Interpreting the Wigner–Eckart Theorem

https://doi.org/10.1016/j.shpsa.2021.01.007Get rights and content

Abstract

The Wigner–Eckart theorem is central to the application of symmetry principles throughout atomic, molecular, and nuclear physics. Nevertheless, the theorem has a puzzling feature: it is dispensable for solving problems within these domains, since elementary methods suffice. To account for the significance of the theorem, I first contrast it with an elementary approach to calculating matrix elements. Next, I consider three broad strategies for interpreting the theorem: conventionalism, fundamentalism, and conceptualism. I argue that the conventionalist framework is unnecessarily pragmatic, while the fundamentalist framework requires more ontological commitments than necessary. Conceptualism avoids both defects, accounting for the theorem’s significance in terms of how it epistemically restructures the calculation of matrix elements. Specifically, the Wigner–Eckart theorem modularizes and unifies matrix element problems, thereby changing what we need to know to solve them.

Introduction

Shortly after the founding papers on quantum mechanics in 1924 and 1925, Eugene Wigner (1927) and Hermann Weyl (1927) published the first articles detailing the application of group theory to quantum mechanics. By 1931, these ideas had matured into a comprehensive treatment of atomic spectra (Wigner, 1931). Central to this application is what came to be known as the Wigner–Eckart theorem. Developed in a review article by Carl Eckart (1930, p. 358), the seed of the theorem first appeared in a landmark paper by Wigner (1927, p. 645), the first paper to apply group representation theory to the rotational symmetry of atoms (Biedenharn & van Dam, 1965, p. 6). Although originally developed only for the special rotation group SO(3) and the special unitary group SU(2), numerous generalizations have extensively extended its scope. Today, it applies to almost every group that has ever found application in physics and chemistry, inspiring a host of applications outside its original domain.1 In the words of the physicist J. F. Cornwell, “the Wigner–Eckart Theorem provides both the most succinct and the most powerful expression in the whole field of application of group theory in physical problems. Indeed, most physical applications depend directly on it” (1984, p. 108). Similarly, in a wide-ranging review of applied group theory, the chemical physicist B. G. Wybourne argued that many chemists and physicists had misunderstood the significance of applied group theory by failing to appreciate the Wigner–Eckart theorem:

The myth that group theory is limited to qualitative results in its application to physical problems persists with the failure of most expositors of group theory to give adequate attention to the celebrated Wigner–Eckart theorem that completes the group theoretical picture and turns group theory into a practical subject. (Wybourne, 1973, p. 39)

Despite its centrality, the theorem has a puzzling feature: it is in fact dispensable, since elementary methods provide the same information codified by the theorem. This points to a larger puzzle surrounding many symmetry arguments in science. While providing fantastic simplifications, symmetry arguments are often not necessary. When it comes to solving problems, more elementary approaches often suffice. This lack of necessity invites comparisons with other methodological conveniences employed by science. What separates a generic symmetry argument from a faster computer or a fortuitous notation, such as the Einstein summation convention? Or, if there is little difference, why are symmetry arguments routinely touted as deep and intellectually significant?

Given its dispensability, there are prima facie only two possible sources of the theorem’s significance. Perhaps the theorem is significant due to its ontological status, such as carving nature more closely at the joints or playing a privileged nomological or explanatory role. At the other extreme, perhaps the theorem has only pragmatic benefits, determined by the goals and values of scientific agents. This conventionalist or pragmatist response would pinpoint the significance of the theorem to the eye of the beholder. Yet, there is in fact a third possible source of the theorem’s significance, located within its epistemic roles. These epistemic roles give rise to the theorem’s methodological advantages, while making the theorem more than just convenient. Furthermore, the epistemic significance of the theorem does not depend on any particular ontological interpretation. I will thus argue that one can remain neutral—or agnostic—regarding these rival metaphysical interpretations, while nonetheless characterizing how the theorem makes a non-pragmatic and objective difference for understanding. I will refer to such differences in understanding as aspects of intellectual significance (as distinct from the theorem’s methodological or ontic significance).

Section 2 begins by describing a generic class of problems that the Wigner–Eckart theorem solves. I contrast the Wigner–Eckart approach with an elementary approach that eschews the theorem. Next, Section 3 develops a pragmatic interpretation of the theorem, based on its convenience for solving matrix element problems. Section 4 introduces my preferred philosophical framework for interpreting the theorem: conceptualism. I characterize the source of the theorem’s intellectual significance, locating it within what I call epistemic dependence relations. Section 4.3 shows how these epistemic dependence relations give rise to the theorem’s methodological advantages. These pragmatic benefits are thereby separable from the non-pragmatic epistemic benefits of the theorem. Finally, Section 5 considers competing ontological interpretations of the theorem in terms of different conceptions of laws of nature, joint-carving, and explanation. I show that the intellectual significance of the theorem does not depend on privileging any particular ontological interpretation. Indeed, it is not clear that the theorem or its use adjudicates between these competing ontologies. By focusing on its intellectual rather than ontic dimensions, conceptualism provides an empiricist-friendly interpretation of the Wigner–Eckart theorem.

Section snippets

Two approaches to matrix elements in spectroscopy

Throughout the varied branches of spectroscopy, an important task is to calculate matrix elements of perturbation operators. Perturbations, such as electric dipole and electric quadrupole radiation, lead to transitions from one physical state to another. For a given initial and final state, a matrix element of the perturbation operator quantifies the probability of transition. The greater the matrix element’s magnitude, the greater the probability of transition. In turn, more probable

Methodological significance

The example in Section 2.3 might give the impression that the Wigner–Eckart theorem merely provides a convenient method for calculating matrix elements. Computationally, it replaces calculating integrals with the far simpler algebraic act of multiplication. According to what I’ll call conventionalism, the significance of the theorem stems entirely from these pragmatic benefits. In Section 4, I will reject conventionalism in favor of a more epistemically robust account of the theorem’s

Intellectual significance

Against conventionalism, I will argue that the Wigner–Eckart approach possesses a non-pragmatic and objective significance, stemming from how it epistemically reorganizes the matrix element problem. This is not a matter of mere presentational differences or pragmatic benefits. It is a matter of transforming what we need to know to solve matrix element problems. The theorem illustrates how different approaches to the same problem can manifest intellectual differences, i.e. objective and

Ontological significance

Against my preferred position, a more ontologically-driven philosopher might argue that conceptualism neglects crucial aspects of the significance of the Wigner–Eckart theorem. A fundamentalist would demand an account of what the Wigner–Eckart theorem is, fundamentally. Is it a law of nature (and if so, what kind)? Is it instead a certain kind of explanatory improvement, be it causal or non-causal? Or, is the fundamental nature of the theorem to carve nature more closely at its joints? Just as

Conclusion

I have considered three broad strategies for making sense of the significance of the Wigner–Eckart theorem. The first and most deflationary strategy—conventionalism—focused entirely on the methodological and pragmatic benefits that the theorem provides. I argued that although these benefits surely exist, they are relative to the values and aims of agents. In contrast, the second and third strategies—conceptualism and fundamentalism—sought an objective and non-pragmatic account of why the

Acknowledgements

I thank Jeremy Butterfield for supervising the Master’s thesis where this project originated and for encouraging me to revise it. Klaas Landsman, Jos Uffink, and two MPhil examiners provided helpful feedback on that thesis. Ken Manders helped with early stage conceptualizations of this research, while Dave Baker, Gordon Belot, Laura Ruetsche, and Adam Waggoner have helped with more recent ones. This material is based upon work supported by the National Science Foundation Graduate Research

References (60)

  • V. Gijsbers

    Understanding, explanation, and unification

    Studies In History and Philosophy of Science Part A

    (2013)
  • M. Lange

    Laws and meta-laws of nature: Conservation laws and symmetries

    Studies In History and Philosophy of Science Part B: Studies In History and Philosophy of Modern Physics

    (2007)
  • L. O’Raifeartaigh

    Broken symmetry

  • E. Scholz

    Introducing groups into quantum theory (1926–1930)

    Historia Mathematica

    (2006)
  • V.K. Agrawala

    Wigner–Eckart theorem for an arbitrary group or Lie algebra

    Journal of Mathematical Physics

    (1980)
  • J. Avigad

    Mathematics and language

  • J. Avigad et al.

    The concept of ‘character’ in Dirichlet’s theorem on primes in an arithmetic progression

    Archive for History of Exact Sciences

    (2014)
  • G. Belot

    Geometric possibility

    (2011)
  • L.C. Biedenharn et al.

    Angular momentum in quantum physics: Theory and application

    (1981)
  • L.C. Biedenharn et al.

    Quantum theory of angular momentum: A collection of reprints and original papers

    (1965)
  • A. Bird

    Nature’s metaphysics: Laws and properties

    (2007)
  • A. Bokulich

    Representing and explaining: The eikonic conception of scientific explanation

    Philosophy of Science

    (2018)
  • L. Bonolis

    From the rise of the group concept to the stormy onset of group theory in the new quantum mechanics. A saga of the invariant characterization of physical objects, events and theories

    La Rivista Del Nuovo Cimento

    (2004)
  • H.R. Brown et al.

    The dynamical approach to spacetime theories

  • P.J. Brussaard et al.

    Shell-model applications in nuclear spectroscopy

    (1977)
  • J.N. Butterfield

    Between laws and models: Some philosophical morals of Lagrangian mechanics

    (2004)
  • E.U. Condon et al.

    The theory of atomic spectra

    (1935)
  • J.F. Cornwell
    (1984)
  • C.F. Craver

    The ontic account of scientific explanation

  • R. Creath

    The gentle strength of tolerance: The Logical Syntax of Language and Carnap’s philosophical programme

  • H. Demarest

    Do counterfactuals ground the laws of nature? A critique of Lange

    Philosophy of Science

    (2012)
  • C. Eckart

    The application of group theory to the quantum dynamics of monatomic systems

    Reviews of Modern Physics

    (1930)
  • A. Frank et al.

    Algebraic methods in molecular and nuclear structure physics

    (1994)
  • V. Gijsbers

    Why unification is neither necessary nor sufficient for explanation

    Philosophy of Science

    (2007)
  • J.S. Griffith

    The irreducible tensor method for molecular symmetry groups

    (1962)
  • M. Hamermesh

    Group theory and its application to physical problems

    (1962)
  • Hunt, J. (forthcoming). Epistemic dependence and understanding: Reformulating through symmetry. The British Journal for...
  • F. Iachello et al.

    Algebraic theory of molecules

    (1995)
  • B.R. Judd

    Operator techniques in atomic spectroscopy

    (1963)
  • B.R. Judd

    History of complex spectra

    Reports on Progress in Physics

    (1985)
  • Cited by (1)

    View full text