those objects or entities. We have presented some examples of such access which analytically spread over a wide variety of (asymptotic) divergent and convergent expressions.
We have, in particular, argued that (asymptotic) divergence of a particular type of perturbation series based on power series expansions could be overcome by other methods of perturbative series; in particular, by inverse factorial series. This still leaves open the consistent existence of quantum field theory, but at least it indicates conceivable convergent access to quantum field theoretical objects and functions.

Funding

This research was funded in whole, or in part, by the Austrian Science Fund (FWF), Project No. I 4579-N. For the purpose of open access, the author has applied a CC BY public copyright licence to any author accepted manuscript version arising from this submission.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

I kindly acknowledge explanations by and considerations with Cristian S. Calude, Alexander Leitsch and Noson S. Yanofsky, as well as discussions with and suggestions by Thomas Sommer. Mea culpa if I got them wrong!

Conflicts of Interest

The author declares no conflict of interest. The funders had no role in the design of the study, in the collection, analyses, or interpretation of data, in the writing of the manuscript, or in the decision to publish the results.

References

  1. Stace, W.T. The Refutation of Realism. Mind 1934, 43, 145–155. [Google Scholar] [CrossRef]
  2. Parsons, C. Platonism and Mathematical Intuition in Kurt Gödel’s Thought. Bull. Symb. Log. 1995, 1, 44–74. [Google Scholar] [CrossRef] [Green Version]
  3. Burgess, J.P.; Rosen, G. A Subject with No Object; Oxford University Press: Oxford, UK, 1999. [Google Scholar] [CrossRef]
  4. Turing, A.M. Intelligent Machinery. In Cybernetics. Key Papers; Evans, C.R., Robertson, A.D.J., Eds.; Butterworths: London, UK, 1968; pp. 27–52. [Google Scholar]
  5. Halmos, P.R. Naive Set Theory; Undergraduate Texts in Mathematics; Springer: New York, NY, USA, 1974. [Google Scholar] [CrossRef]
  6. Hrbacek, K.; Jech, T. Introduction to Set Theory, 3rd ed.; CRC Press: Boca Raton, FL, USA, 1999. [Google Scholar] [CrossRef]
  7. Suppes, P. The Transcendental Character of Determinism. Midwest Stud. Philos. 1993, 18, 242–257. [Google Scholar] [CrossRef]
  8. Specker, E. Der Satz vom Maximum in der rekursiven Analysis. In Constructivity in Mathematics: Proceedings of the Colloquium Held at Amsterdam, 1957; Heyting, A., Ed.; North-Holland Publishing Company: Amsterdam, The Netherlands, 1959; pp. 254–265. [Google Scholar] [CrossRef]
  9. Specker, E. Selecta; Birkhäuser Verlag: Basel, Switzerland, 1990. [Google Scholar] [CrossRef]
  10. Kreisel, G. A notion of mechanistic theory. Synthese 1974, 29, 11–26. [Google Scholar] [CrossRef]
  11. Calude, C.S.; Dinneen, M.J. Exact Approximations of Omega Numbers. Int. J. Bifurc. Chaos 2007, 17, 1937–1954. [Google Scholar] [CrossRef]
  12. Yanofsky, N.S. Paradoxes, Contradictions, and the Limits of Science. Am. Sci. 2016, 104, 166. [Google Scholar] [CrossRef]
  13. Moore, C.D. Unpredictability and undecidability in dynamical systems. Phys. Rev. Lett. 1990, 64, 2354–2357. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Bennett, C.H. Undecidable dynamics. Nature 1990, 346, 606–607. [Google Scholar] [CrossRef]
  15. Svozil, K. Omega and the time evolution of the N-body problem. In Randomness and Complexity, from Leibniz to Chaitin; Calude, C.S., Ed.; World Scientific: Singapore, 2007; pp. 231–236. [Google Scholar] [CrossRef] [Green Version]
  16. Cubitt, T.S.; Perez-Garcia, D.; Wolf, M.M. Undecidability of the spectral gap. Nature 2015, 528, 207–211. [Google Scholar] [CrossRef] [Green Version]
  17. Rado, T. On Non-Computable Functions. Bell Syst. Tech. J. 1962, 41, 877–884. [Google Scholar] [CrossRef]
  18. Abbott, A.A.; Calude, C.S.; Svozil, K. Value-indefinite observables are almost everywhere. Phys. Rev. A 2014, 89, 032109. [Google Scholar] [CrossRef] [Green Version]
  19. Bretto, A. Hypergraph Theory; Mathematical Engineering; Springer: Cham, Switzerland, 2013; pp. xiv, 119. [Google Scholar] [CrossRef]
  20. Greechie, R.J. Orthomodular lattices admitting no states. J. Comb. Theory Ser. A 1971, 10, 119–132. [Google Scholar] [CrossRef] [Green Version]
  21. Pitowsky, I. Infinite and finite Gleason’s theorems and the logic of indeterminacy. J. Math. Phys. 1998, 39, 218–228. [Google Scholar] [CrossRef]
  22. Hrushovski, E.; Pitowsky, I. Generalizations of Kochen and Specker’s theorem and the effectiveness of Gleason’s theorem. Stud. Hist. Philos. Sci. Part Stud. Hist. Philos. Mod. Phys. 2004, 35, 177–194. [Google Scholar] [CrossRef] [Green Version]
  23. Abbott, A.A.; Calude, C.S.; Svozil, K. A variant of the Kochen-Specker theorem localising value indefiniteness. J. Math. Phys. 2015, 56, 102201. [Google Scholar] [CrossRef] [Green Version]
  24. Heaviside, O. Electromagnetic Theory; “The Electrician” Printing and Publishing Corporation: London, UK, 1894–1912. [Google Scholar]
  25. Dyson, F.J. Divergence of Perturbation Theory in Quantum Electrodynamics. Phys. Rev. 1952, 85, 631–632. [Google Scholar] [CrossRef]
  26. Le Guillou, J.C.; Zinn-Justin, J. Large-Order Behaviour of Perturbation Theory; Current Physics-Sources and Comments; Elsevier: Amsterdam, The Netherlands, 1990; Volume 7. [Google Scholar]
  27. Vainshtein, A.I. Decaying Systems and Divergence of the Series of Perturbation Theory. In Continuous Advances in QCD 2002; World Scientific Publishing Co. Pte. Ltd.: Singapore, 2002. [Google Scholar] [CrossRef]
  28. Wigner, E.P. The unreasonable effectiveness of mathematics in the natural sciences. Richard Courant Lecture delivered at New York University, May 11, 1959. Commun. Pure Appl. Math. 1960, 13, 1–14. [Google Scholar] [CrossRef]
  29. Landau, E. Über die Grundlagen der Theorie der Fakultätenreihen. Sitzungsberichte Bayer. Akad. Wiss. 1906, 36, 151–218, 482. [Google Scholar]
  30. Bleistein, N.; Handelsman, R.A. Asymptotic Expansions of Integrals; Dover Books on Mathematics: Dover, Delaware, 1975. [Google Scholar]
  31. Aoyama, T.; Hayakawa, M.; Kinoshita, T.; Nio, M. Tenth-Order QED Contribution to the Electron g − 2 and an Improved Value of the Fine Structure Constant. Phys. Rev. Lett. 2012, 109, 111807. [Google Scholar] [CrossRef] [Green Version]
  32. Hanneke, D.; Fogwell, S.; Gabrielse, G. New Measurement of the Electron Magnetic Moment and the Fine Structure Constant. Phys. Rev. Lett. 2008, 100, 120801. [Google Scholar] [CrossRef] [Green Version]
  33. Hanneke, D.; Fogwell Hoogerheide, S.; Gabrielse, G. Cavity control of a single-electron quantum cyclotron: Measuring the electron magnetic moment. Phys. Rev. A 2011, 83, 052122. [Google Scholar] [CrossRef] [Green Version]
  34. Aoyama, T.; Hayakawa, M.; Kinoshita, T.; Nio, M. Complete Tenth-Order QED Contribution to the Muon g − 2. Phys. Rev. Lett. 2012, 109, 111808. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Keshavarzi, A.; Khaw, K.S.; Yoshioka, T. Muon g − 2: A review. Nucl. Phys. B 2022, 975, 115675. [Google Scholar] [CrossRef]
  36. Janka, G.; Ohayon, B.; Crivelli, P. Muonium Lamb shift: Theory update and experimental prospects. EPJ Web Conf. 2022, 262, 01001. [Google Scholar] [CrossRef]
  37. Bezginov, N.; Valdez, T.; Horbatsch, M.; Marsman, A.; Vutha, A.C.; Hessels, E.A. A measurement of the atomic hydrogen Lamb shift and the proton charge radius. Science 2019, 365, 1007–1012. [Google Scholar] [CrossRef]
  38. Ohayon, B.; Janka, G.; Cortinovis, I.; Burkley, Z.; Borges, L.d.S.; Depero, E.; Golovizin, A.; Ni, X.; Salman, Z.; Suter, A.; et al. Precision Measurement of the Lamb Shift in Muonium. Phys. Rev. Lett. 2022, 128, 011802. [Google Scholar] [CrossRef]
  39. Boyd, J.P. The Devil’s Invention: Asymptotic, Superasymptotic and Hyperasymptotic Series. Acta Appl. Math. 1999, 56, 1–98. [Google Scholar] [CrossRef] [Green Version]
  40. Rousseau, C. Divergent series: Past, present, future. Math. Rep. C. R. Math. 2016, 38, 85–98. [Google Scholar]
  41. Flory, M.; Helling, R.C.; Sluka, C. How I Learned to Stop Worrying and Love QFT. arXiv 2012, arXiv:1201.2714. [Google Scholar]
  42. Costin, O. Asymptotics and Borel Summability; Monographs and Surveys in Pure and Applied Mathematics; Chapman & Hall/CRC; Taylor & Francis Group: Boca Raton, FL, USA, 2009; Volume 141. [Google Scholar]
  43. Zinn-Justin, J. Summation of divergent series: Order-dependent mapping. Appl. Numer. Math. 2010, 60, 1454–1464. [Google Scholar] [CrossRef] [Green Version]
  44. Costin, O.; Dunne, G.V. Convergence from divergence. J. Phys. A Math. Theor. 2017, 51, 04LT01. [Google Scholar] [CrossRef] [Green Version]
  45. Brüning, E.A.K. How to reconstruct an analytic function from its asymptotic expansion? Complex Var. Theory Appl. Int. J. 1996, 30, 199–220. [Google Scholar] [CrossRef]
  46. Euler, L. De seriebus divergentibus. Novi Comment. Acad. Sci. Petropolitanae 1760, 5, 205–237. Available online: https://scholarlycommons.pacific.edu/euler-works/247/ (accessed on 6 January 2023).
  47. Weniger, E.J. Summation of divergent power series by means of factorial series. Appl. Numer. Math. 2010, 60, 1429–1441. [Google Scholar] [CrossRef] [Green Version]
  48. Abramowitz, M.; Stegun, I.A. (Eds.) Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables; Number 55 in National Bureau of Standards Applied Mathematics Series; U.S. Government Printing Office: Washington, DC, USA, 1964; pp. xiv, 1046.
  49. Sloane, N.J.A. A001620 Decimal Expansion of Euler’s Constant (or the Euler-Mascheroni Constant), Gamma. (Formerly M3755 N1532). The On-Line Encyclopedia of Integer Sequences. 2019. Available online: https://oeis.org/A027642 (accessed on 17 July 2019).
  50. Pernice, S.A.; Oleaga, G. Divergence of perturbation theory: Steps towards a convergent series. Phys. Rev. D 1998, 57, 1144–1158. [Google Scholar] [CrossRef] [Green Version]
  51. Pittnauer, F. Vorlesungen über Asymptotische Reihen; Lecture Notes in Mathematics; Springer: Berlin/Heidelberg, Germany, 1972; Volume 301. [Google Scholar] [CrossRef]
  52. Remmert, R. Theory of Complex Functions, 1st ed.; Graduate Texts in Mathematics; Springer: New York, NY, USA, 1991; Volume 122. [Google Scholar] [CrossRef]
  53. Tao, T. Compactness and Contradiction; American Mathematical Society: Providence, RI, USA, 2013. [Google Scholar]
  54. Watson, G.N. The transformation of an asymptotic series into a convergent series of inverse factorials [Memoir crowned by the Danish Royal Academy of Science]. Rend. Circ. Mat. Palermo 1912, 34, 41–88. [Google Scholar] [CrossRef]
  55. Doetsch, G. Handbuch der Laplace-Transformation: Band II Anwendungen der Laplace-Transformation; Springer Basel AG (Birkhäuser): Basel, Switzerland, 1972. [Google Scholar] [CrossRef]
  56. Costin, O.; Costin, R.D. A new type of factorial series expansions and applications. arXiv 2016, arXiv:1608.01010. [Google Scholar] [CrossRef]
  57. Nielsen, N. Die Gammafunktion; AMS Chelsea Publishing: Bronx, New York, NY, USA, 1965. Reprint of “Handbuch der Theorie der Gammafunktion”, first published in 1906, and “Theorie des Integrallogarithmus und verwandter Transzendenten”, first published in 1906. [Google Scholar]
  58. Knoop, K. Theorie und Anwendung der Unendlichen Reihen; Springer: Berlin/Heidelberg, Germany, 1996. [Google Scholar] [CrossRef]
  59. Dyson, F.J. The Radiation Theories of Tomonaga, Schwinger, and Feynman. Phys. Rev. 1949, 75, 486–502. [Google Scholar] [CrossRef]
  60. Bell, J.S. Against ‘measurement’. Phys. World 1990, 3, 33–41. [Google Scholar] [CrossRef]
  61. Berry, M.V.; Costin, O.; Costin, R.D.; Howls, C.J. Borel Plane Resurgence in Hyperasymptotics and Factorial Series, 2016. Talk at the Resurgence Meeting, presented by Ovidiu Costin on July 19, 2016. Available online: https://math.tecnico.ulisboa.pt/seminars/resurgence/?action=show&id=4371 (accessed on 7 July 2022).
  62. Costin, O. Exponential asymptotics, transseries, and generalized Borel summation for analytic, nonlinear, rank-one systems of ordinary differential equations. Int. Math. Res. Not. 1995, 1995, 377. [Google Scholar] [CrossRef]
  63. Edgar, G.A. Transseries for Beginners. Real Anal. Exch. 2009, 35, 253–310. [Google Scholar] [CrossRef]
  64. Baker, G.A.; Graves-Morris, P. Pad’e Approximants, 2nd ed.; Encyclopedia of Mathematics and Its Applications; Cambridge University Press: Cambridge, UK, 1996. [Google Scholar] [CrossRef]
  65. Aigner, M.; Ziegler, G.M. Proofs from THE BOOK, 4th ed.; Springer: Heidelberg, Germany, 1998–2010. [Google Scholar] [CrossRef]
  66. Kreisel, G. Kurt Gödel. 28 April 1906–14 January 1978. Biogr. Mem. Fellows R. Soc. 1980, 26, 148–224. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Svozil, K. Functional Epistemology “Nullifies” Dyson’s Rebuttal of Perturbation Theory. Axioms 2023, 12, 72. https://doi.org/10.3390/axioms12010072

AMA Style

Svozil K. Functional Epistemology “Nullifies” Dyson’s Rebuttal of Perturbation Theory. Axioms. 2023; 12(1):72. https://doi.org/10.3390/axioms12010072

Chicago/Turabian Style

Svozil, Karl. 2023. "Functional Epistemology “Nullifies” Dyson’s Rebuttal of Perturbation Theory" Axioms 12, no. 1: 72. https://doi.org/10.3390/axioms12010072

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop