Download as pdf or txt
Download as pdf or txt
You are on page 1of 37

Interpreting the Wigner-Eckart

Theorem

Joshua R. Hunt

Submitted in partial fulfillment of


MPhil in History and Philosophy of Science

Under the supervision of Dr. Jeremy Butterfield

University of Cambridge
United Kingdom
June 1, 2015
Abstract

The Wigner-Eckart theorem expresses a restraint that symmetry places on quantum-


mechanical operators, vastly simplifying matrix element computations. Originally proven
for the rotation group, the Wigner-Eckart theorem applies to most symmetry groups used
in physics and chemistry. However, it is never necessary to use this theorem. This lack
of necessity leads to a philosophical puzzle: why is the Wigner-Eckart theorem concep-
tually significant, rather than merely convenient? This puzzle applies to many cases of
theoretical equivalence in science, cases where adopting alternative expressive means is
convenient but not necessary.
After introducing the Wigner-Eckart theorem, I argue that it is conceptually sig-
nificant because it performs a novel functional role: it shows how symmetry restrains
matrix elements. Via this restraint, the Wigner-Eckart theorem decomposes the problem
of computing matrix elements into more general sub-problems that are collectively more
tractable than the original problem, making the theorem convenient.
Generalizing these morals, I conclude by proposing a criterion for distinguishing the-
oretically equivalent expressive means. I argue that a difference in expressive means is
conceptually significant if and only if it leads to different functional roles.

Contents

1 Introducing the Wigner-Eckart Theorem 2


1.1 Two Approaches to Matrix Elements . . . . . . . . . . . . . . . . . . . . 3
1.1.1 An elementary approach . . . . . . . . . . . . . . . . . . . . . . . 3
1.1.2 The traditional Wigner-Eckart theorem . . . . . . . . . . . . . . . 5
1.2 An Historical Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.2.1 Genesis of the Wigner-Eckart theorem . . . . . . . . . . . . . . . 7
1.2.2 Applications to atoms, nuclei, and molecules . . . . . . . . . . . . 8
1.3 Connections and Inspirations . . . . . . . . . . . . . . . . . . . . . . . . 10
1.3.1 Schemes, laws of nature, and models . . . . . . . . . . . . . . . . 10
1.3.2 Idealization and approximation . . . . . . . . . . . . . . . . . . . 11
1.3.3 Scientific explanation . . . . . . . . . . . . . . . . . . . . . . . . . 12

2 Methodological and Conceptual Significance 15


2.1 A Puzzle: Theoretical Equivalence, Expressive Means, and Convenience . 15
2.2 Modularization, Generality, and Tractability . . . . . . . . . . . . . . . . 18
2.2.1 Modularization . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.2.2 Solving three sub-problems . . . . . . . . . . . . . . . . . . . . . . 20
2.3 The Conceptual Significance of the Wigner-Eckart Theorem . . . . . . . 23
2.3.1 A standard interpretation of the Wigner Eckart theorem . . . . . 23
2.3.2 Information, restraints, and functional roles . . . . . . . . . . . . 24
2.4 A Spectrum of Expressive Means . . . . . . . . . . . . . . . . . . . . . . 26
2.4.1 A proposal: (Restraint) . . . . . . . . . . . . . . . . . . . . . . . . 27
2.4.2 My proposal: (Function) . . . . . . . . . . . . . . . . . . . . . . . 28
2.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
1
Chapter 1

Introducing the Wigner-Eckart


Theorem

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. (Cornwell
1984, 108)

Although no one disputes the importance of symmetry arguments in chemistry and


physics, many of these arguments have a puzzling character. Despite 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 com-
parisons with numerous 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 conceptually significant?
Prima facie, symmetry arguments occupy a space somewhere between convenience
and necessity, yet this vague categorization does little to explain their importance. It
would be nice to have a general account of why symmetry arguments—especially the
unnecessary ones—are conceptually significant. Focusing on prominent formulations of
classical mechanics, Butterfield has provided the beginnings of an answer (2004a, 2004b,
2006). To avoid myopia, a full account would have to dive down from quantum chem-
istry to quantum field theory, while not forgetting the broad expanse between classical
mechanics and cosmology. Thankfully, there is no veto against working on one example
at a time. In the spirit of this broader project, the present essay analyzes a prominent
class of symmetry arguments originating in atomic physics, collectively known as “the
Wigner-Eckart theorem.”
In its traditional form, the Wigner-Eckart theorem applies to any spherically symmet-
ric quantum-mechanical system. This includes any sufficiently isolated atom, nucleus, or
molecule. Since the energy of an isolated system is invariant under arbitrary rotations
in three-dimensional space, we can classify states using the rotation group in three di-
mensions, SO(3). The Wigner-Eckart theorem also generalizes to symmetry groups other
than SO(3), including arbitrary finite groups and compact Lie groups. Extended to finite
groups, the theorem applies to the symmetry groups of molecules and crystals. Extended
to compact Lie groups, the theorem applies to the complex spectra of atoms and nuclei.
Compact Lie groups also arise in the more advanced aspects of molecular spectra and
ligand field theory, which deals with crystals and transition metal ions.

2
Despite these varied applications, the Wigner-Eckart theorem remains neglected in
the philosophical literature on physical symmetries. To introduce the Wigner-Eckart the-
orem, Section 1.1 provides essential details. I begin in Section 1.1.1 by describing an
elementary approach for computing matrix elements. Then, using the case of SO(3) as
an example, Section 1.1.2 describes the Wigner-Eckart theorem and its basic function.
Section 1.2 describes the aforementioned applications, providing a selective overview of
the theorem’s historical development and extensions in various branches of physics and
chemistry. Although the Wigner-Eckart theorem—and applied group representation the-
ory more generally—cries out for historical engagement, this task lies beyond my scope.
Instead, my historical discussion is designed to situate the philosophical puzzle broached
above. Before developing this puzzle in Chapter 2, Section 1.3 connects the Wigner-
Eckart theorem to related topics in philosophy of science. These include laws of nature,
idealization, approximation, and explanation. I justify why none of these topics forms
the focus of my philosophical discussion.
This chapter aims to introduce the Wigner-Eckart theorem as a promising case study
for philosophy of science. In Chapter 2, I make good on this promise by developing a
puzzle about theoretical equivalence. This puzzle arises because both the elementary
approach to matrix elements and the Wigner-Eckart theorem solve the same class of
problems and share the same physical ontology. In other words, they are theoretically
equivalent. As a result, it is never necessary to invoke the theorem: the problems it solves
can be solved without it. This illustrates the question raised above: why is the theorem
conceptually significant rather than simply a matter of convenience? To generalize this
question, it is helpful to adopt Kenneth Manders’ notion of “expressive means,” intro-
duced in the context of Euclidean diagrammatic reasoning (2008a, 2008b). Expressive
means comprise the notations, diagrams, and conceptual resources that a problem-solving
framework makes available. For example, Descartes’ analytic geometry uses polynomial
equations, an expressive means not available in traditional Euclidean geometry (Manders
2008a, 76). Likewise, the Wigner-Eckart theorem uses different expressive means than
the elementary approach. This motivates a general philosophical question: in cases of
theoretically equivalent expressive means, what distinguishes a conceptually significant
expressive difference from one that is merely convenient? Chapter 2 proposes an answer
to this question.

1.1 Two Approaches to Matrix Elements


This section introduces two approaches for computing matrix elements, fortifying the
reader for more detailed discussions ahead. I begin by describing an elementary approach,
which serves as a foil for the Wigner-Eckart theorem. Next, I describe the ingredients
that go into the traditional Wigner-Eckart theorem, which holds for rotationally invariant
systems.

1.1.1 An elementary approach


In all branches of spectroscopy, an important theoretical task is to calculate matrix el-
ements of perturbation operators. Perturbations, such as electric dipole and electric

3
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 prob-
ability of transition. The greater the matrix element’s magnitude, the greater the prob-
ability of transition. In turn, more probable transitions lead to more intense emission or
absorption lines compared to less probable transitions.
As an example, consider the electric dipole operator P ≡ −ex, where e is the charge of
an electron, and x is the position operator. This perturbation operator leads to the most
intense electronic transitions in atoms, orders of magnitude more intense than electric
quadrupole radiation. To calculate these transition intensities, we calculate the matrix
elements of P: hα0 | P |αi, where |αi and |α0 i are the initial and final states of a transition-
ing electron, respectively. Squaring the absolute value of this matrix element provides
the probability for the state |αi to transition to |α0 i under the action of electric dipole
radiation: Transition probability = |hα0 | P |αi|2 .
An elementary approach for calculating matrix elements is to compute an inner prod-
uct in Hilbert space for each desired matrix element. For example, using the position
representation of elementary wave mechanics, we can compute an integral over three
dimensional space:
Z
hα | P |αi = ψα∗ 0 (x)Pψα (x)dx3 .
0
(1.1)

This computation requires integrating the product of the wave-mechanical function for
the perturbation operator and the initial and final wavefunctions. By “the elementary
approach,” I mean just this procedure of computing each matrix element seriatum using
Equation 1.1.
As Butterfield has emphasized, it is important to specify at the outset what it means
to “solve a problem” in physics (2004a, 10-11). The meaning of this phrase varies across
different problem-solving contexts, forming a spectrum from in-principle solvable (the
most liberal) to the most stringent requirements, such as solvable via analytic functions
or in polynomial time. Throughout my discussion of computing matrix elements of oper-
ators, I intend a moderately restrictive sense of problem-solving. I consider the problem
to be solved only if one can actually “get hold of” the relevant matrix elements. This
requires that the inner products are at least numerically computable. As shown below,
if the elementary approach satisfies this requirement, then the Wigner-Eckart theorem
does as well, and vice versa. Thus, this requirement does not distinguish between the
approaches.
Using the elementary approach, one has complete freedom to choose a basis {|αi} for
expressing matrix elements. In contrast, the Wigner-Eckart theorem is restricted to a par-
ticular kind of basis—a standard angular momentum basis—defined in the next section.
This difference arises because the elementary approach applies to quantum-mechanical
systems regardless of their symmetry, whereas the Wigner-Eckart theorem only applies
to systems with well-defined symmetry properties. This difference illustrates a moral
that Butterfield has aptly termed “(Restrict)” (2004a, 29). Namely, the Wigner-Eckart
theorem provides a conceptual advance in part by restricting attention to systems with
well-defined symmetry properties. Nevertheless, the basis used by the Wigner-Eckart the-
orem is usually the most convenient for computing matrix elements via the elementary
approach as well. Hence, when it comes to solving problems within their mutual scope,
this restriction does not actually distinguish between the approaches.

4
1.1.2 The traditional Wigner-Eckart theorem
The traditional Wigner-Eckart theorem applies to quantum-mechanical systems with
spherical symmetry. This includes any sufficiently isolated quantum-mechanical sys-
tem: protons, neutrons, nuclei, atoms, and molecules (Rose 1957, 22-23). For these
systems, the Hamiltonian operator is invariant under members of the compact Lie group
SO(3)—the special orthogonal group of degree three—identical to the rotation group in
three-dimensional space. These systems possess a total angular momentum j(j + 1), j an
integer, that is a constant of the motion. More generally, the relevant symmetry group
is the compact Lie group SU (2), the special unitary group of degree two. Systems whose
Hamiltonian is invariant under members of SU (2) possess a total angular momentum
j(j + 1) where j is either an integer or half-integer.
The group-theoretic viewpoint leads naturally to a general Wigner-Eckart theorem
considered in Section 1.2.2. However, to simplify the exposition, I mostly eschew group-
theoretic language. In the history of atomic spectra, this strategy has often been adopted
for pedagogic reasons, although it was not perfected until Racah’s (1942) algebraic deriva-
tion of the Wigner-Eckart theorem, discussed in Section 1.2.1. Among its many applica-
tions, the traditional Wigner-Eckart theorem is used to compute matrix elements relevant
for the fine structure and magnetic and electric hyperfine structures in atoms (cf. e.g.
Tinkham 1964, Chapter 6).
For isolated quantum-mechanical systems, it is preferable to work in a standard angu-
lar momentum basis: {|αjmi}. This basis parameterizes kets by two angular momentum
quantum numbers, j and m. To construct this basis, we determine simultaneous eigenvec-
tors of the square of the total angular momentum operator, J 2 , and one of the components
of the total angular momentum operator J. Conventionally, the z-component of J is cho-
sen: Jz . j and m parameterize the eigenvalues of J 2 and Jz , respectively. Finally, the
parameter α corresponds to the eigenvalues of a set of operators A that, together with
the operators J 2 and Jz , form a complete set of commuting observables. The operators
in A vary from system to system, so α serves as a generic parameter.
We can construct an angular momentum basis for any space on which J 2 provides a
good quantum number, i.e. a conserved quantity that labels stationary states. Physically,
this requires that the system be sufficiently isolated from external perturbations so that
its energy is invariant under arbitrary rotations in space. Isolated atoms, nuclei, and
molecules all satisfy this requirement, so adopting an angular momentum basis requires
only a small loss of generality. m, the magnetic quantum number, ranges over the 2j + 1
values −j, −j + 1, . . . , j − 1, j. In an isolated system, states with the same values of j
and α have the same energy. In a magnetic field, these states split into 2j + 1 values
distinguished by their different m-values, giving rise to the Zeeman effect and the name
“magnetic quantum number.”
For isolated systems, we can also construct irreducible tensor operators T k . These
provide angular momentum bases for operators. For a rank k tensor, k an integer, an
irreducible tensor operator T k consists of 2k + 1 components Tqk . q, analogously to m,
ranges over the 2k + 1 values −k, −k + 1, . . . , k − 1, k. These Tqk components form an
angular momentum basis. Although using an angular momentum basis of irreducible
tensor operators seems restrictive, it actually presents no loss of generality. Despite the
fact that not all quantum-mechanical operators are irreducible tensor operators, all phys-
ically well-defined operators are expressible as direct sums of irreducible tensor operators.
To be well-defined, an operator’s expectation values must be invariant under arbitrary

5
rotations in space. This makes a well-defined operator a tensor operator : it transforms
appropriately under rotations. Since the irreducible tensor operators form a basis for the
space of tensor operators, all operators are expressible in this basis.
The Wigner-Eckart theorem applies to
matrix
elements
of irreducible tensor operators
0 0 0 k
in an angular momentum basis of kets: α j m Tq αjm . The theorem states that each

0 0 k
matrix element equals a reduced matrix element α j T αj times a Clebsch-Gordan

coefficient hj 0 m0 | jkmqi:

0 0 0 k
α j m Tq αjm = α0 j 0 T k αj hj 0 m0 | jkmqi


Wigner-Eckart theorem: (1.2)

The reduced matrix element α0 j 0 T k αj is in fact not an operator matrix element.

But it behaves like one under changes in basis (Biedenharn and Louck 1981a, 97). Al-
though closed-form expressions for reduced matrix elements exist,
in practice
we compute
0 0 0 k

a reduced matrix element by dividing a known matrix element α j m Tq αjm by the
relevant Clebsch-Gordan coefficient hj 0 m0 | jkmqi.
Computing the reduced matrix element via a matrix element initially seems the wrong
way around: isn’t the Wigner-Eckart theorem supposed to help us compute matrix ele-
ments? Nevertheless, this procedure still provides an overall reduction in computation.
As discussed in the next chapter, this computational reduction results from a separation
of degrees of freedom. Whereas the Clebsch-Gordan coefficient hj 0 m0 | jkmqi
0 depends
on
0 0 k
the magnetic quantum numbers m, q, and m , the reduced matrix element α j T αj
is independent of these parameters. Thus, all matrix elements
0 with
the same param-
0 0 0 k
eters α, α , j and j share the same reduced matrix element α j T αj . There are
(2j 0 + 1) × (2k + 1) × (2j + 1) such matrix elements, due to the corresponding ranges of
the magnetic quantum numbers m, q, and m0 .
If we use the elementary approach, then we must compute each of these matrix
elements via an inner product. Fortunately, we can reduce this to calculating only
(2k + 1) × (2j + 1) inner products by applying the selection rule m + q = m0 . Ma-
trix elements that do not satisfy this selection rule, i.e. where m0 does not equal m + q,
necessarily equal zero. But if instead we use the Wigner-Eckart theorem, then it is nec-
essary to compute only one matrix element via an inner product, provided the one we
choose is nonzero. We then use this matrix element to determine the reduced matrix
element. By sequentially multiplying this reduced matrix element with each Clebsch-
Gordan coefficient, we obtain the remaining [(2k + 1) × (2j + 1)] − 1 matrix elements.
Section 2.2 elaborates this computational procedure, while Section 2.3 explains why it is
conceptually significant.
Although there are many ways to define Clebsch-Gordan coefficients, the quantum-
mechanical addition of angular momenta provides an elementary definition. To add
angular momenta, we form the tensor product of two Hilbert spaces on which angular
momentum bases are defined: H1 ⊗ H2 ≡ H. Here, H1 is a Hilbert space of kets {|jmi},
and H2 is a Hilbert space of irreducible tensor operators {Tqk }. This latter Hilbert space
behaves like a standard angular momentum space, a nontrivial result first shown by
Eckart (1930) (Biedenharn and Louck 1981a, 95).
Two prominent standard angular momentum bases exist for the tensor product space
H: the uncoupled and coupled bases. The uncoupled basis comprises tensor products of
basis kets of H1 and H2 : |jmi ⊗ |kqi. Generally, physicists omit the symbol “⊗,” writing
this product as |jmi |kqi or |jkmqi. I adopt the latter notation for convenience. Fur-
thermore, to preserve continuity with standard treatments of angular momenta, I abuse

6
notation, writing |kqi for the irreducible tensor component Tqk .
Alternatively, we can view H as an angular momentum space in its own right and
construct a coupled angular momentum basis {|j 0 m0 i}. As usual, j 0 and m0 parameterize
the eigenvalues of the operators J 02 and Jz0 , respectively. As follows from the standard
theory of angular momentum for the groups SO(3) and SU (2), m0 = m + q, and j 0 is
a member of the Clebsch-Gordan series {j + k, j + k − 1, j + k − 2, · · · , |j − k|} (Rose
1957, 32ff.). Finally, we define the Clebsch-Gordan coefficients as the matrix elements of
a unitary transformation C from the coupled basis to the uncoupled basis:
j+k j 0
X X
|jkmqi = |j 0 m0 i hj 0 m0 |jkmqi (1.3)
j 0 =|j−k| m0 =−j 0

As the notation hj 0 m0 | jkmqi emphasizes, Clebsch-Gordan coefficients are also the inner
product of a basis ket from the coupled representation with a basis ket from the uncou-
pled representation. As the elements of a unitary matrix, the Clebsch-Gordan coefficients
are unique up to a phase factor. Standard phase conventions make the coefficients real.

1.2 An Historical Overview


To illustrate the importance of the Wigner-Eckart theorem for spectroscopy, I now pro-
vide a potted intellectual history of its genesis in atomic spectroscopy and extensions
to other branches of physics and chemistry. The history of the Wigner-Eckart theorem
largely follows the history of theoretical developments in atomic, nuclear, and molecular
spectroscopy. The secondary literature consists primarily of historical accounts and brief
remarks written by physicists, chemists, and mathematicians who worked in these areas.1
Unfortunately, the Wigner-Eckart theorem has received little attention from professional
historians of science. Nevertheless, I do not provide a thorough history here. Rather
than striving for historical completeness, I aim to situate and motivate the philosophical
argument of Chapter 2.
This brief history performs two functions. First, it shows that the Wigner-Eckart
theorem is central to the application of group theory to models of atoms, nuclei, and
molecules. Consequently, any thorough philosophical account of symmetry principles
in physics and chemistry must necessarily engage with it. Second, the history of the
Wigner-Eckart theorem shows that the philosophical puzzle of Chapter 2 is not restricted
to atomic spectroscopy. Instead, it arises in a wide variety of different theoretical con-
texts. Although I construct my argument using the atomic case, it extends, mutatis
mutandis, to this wider context.

1.2.1 Genesis of the Wigner-Eckart theorem


Unlike many names of theorems, the appellation “Wigner-Eckart” is historically accu-
rate. One of the theorem’s underlying ideas first appeared in a landmark paper by
Wigner (1927, 645)—the first paper to apply group representation theory to the rota-
tional symmetry of atoms (Biedenharn and van Dam 1965, 6). The first statement of
1
cf. e.g. Mackey (1981); Judd (1985); Zeldes (2009); Kibler (2011).

7
what we now call the Wigner-Eckart theorem occurs in a review article by Carl Eckart
(1930, 358), where he presented applications of group representation theory to atoms.
Eckart stated a special case of the traditional Wigner-Eckart theorem for vector opera-
tors, such as the electric dipole operator −ex, which transform as vectors under rotations
in three-dimensional space. In his 1931 monograph on applications of group theory to
atomic spectra, Wigner was the first to state the Wigner-Eckart theorem for arbitrary
tensor operators transforming under SU (2). However, it is arguable that in Eckart’s pre-
sentation, “the general result is clearly implied” (Biedenharn and Louck 1981a, 95).
Whereas Wigner’s proof relies heavily on the representation theory of SO(3) and
SU (2), physicists developed an alternative approach that avoids representation theory.
Implicitly, these physicists took advantage of the mathematical relationship existing be-
tween the Lie group SO(3) and its Lie algebra so(3), which is isomorphic to the Lie
algebra su(2) of the Lie group SU (2) (Mackey 1981, xxxviii). Using this Lie algebra,
one works with quantum-mechanical operators and their commutation relations, which
were by 1930 much more familiar to physicists than group representation theory. Largely
for this pedagogic reason, Condon and Shortley’s monumental Theory of Atomic Spectra
(1935) does not provide Wigner’s version of the Wigner-Eckart theorem for arbitrary ten-
sor operators. Instead, Condon and Shortley derived the special case for vector operators
originally given by Eckart, which was sufficient for providing a thorough account of atomic
spectra (1935, 63). In 1942, Racah extended Condon and Shortley’s non-representation-
theoretic approach to derive the general form of the Wigner-Eckart theorem given by
Wigner.2 Additionally, Racah further developed the algebra of tensor operators, vastly
extending its applications to the spectra of complex atoms.

1.2.2 Applications to atoms, nuclei, and molecules


To convey a fuller sense of the theorem’s scope, I review various directions that physicists,
chemical physicists, and chemists explored after early applications in atomic spectra.
Many of these applications involve generalizations of the Wigner-Eckart theorem beyond
the groups SO(3) and SU (2). Group-theoretic proofs of the traditional Wigner-Eckart
theorem easily generalize for arbitrary finite groups and compact Lie groups (Cornwell
1984, 113). In a series of papers in the 1970s, the mathematician A.U. Klimyk further
generalized the Wigner-Eckart theorem to any non-compact, semi-simple Lie group and
even non-semi-simple, non-compact Lie groups (Cornwell 1984, 108-109). Consequently,
the Wigner-Eckart theorem applies to almost every group that has ever found application
in physics and chemistry.
Racah’s fourth paper on the theory of complex spectra (1949) inspired a host of
developments in nuclear, atomic, and molecular physics. In this paper, Racah used what
we now call a chain of groups to label the valence electrons of lanthanides and actinides
(Judd 1963, 106)3 :

GL(7) ⊃ U (7) ⊃ SU (7) ⊃ O(7) ⊃ G2 ⊃ SO(3) (1.4)


In lanthanides and actinides, valence electrons occupy an f -shell. Collectively, the groups
in the chain 1.4 provide quantum numbers for uniquely labeling these states. The largest
2
In a separate project, I am comparing conceptual differences between Lie-group-theoretic derivations
in the tradition of Wigner and Lie-algebraic derivations in the tradition of Racah.
3
G2 is the compact real form of one of the five exceptional simple Lie groups.

8
group, GL(7)—the general linear group of degree seven over the real numbers—serves
as a supergroup: together, the atomic f -shell states form a basis of a single irreducible
representation of GL(7) (Wybourne 1974, 349).4 After labeling these states, the next
step is to express operators of interest via irreducible tensor operators transforming un-
der these groups. We can then apply the Wigner-Eckart theorem to each group in this
chain to calculate matrix elements (Judd 1963, 196; Wybourne 1970, 94). Here, we apply
the Wigner-Eckart theorem six times, once for each group. Each matrix element is then
expressed as a finite sum, where each summand is a product of a reduced matrix element
and six Clebsch-Gordan coefficients, one for each group.5 Like the traditional reduced
matrix element, this reduced matrix element is independent of the quantum numbers
associated with each subgroup.
Although Racah developed his approach for atoms, the chemical physicist Brian
Wybourne has noted that,“atomic spectroscopists were surprisingly slow at realizing
the full significance of Racah’s work and much of the further development was made
by nuclear physicists, notably Jahn, Elliott, and Flowers” (1970, vii). In nuclear physics,
Racah’s approach was quickly applied to the nuclear shell model. Many of these early ap-
plications rely on the traditional Wigner-Eckart theorem for SU (2) (de-Shalit and Talmi
1955, 117). The Wigner-Eckart theorem was also extended to states and operators defined
in isospin space (Brussaard and Glaudemans 1977, 412). Later developments include the
interacting boson model introduced by Arima and Iachello in 1974, which uses the unitary
Lie algebra u(6) (Frank and Isacker 1994, 295, 298). Again, the Wigner-Eckart theorem
is useful for calculating matrix elements in this model (Frank and Isacker 1994, 375).
By 1960, chemists had adapted the methods of Wigner and Racah to treat molecules.
John S. Griffith published foundational work, culminating in his 1962 monograph The
Irreducible Tensor Method for Molecular Symmetry Groups. Griffith provides a general
Wigner-Eckart theorem applicable to the finite symmetry groups of molecules (1962, 21).
A number of applications follow, such as calculating electron transition intensities for
metal ions in an octahedral ligand field (1962, 96). As mentioned in Section 1.1.2, the
traditional Wigner-Eckart theorem also remains relevant for molecules (Griffith 1962,
77).
By the 1970s, chemical physicists were applying Lie groups to ligand field theory.
Brian Wybourne was one of the main innovators, applying supergroups to label the elec-
tron shells of transition metals in ligand fields (Wybourne 1972, 365). The group structure
is determined through an algebra of creation and annihilation operators—already familiar
from their use in quantum field theory (Wybourne 1973b, 1120). For instance, one can use
the supergroup SO(9) to label the states of the e-shell for a transition metal complex in an
octahedral ligand field. We obtain a complete set of quantum numbers for the states by
determining the relevant subgroup structure of the supergroup (Wybourne 1973b, 1122),
which in this case is given by the chain SO(9) ⊃ SO(8) ⊃ SU (4) ⊃ SU (2)⊗Oh , where Oh
is the octahedral group used in elementary ligand field theory (Wybourne 1972, 366). As
in the atomic case, one advantage of the supergroup approach is that the Wigner-Eckart
theorem applies to the entire chain (Wybourne 1972, 367).
In molecular physics, Iachello and Levine developed the “vibron model,” which uses
4
A matrix representation, Γr , is a map from elements of the group to invertible matrices that preserves
the multiplicative structure of the group. In other words, Γr is a group homomorphism from the group
to matrices. A representation Γj is irreducible if it cannot be decomposed into a direct sum of lower
dimensional, nontrivial representations.
5
The summation stems from the general Wigner-Eckart theorem (Cornwell 1984, 108).

9
the unitary Lie algebra u(4) to label the vibrational and rotational states of molecules
(1995, 26; Wulfman 2011, 401). This model enables the calculation of transition prob-
abilities for emitted and absorbed radiation due to changes in molecular states. This,
in turn, enables the prediction and interpretation of vibrational and rotational spectra
for diatomic and polyatomic molecules. As in the previous examples, the Wigner-Eckart
theorem facilitates matrix element computations (Iachello and Levine 1995, 11).
Despite these varied applications, the practical significance of the Wigner-Eckart the-
orem has not always been recognized. In a wide-ranging review of applied group theory
up to 1973, Wybourne argued that many of his colleagues failed to appreciate the signif-
icance of the Wigner-Eckart theorem and, in doing so, misunderstood the significance of
applied group theory:

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 1973a, 39)

As this section has shown, the Wigner-Eckart theorem is immensely useful, both in its
traditional and generalized forms. The philosophical puzzle of Chapter 2 arises because—
despite its practical importance—the Wigner-Eckart theorem is not necessary for comput-
ing matrix elements. Before developing this puzzle, I situate the Wigner-Eckart theorem
in a broader philosophical landscape.

1.3 Connections and Inspirations


The preceding sections have aimed to show the importance of the Wigner-Eckart theo-
rem and motivate it as a worthy candidate for extended philosophical analysis. I now
characterize the shape that my philosophical analysis will take. To this end, it is helpful
to compare the Wigner-Eckart theorem with formulations of classical mechanics, which
Butterfield (2004a) refers to as schemes. Butterfield (2004a, 2004b, 2006) extensively
discusses three: the Lagrangian, Hamiltonian, and Hamilton-Jacobi schemes. Like my
focus here, Butterfield argues that philosophers should focus on these schemes’ “various
merits for solving problems” (2004a, 9-10). The puzzle for these schemes is much the
same as the puzzle for the Wigner-Eckart theorem and its foil, the elementary approach:
restricted to certain classes of problems, these schemes are equally powerful. Yet, they
appear to be conceptually distinct. What accounts for these differences? In this section,
I consider three topics that might be suggested for answers: laws of nature vs. models,
idealization vs. approximation, and scientific explanation. I argue that, in fact, none of
these topics provides a promising way forward.

1.3.1 Schemes, laws of nature, and models


Throughout, I adopt terminology and insights from Butterfield’s investigation of classi-
cal mechanics. Much of Butterfield’s motivation comes from recognizing that the three
most well-established schemes for classical mechanics—Lagrangian, Hamiltonian, and

10
Hamilton-Jacobi—fall “between two topics often emphasized by philosophers, ‘laws of
nature’ and ‘models’” (2004a, 10). Butterfield’s appraisal emphasizes that these classical-
mechanics schemes are conceptual frameworks for solving problems, rather than specific
models. Although they are too general to be models, these schemes are not sufficiently
general to be laws of nature; they include formulations of the laws of classical mechanics
but are not just a statement of those laws (Butterfield 2004a, 7).
Like these schemes, it is doubtful whether the Wigner-Eckart theorem deserves the
honorific title “law of nature.” The theorem is restricted to a particular class of models,
and indeed, a particular kind of basis. Nevertheless, due to the complexity of the philo-
sophical debate over laws of nature (Earman 2004, 1228), I am reluctant to tether my
interpretation to a particular account of laws. Among many other positions, David Arm-
strong (1983) characterizes a law as a necessary relationship between universals, while
David Lewis (1973) argues that a law is an axiom or theorem of a best deductive system.
To make my conclusions more robust, I develop my account in Chapter 2 without mak-
ing any assumptions concerning lawhood: there is much to say about the Wigner-Eckart
theorem independently of whether or not it is a law of nature. Nonetheless, accounts of
laws have important implications for my project. For instance, lawhood would provide
an easy answer to why the Wigner-Eckart theorem is conceptually significant: it is a law.
Thus, if the Wigner-Eckart theorem is a law, then much of my analysis in Chapter 2 may
seem superfluous. But regardless, my detailed discussion leads to a general theory of con-
ceptual significance; this remains philosophically interesting even if the Wigner-Eckart
theorem is a law.
Returning to Butterfield’s analysis of classical mechanics, the Wigner-Eckart theorem
is not aptly classified as a scheme in the grand sense of his chosen formulations of mechan-
ics. In this sense, a scheme provides broad-ranging methods for solving central problems
of interest. For instance, the Lagrangian scheme applies to most problems in classical
mechanics. In contrast, the Wigner-Eckart theorem is of much narrower scope. The
analogous scheme that contains the theorem is the Racah-Wigner algebra. This scheme
provides a general framework for manipulating irreducible tensor operators (Biedenharn
and Louck 1981a, b). Rather than through schemes, the Wigner-Eckart theorem is best
described using what Batterman and Rice call a “minimal model” (2014, 358). A mini-
mal model is a submodel shared by a large class of more specific models, which augment
the minimal model by adding details. The Wigner-Eckart theorem is a consequence of
the following minimal model: a quantum-mechanical Hamiltonian that is invariant under
arbitrary rotations. Thus, any more specific quantum-mechanical model whose Hamilto-
nian is rotationally invariant falls within the scope of the Wigner-Eckart theorem.
Many more aspects of Racah-Wigner algebras could be treated along the lines of
Chapter 2. I restrict attention to the Wigner-Eckart theorem for clarity: this theorem is
by itself rich enough to illustrate a general philosophical problem. Furthermore, it exem-
plifies the general strategy of Racah-Wigner algebras: exploit the symmetry properties
of tensor operators.

1.3.2 Idealization and approximation


Despite not being a scheme, the Wigner-Eckart theorem bears the same relationship to
idealization and approximation as formulations of mechanics do. Namely, it relies on
idealizations, rather than approximations. Following Butterfield (2004a), I take this dis-

11
tinction to mark whether simplification occurs (i) before applying a model (idealization)
or (ii) in the course of solving a well-defined problem (approximation). As Butterfield
puts it, “idealization is neglecting what you believe negligible when you first pose a
problem, while approximation is doing so while solving it” (2004a, 24-25). The key sim-
plification in the Wigner-Eckart theorem is the assumption of spherical symmetry for
sufficiently isolated systems. This assumption occurs before solving any specific problem
and is therefore an idealization.
Although philosophers debate whether certain methods in statistical mechanics in-
volve idealizations (Batterman 2002, Norton 2012), this debate does not affect interpre-
tations of the Wigner-Eckart theorem. By an “idealization,” Norton means a “real or
fictitious system, distinct from the target system, some of whose properties provide an
inexact description of some aspects of the target system” (2012, 209). To apply the
Wigner-Eckart theorem, we refer to a truly isolated quantum-mechanical system; this
provides an inexact description of the target system because it is not truly isolated. Sim-
ilarly, Batterman uses “idealization” to mean a system that does not exist in the physical
world (2010, 10). This also fits the Wigner-Eckart theorem insofar as it is impossible
to have a truly isolated quantum-mechanical system. Thus, this debate—which mainly
arises in the context of infinite systems—is not relevant for interpreting the Wigner-Eckart
theorem.
Additionally, various philosophical problems about idealization are not germane to
this case study. Although the Wigner-Eckart theorem relies on an idealization of spher-
ical symmetry, applications of the elementary approach make the same idealization. In
practice, we apply the elementary approach to compute matrix elements in a standard
angular momentum basis. As described in Section 1.1.2, using this basis involves an
assumption of spherical symmetry. Since both approaches make the same idealization,
idealization does not distinguish them. Thus, I will not consider why this idealization is
so successful, whether its success warrants commitment to spherically symmetric systems,
or similar questions.

1.3.3 Scientific explanation


Finally, I distinguish my topic from another central philosophical topic: scientific ex-
planation. As in the case of idealization, I claim that explanatory differences do not
distinguish between the elementary approach and the Wigner-Eckart theorem. If our
explanandum is an atomic transition intensity, then both approaches explain equally
well—insofar as computing the appropriate matrix element is explanatory. More pre-
cisely, both approaches satisfy Hempel and Oppenheim’s deductive-nomological account
of explanation (1965 [1949]): from the laws of quantum mechanics and initial conditions
regarding atomic states, they deduce atomic transition intensities.
However, there is a sense in which the Wigner-Eckart theorem explains more than
the elementary approach: it unifies the matrix elements of irreducible tensor operators
via Clebsch-Gordan coefficients and reduced matrix elements. This invites an analysis
using Philip Kitcher’s account of explanatory unification (1989). In this section, I explain
why Kitcher’s account of unification provides no resources for distinguishing between the
elementary approach and the Wigner-Eckart theorem. To do so, I consider two potential
explanatory differences between the approaches. These differences arise if we modify the
explanandum from a specific transition intensity to patterns that these intensities display.

12
Although I spend more time discussing the first alternative explanandum, we will see that
it is contrived. The second is a big topic—selection rules—but it is more easily addressed.
The first potential explanatory difference involves the following “explanandum:” for
an irreducible tensor operator T k , the ratio of two matrix elements with fixed parameters
α, α0 , j, and j 0 equals a ratio of Clebsch-Gordan coefficients. I use scare-quotes because it
is unclear that this fact should count as an explanandum. Regardless, the Wigner-Eckart
theorem entails this fact: this ratio of matrix elements contains the same reduced matrix
element in the numerator and denominator, which cancels, leaving a ratio of Clebsch-
Gordan coefficients. Although the elementary approach can demonstrate this fact for
particular cases, it cannot provide a deductive explanation that holds for arbitrary ma-
trix elements.
At first glance, Kitcher’s account of unification seems to capture this difference. Ac-
cording to Kitcher, an argument is explanatory if and only if it belongs to the explana-
tory store—the set of derivations that unifies the set of accepted scientific beliefs. For
Kitcher, an explanatory store unifies if it “makes the best trade-off between minimizing
the number of patterns of derivation employed and maximizing the number of conclusions
generated” (1989, 432). Arguably, the Wigner-Eckart theorem belongs in the explanatory
store. Although using the theorem introduces new argument patterns—namely, the com-
putation of Clebsch-Gordan coefficients and reduced matrix elements—it provides new
conclusions. These include the above “explanandum,” selection rules (discussed below),
and of course the bare content of the theorem: matrix elements factor into a product of
a reduced matrix element and a Clebsch-Gordan coefficient. Thus, the Wigner-Eckart
theorem provides a fair trade-off between new argument patterns and new conclusions.
Nonetheless, Kitcher’s account of unification suffers from a serious difficulty in this
case. Although it sensibly decrees that the Wigner-Eckart theorem unifies, it is commit-
ted to saying that the elementary approach unifies as well. This is because the elementary
approach is a constituent argument pattern of the Wigner-Eckart theorem: to apply the
Wigner-Eckart theorem, we rely on the elementary approach to compute an initial ma-
trix element. Thus, if the Wigner-Eckart theorem belongs to the explanatory store, then
the elementary approach necessarily belongs as well. As a result, Kitcher’s account of
unification fails to distinguish between the elementary approach and the Wigner-Eckart
theorem.
Indeed, to distinguish the approaches on the basis of Kitcherian unification, only one
approach could belong to the explanatory store. This leaves only one option: includ-
ing the elementary approach in the explanatory store while excluding the Wigner-Eckart
theorem. In this case, the elementary approach would unify and thereby explain, while
the Wigner-Eckart theorem would fail to be explanatory. However, this verdict neglects
the sense in which the Wigner-Eckart theorem unifies matrix elements. Section 2.2 dis-
cusses this sense in terms of generality: the same Clebsch-Gordan coefficients and reduced
matrix elements apply to many different matrix element problems. As we will see in Sec-
tion 2.2, generality captures much of the spirit of Kitcherian unification.
A related feature of Kitcher’s account that I preserve is his emphasis on the num-
ber of conclusions an argument pattern generates. Section 2.3 discusses this insight in
terms of the information that a solution procedure provides. The Wigner-Eckart the-
orem provides more information about matrix elements than the elementary approach
does. This difference is essential for understanding why the Wigner-Eckart theorem is
conceptually significant. One example of additional information is the “explanandum”
introduced above. However, as an explanandum, this fact is highly contrived. In practice,

13
spectroscopists are not concerned with explaining this fact. Indeed, it is doubtful that
one would have recognized this fact without having already proven the Wigner-Eckart
theorem. It is perhaps more aptly termed a prediction that the theorem makes. This
example indicates that accounts of scientific explanation are ill-suited for characterizing
some conceptually significant differences between solution procedures.
I turn now to a second potential explanatory difference, involving the explanandum
of selection rules. Selection rules provide general criteria for non-zero matrix elements:
if two states do not satisfy all relevant selection rules for a given perturbation, then the
matrix element between them necessarily vanishes, i.e. equals zero. If the matrix ele-
ment between two states vanishes, then a transition does not occur between those states
via that perturbation operator. One common application of the Wigner-Eckart theorem
is to derive selection rules for angular momentum quantum numbers (Cornwell 1984,
135). However, as in the case for computing matrix elements, it is not necessary to use
the Wigner-Eckart theorem to derive selection rules. A general approach exists that is
more direct than the Wigner-Eckart theorem; this approach relies on computing Clebsch-
Gordan series for the groups whose irreducible representations provide quantum numbers
(Hamermesh 1962, 166ff.; Smith and Wybourne 1967). Since this approach is distinct
from the Wigner-Eckart theorem, it is possible to append it to the elementary approach,
giving, in other words, an elementary approach for deriving selection rules. This leads to
a philosophical problem analogous to that of Chapter 2: is using the Wigner-Eckart theo-
rem to derive selection rules conceptually significant, or merely a matter of convenience?
Hence, considering selection rules in terms of explanation does not solve this overarching
philosophical problem. In the next chapter, I show that more appropriate resources exist
for tackling these questions than those supplied by philosophical accounts of explanation.
Before moving on, I consider one more related topic: scientific understanding. The dif-
ferences examined in Chapter 2 between the elementary approach and the Wigner-Eckart
theorem deal fundamentally with the viewpoint one adopts in solving a problem. Recog-
nizing this, it seems natural to discuss these differences in terms of understanding. For
instance, we can characterize the Wigner-Eckart theorem as providing a different under-
standing of the matrix element problem. Similarly, different classical-mechanics schemes
provide different understandings of many of the same problems and solutions. These dif-
ferences in problem-solving procedures arise from differences in expressive means, a term
I define in Section 2.1. Along these lines, Manders has explicitly connected expressive
means with understanding in mathematics (unpublished). As developed there and in this
essay, this account of expressive means does not depend on cognitive features of agents.
This is in contrast with recent literature on scientific understanding that considers un-
derstanding as a mental state of agents (Dieks and de Regt 2005, de Regt 2009, Chang
2009, Khalifa 2012). Although we commonly use “understanding” to reference a partic-
ular kind of mental state, there is much to be said about understanding that does not
depend on features of agents. By analyzing differences in expressive means, it is possible
to distinguish the conceptual content of problem-solving procedures without recourse to
mental states. Nevertheless, to avoid the psychological connotations of “understanding,”
I will instead develop an account of “conceptual significance,” discussed in Section 2.1.

14
Chapter 2

The Methodological and Conceptual


Significance of the Wigner-Eckart
Theorem

This chapter addresses the following question: what makes a difference between theoreti-
cally equivalent expressive means conceptually significant, rather than merely convenient?
To sharpen this question, four terms need to be unpacked: theoretical equivalence, expres-
sive means, conceptual significance, and convenience. Section 2.1 clarifies these terms.
As stated, this guiding question goes far beyond the particulars of the Wigner-Eckart the-
orem and other symmetry arguments. Nevertheless, the theorem thoroughly illuminates
this question, and it supports a tentative answer. Section 2.2 uses the Wigner-Eckart
theorem to illustrate how a change in expressive means can be convenient for solving
a problem. However, I contend that convenience does not constitute conceptual signifi-
cance. In Section 2.3, I argue that the Wigner-Eckart theorem is conceptually significant
because it provides a novel restraint: it controls the form of solutions to matrix element
problems.1 This restraint leads to novel functional roles for the expressive means of the
Wigner-Eckart theorem, which support problem-solving strategies that are unavailable to
the elementary approach. These functional roles lead to the methodological advantages
described in Section 2.2, providing a connection between conceptual significance and con-
venience. Finally, Section 2.4 proposes a tentative answer to this chapter’s overarching
question: a theoretically equivalent change in expressive means is conceptually signifi-
cant if and only if it leads to novel functional roles for solving a problem. I motivate this
proposal with a number of brief examples.

2.1 A Puzzle: Theoretical Equivalence, Expressive


Means, and Convenience
The Wigner-Eckart theorem exemplifies a puzzle that applies to many theoretical ad-
vances. It is an example of what Butterfield calls “theory development without an in-
crease in logical strength” (2004a, 28). Speaking broadly, Butterfield takes this aspect of
scientific theories to be “old news:”

That is, it is not controversial that providing equivalent formulations of a


1
I avoid using the related word “constraint” due to its technical meaning in mechanics.

15
theory can be very significant, both methodologically and ontologically. For
one of a pair of theoretically equivalent formulations might extend better than
the other to another domain of phenomena; or deal better, in some sense,
with the given domain. This can even be so when theoretical equivalence is
construed strongly enough that theoretical equivalence implies, or near enough
implies, that the two formulations have the same ontology. (2004a, 29)

The Wigner-Eckart theorem and the elementary approach fall into this latter class of
theoretical equivalence: they posit the same ontology. The ontology for both approaches
is determined by choosing an interpretation of quantum mechanics (Bell 1987). Fur-
thermore, they cover the same range of phenomena, so a difference in fruitfulness, or
problem-solving power, also does not distinguish them.2
These observations motivate a definition for theoretical equivalence: two problem-
solving approaches are theoretically equivalent if (i) they share the same ontology and
(ii) they solve the same class of problems. The fact that this definition is tailor-made
for analyzing the Wigner-Eckart theorem should not trouble us. For it places stringent
conditions on any philosophical account of how the Wigner-Eckart theorem—and similar
cases of theoretical equivalence—acquire conceptual significance. (i): Ontological equiva-
lence means that no appeal to an ontological difference can distinguish the Wigner-Eckart
theorem from the elementary approach. This restriction does not apply to theoretical re-
formulations that have ontological differences, such as various interpretations—as opposed
to formulations—of quantum mechanics. (ii): Furthermore, since the approaches solve
the same class of problems, no obvious epistemic difference immediately distinguishes
them, although I argue in Section 2.3 that there is a more subtle kind of epistemic differ-
ence at play. Problem-solving equivalence forestalls an easy answer for why theoretically
equivalent problem-solving frameworks are conceptually distinct. In contrast, an easy
answer holds for many theoretically inequivalent approaches, such as calculus-based me-
chanics vs. purely algebraic approaches. These inequivalent approaches are conceptually
distinct because they solve different classes of problems.
Despite being uncontroversial, the significance of theoretically equivalent formulations
remains puzzling. Since the Wigner-Eckart theorem neither posits a new ontology nor
makes previously insoluble problems soluble, why is it significant? Why is it not a philo-
sophically uninteresting convenience? To make headway on these questions, it is helpful
to focus on “expressive means,” an idea introduced by Kenneth Manders in the philoso-
phy of mathematics (2008a, b; unpublished). Expressive means comprise the notational,
diagrammatic, and conceptual resources that a problem-solving approach makes avail-
able. The Wigner-Eckart theorem uses the same expressive means as the elementary
approach, except for a few additions. These include Clebsch-Gordan coefficients and re-
duced matrix elements, which do not figure in the elementary approach. Prima facie,
any conceptual difference between these two approaches should arise from differences in
expressive means. However, explaining the conceptual differences between the Wigner-
Eckart theorem and the elementary approach is not as simple as identifying differences
in expressive means. Complications arise because not all changes in expressive means are
conceptually significant. For instance, substituting the symbol “y” for the symbol “x” in
the expression x2 −1 is a trivial change in expressive means. It is an example of what I will
call a trivial notational variant. With this in mind, we can rephrase the guiding question
2
This is true provided we append a procedure for deriving selection rules to the elementary approach,
as discussed in Section 1.3.

16
of this chapter: what differences between theoretically equivalent expressive means are
conceptually significant?
To begin answering this question, I consider a pre-theoretical notion of conceptual
significance. Ideally, we desire an account of conceptual significance that distinguishes
trivial from nontrivial notational variants. Throughout, I focus on variants that are
theoretically equivalent, since this is the most puzzling case. Although “trivial vs. non-
trivial” points to a dichotomy, we should not expect differences in expressive means to be
so simple. Instead, we expect expressive differences to lie on a spectrum of significance,
ranging from the triviality of symbol-substitutions to prima facie nontrivial differences
such as the Wigner-Eckart theorem. In between, we can expect alternative expressive
means such as different coordinate systems and some diagrammatic techniques. Much of
my task in this chapter is to articulate a philosophical account of conceptual significance.
I will ground this account on concrete notions such as the restraints and functional roles
that expressive means provide.
Before developing this account of conceptual significance, I will, in Section 2.2, ar-
ticulate why the Wigner-Eckart theorem is convenient. By convenient, I mean that an
expressive means is easier to use than an alternative. I will describe how, compared to
the elementary approach, the Wigner-Eckart theorem provides three benefits: modular-
ization, generality, and tractability. The theorem modularizes the matrix element problem
into sub-problems, uses more general expressive means, and replaces integrals with a more
tractable multiplication problem. Adopting a phrase that Butterfield uses in passing, I
refer to these benefits as “methodological advantages” (2004a, 29). Collectively, these
three methodological advantages make the Wigner-Eckart theorem more convenient than
the elementary approach. My analysis in Section 2.2 contributes to a small literature on
the methodological importance of symmetry arguments in physics. These methodologi-
cal advantages also arise in examples considered by van Fraassen (1989, 236ff.), Morrison
(2000), and Bangu (2013), although I do not pursue these connections here.
Although convenience underlies the practical importance of the Wigner-Eckart theo-
rem, it is clear that convenience is not sufficient for conceptual significance. There are
many cases where an alternative expressive means is more convenient to use but lacks
any underlying conceptual difference. For example, it is convenient to use the Einstein
summation convention, but this notation plays the same role as a Σ-summation symbol.
Similarly, Dirac bra-ket notation introduces no new functions compared with earlier inner
product notations, although it is more convenient to use. Many technological advances
provide similar examples. It is convenient to use an integrated mass spectrometer and
gas chromatograph. Yet, this device is no more conceptually significant than the two
devices separately. Likewise, increased computing power has no doubt been—in recent
decades—the most significant methodological advance across all of science. Yet, a faster
computer is no more conceptually significant than the code it implements—assuming
that our problem of interest is not building a faster computer. These examples indicate
that convenience is not sufficient for a conceptually significant difference in expressive
means. They also motivate extending Manders’ notion of expressive means to include
some aspects of technology as well, although I will not make this extension precise here.
Recognizing the shortcomings of convenience leads to a sharpened philosophical puz-
zle: what distinguishes conceptually significant differences in expressive means from dif-
ferences that are merely convenient? This puzzle arises for many more examples than the
Wigner-Eckart theorem. It applies to theoretically equivalent formulations of classical
mechanics (Newtonian, Lagrangian, Hamiltonian, etc.), various formulations of quantum

17
mechanics (Schrödinger, Heisenberg, density operators, etc.), and many applications of
symmetry in quantum chemistry, general relativity, and particle physics. Nevertheless,
this puzzle remains unarticulated in discussions of how symmetries, and theoretical re-
formulations more broadly, acquire conceptual significance.
Without having stated this puzzle as such, Butterfield (2004a) nevertheless contributes
to its solution. His insight is that classical-mechanics schemes provide not only solutions
but also information about those solutions: “These schemes give information—indeed,
an amazing amount of information—not just about solutions to individual problems, but
about the structure of the set of solutions to all problems of a large class” (2004a, 25). Fur-
thermore, Butterfield recognizes that “the information a scheme can provide varies,” and
these variations can support nontrivial conceptual differences between schemes (2004a,
25).
Accordingly, Section 2.3 applies this notion of information to the Wigner-Eckart theo-
rem. First, in Section 2.3.1, I present and endorse a standard interpretation of the Wigner-
Eckart theorem. According to this interpretation, the theorem is significant because it
separates rotational and non-rotational degrees of freedom. Then, in Section 2.3.2, I show
how this separation of degrees of freedom is a restraint that the Wigner-Eckart theorem
places on matrix elements. This restraint is a particular kind of information. I show how
this restraint leads to novel functional roles for Clebsch-Gordan coefficients and reduced
matrix elements, in turn leading to the theorem’s methodological advantages. Section 2.4
further articulates this notion of functional role and argues that it provides a general so-
lution to the puzzle.

2.2 Modularization, Generality, and Tractability


The Wigner-Eckart theorem owes its convenience to three methodological advantages:
modularization, generality, and tractability. The theorem modularizes the problem of
computing matrix elements into a series of sub-problems that are either more general
or more tractable than the original problem. The first sub-problem is to compute the
Clebsch-Gordan coefficients. These coefficients apply to any system with coupled angular
momenta, giving them the advantage of generality. The second sub-problem is to com-
pute the reduced matrix element, which is no more difficult than the original problem
but more general. Finally, the third sub-problem involves multiplying a Clebsch-Gordan
coefficient with the reduced matrix element to compute each matrix element of interest.
This multiplication is more tractable than performing integrals with the elementary ap-
proach. By itself, none of these advantages makes solving the matrix element problem
more convenient. Yet, together, they lead to a more convenient solution procedure.
In passing, Butterfield has used the phrase “methodological advantage” to describe
the benefits of different schemes in classical mechanics: “one of the equivalent formu-
lations might have the specific methodological advantage of providing such a scheme”
(2004a, 29). This phrase is a natural one for describing the convenience that symmetries
often provide. It also serves to distinguish the focus of a methodological investigation
from those of an ontological or epistemic character. This difference in focus explains why
I do not adopt the more popular “theoretical virtue.” Theoretical virtues are generally
discussed in the context of inference to the best explanation and the underdetermination
of theory by data (Lipton 2004). However, in the context of this case study, neither of

18
these issues arises. We are not faced with an ontological or epistemological choice between
the elementary approach and the Wigner-Eckart theorem. Instead, we are faced with a
methodological choice as to which approach to apply. Thus, although these methodolog-
ical advantages may strike some readers as theoretical virtues in disguise, they perform
a distinct role. Methodological advantages explain why the Wigner-Eckart theorem is
convenient, and they point to sources of the theorem’s conceptual significance.
As described in Section 1.1.2, the Wigner-Eckart theorem takes a matrix element and
decomposes it into two components: a reduced matrix element and a Clebsch-Gordan
coefficient. This decomposition modularizes the problem of computing matrix elements:
it breaks this problem into sub-problems. The first sub-problem (W1 ) is to compute
the relevant Clebsch-Gordan coefficients. The second sub-problem (W2 ) is to compute
the relevant reduced matrix element for fixed parameters j, j 0 , α, and α0 . Finally, the
third sub-problem (W3 ) is to compute each matrix element of interest by multiplying
the reduced matrix element by the relevant Clebsch-Gordan coefficient; one multiplica-
tion is performed per matrix element. We must solve the sub-problems in this order
because we calculate the reduced matrix element by using a Clebsch-Gordan coefficient.
To summarize, the Wigner-Eckart theorem modularizes the original problem into three
directives: (W1) calculate the Clebsch-Gordan coefficients; (W2) calculate the reduced
matrix element; (W3) calculate the matrix element(s).

2.2.1 Modularization
In his investigation of classical mechanics, Butterfield has considered a methodological
advantage similar to modularization, which he calls “(Separate)” (2004a, 27). (Separate)
describes the ubiquitous procedure of working in variables that de-couple a previously
coupled system of equations:

(Separate): The ability to change to variables that are ‘de-coupled’, in that


one has to solve, either:
(a) ideally, independent rather than coupled equations; or
(b) much more commonly, equations that are coupled less strongly (at least
not pairwise!) to one another than was the originally given set. (Butterfield
2004a, 27)

(Separate) is a specific form of modularization. It captures the central kind of mod-


ularization that symmetries in classical mechanics provide by reducing the operative
degrees of freedom and separating variables (2004a, 26). In spirit, it fits the character
of the Wigner-Eckart theorem, where we “de-couple” matrix elements into two compo-
nents, only one of which depends on the rotational degrees of freedom. This de-coupling
is akin to a separation of variables, which—as in classical mechanics—arises thanks to a
symmetry. However, de-coupling a system of equations and the method of separation of
variables are in detail different procedures than the de-coupling provided by the Wigner-
Eckart theorem. Partly for this reason, I adopt the more general term “modularization.”
By introducing “modularization” as a methodological advantage, I emphasize that
this kind of “divide and conquer” strategy is often advantageous but not necessarily so.
One can imagine breaking a problem into sub-problems that are actually more difficult
than the original. Indeed, as I explain below, this is how the Wigner-Eckart theorem
initially makes things look, due to the tedium of calculating Clebsch-Gordan coefficients.

19
Whether a particular modularization is advantageous depends on the tractability and
generality of it sub-problems. Collectively, these sub-problems should be more tractable
than the original problem. Ideally, at least one sub-problem should also be more general
than a specific case. Generality enables a sub-problem to be solved once and then used
in subsequent applications.
To illustrate these motivating points, I briefly step outside the confines of the Wigner-
Eckart theorem. Discussing modularization in computer science, Jeremy Avigad (forth-
coming) has emphasized the importance of modular programming. By modularizing
a code into separate sub-units, “the components themselves [become] more adaptable,
reusable, and generalizable” (Avigad, forthcoming). These advantages frequently occur
in mathematics as well. In a case study from the history of analytic number theory, Avi-
gad and Rebecca Morris (2014) have documented the importance of modularization for
improving proofs. Tracing the development of proofs of Dirichlet’s theorem, they found
that modularization “makes it easier to read the proof, understand it, and verify its
correctness” (Avigad, forthcoming). Dirichlet’s theorem states that any arithmetic pro-
gression whose first two terms are relatively prime contains an infinite number of prime
numbers. Later proofs rely on a kind of function now known as a “character,” which
Dirichlet’s original proof uses but does not identify for special scrutiny. Mathematicians
made progress in analytic number theory by focusing separately on characters: “devel-
oping their properties independently makes those proofs more modular. Dependencies
between pieces of a mathematical text are minimized, so that one can verify properties
of the characters independently, and then suppress those details later on” (Avigad, forth-
coming). In mathematics generally, we see modularization in the universal strategy of
proving difficult theorems by a succession of lemmas, many of which are themselves of
general utility.

2.2.2 Solving three sub-problems


I return now to the Wigner-Eckart theorem. To understand how modularization can
contribute to convenience, I describe each of the three sub-problems in detail. We will
see that computing the Clebsch-Gordan coefficients (W1) is nontrivial but that the co-
efficients have general utility both within and outside the context of the Wigner-Eckart
theorem. Calculating the reduced matrix element (W2) is as difficult as a single instance
of the original problem, i.e. computing one matrix element. Finally, given steps W1 and
W2, multiplying a Clebsch-Gordan coefficient by the reduced matrix element to obtain
additional matrix elements (W3) is much more tractable than the original problem. After
describing what goes into each step and illustrating them with an example, it will become
clear how they work in tandem to provide convenience.
The most tedious step in applying the Wigner-Eckart theorem is computing Clebsch-
Gordan coefficients. There are at least six distinct methods for calculating these coef-
ficients (Biedenharn and Louck 1981a, 76). As Biedenharn and Louck recognize, these
different approaches are conceptually interesting and “indicative of the broad scope of
interpretations and viewpoints that can be ascribed to the mathematical apparatus of
angular momentum theory” (1981a, 76). Two of the approaches exemplify the group-
theoretic and algebraic approaches to atomic spectroscopy. Wigner developed the group-
theoretic approach first by decomposing a tensor
P product Dj1 (U ) ⊗ Dj2 (U ) of irreducible
representations of SU (2) into a direct sum j ⊕Dj (U ) of irreducible representations.

20
Alternatively, a popular algebraic construction uses the action of raising and lowering
operators on a tensor product of two standard angular momentum bases. Both of these
approaches lead to closed-form solutions for Clebsch-Gordan coefficients. Wigner’s equa-
tion gives a taste of how tedious it is to compute Clebsch-Gordan coefficients from these
solutions (1959, 191):

s
j1 j2 J (J + j1 − j2 )!(J − j1 + j2 )!(j1 + j2 − J)!(J + m1 + m2 )!(J − m1 − m2 )!
Am 1 m2 M
= δM,m1 +m2
(J + j1 + j2 + 1)!(j1 − m1 )!(j1 + m1 )!(j2 − m2 )!(j2 + m2 )!
(2.1)
p
X (−1)κ+j2 +m2 (2J + 1)(J + j2 + m1 − κ)!(j1 − m1 + κ)!
×
κ
(J − j1 + j2 − κ)!(J + m1 + m2 − κ)!κ!(j1 − j2 − m1 − m2 + κ)!
j1 j2 J
Here, Am 1 m2 M
is an alternative notation for the Clebsch-Gordan coefficient hJM | j1 j2 m1 m2 i.
Different approaches to computing Clebsch-Gordan coefficients lead to different closed-
form expressions, but these expressions are related to each other by symmetry transfor-
mations of the coefficients. Less complicated recursion relations enable one to leapfrog
from a known Clebsch-Gordan coefficient to an unknown coefficient.
If one had to calculate the Clebsch-Gordan coefficients anew for each problem, then
this procedure would be manifestly inconvenient. The Clebsch-Gordan calculations are
arguably less tractable than an elementary matrix element computation. Thus, it initially
appears as though the Wigner-Eckart theorem is less convenient than the elementary ap-
proach. However, the generality of the Clebsch-Gordan coefficients makes it worthwhile
to calculate them. The same coefficients are relevant for any problem where one wishes to
decompose a tensor product space described by SU (2) ⊗ SU (2), i.e. any problem where
two spherically symmetric systems are coupled. Thus, the coefficients for a particular ten-
sor product space need to be computed only once, and, in practice, the Clebsch-Gordan
coefficients are found by consulting a table of coefficients. These same coefficients also
apply to other problems stemming from the addition of angular momentum. For these
reasons, early workers in the quantum theory of angular momentum devoted much effort
to computing Clebsch-Gordan coefficients for ever larger tensor product spaces, thereby
expanding the domain of easily treatable problems. The generality of the Clebsch-Gordan
coefficients contributes to an overall-more-convenient procedure for calculating matrix el-
ements.
In the second step (W2), we compute the reduced matrix element. For this step,
and this step alone, it is necessary to calculate a matrix element “in the raw,” i.e., by
computing an inner product in some basis. Typically, we integrate over three dimensional
space in a position representation. This integration is identical to an elementary matrix
element computation, and thus W2 is just as difficult. As in step W1, it initially appears
as though the Wigner-Eckart theorem has made things no better than the elementary
approach. Again, this step becomes advantageous only when we consider calculating
more than one matrix element. In the eyes of W2, all nontrivial matrix elements are
created equal. Hence, we initially choose whichever looks easiest to compute. This will
generally be a matrix element whose magnetic quantum numbers—m, q, m0 —are zero or
one. Thanks to this choice, W2 provides a gain in tractability: for the one matrix element
we must compute naı̈vely, we can compute the easiest one. That being said, it is possible
that our choice of matrix element will be trivial, i.e. equal to zero. In this case, we choose
again, until we compute a nontrivial matrix element. A nontrivial matrix element en-

21
ables us to divide by the relevant Clebsch-Gordan coefficient, yielding the reduced matrix
element. This division is simply a statement of the Wigner-Eckart theorem with terms
rearranged: the reduced matrix element equals a nonzero matrix element divided by a
Clebsch-Gordan coefficient.
Compared to W1, W2 produces a less general result. The reduced matrix element
depends on the particulars of the system at hand. Nevertheless, for a given operator, the
reduced matrix element possesses a generality unlike any particular matrix element: all
matrix elements with the same parameters j, j 0 , α, and α0 are proportional to the same
reduced matrix element. Here, we again witness how modularization can lead to increases
in generality.
The final step, W3, contains the “turning point” that motivates the entire enterprise.
Knowing both the reduced matrix element and the relevant Clebsch-Gordan coefficients,
we can compute all matrix elements—for fixed parameters j, j 0 , α, and α0 —by taking
products of these. This re-expression via products is the surface level content of the
Wigner-Eckart theorem. W3 shows how the theorem transforms a class of integrals into
a class of two-term products. This provides a drastic gain in tractability, provided we
have the Clebsch-Gordan coefficients readily available.
Before continuing, I elaborate on the methodological advantage of tractability. By
tractability, I mean the ease of problem-solving. I do not intend a dichotomy between
problems that are feasible and those that are infeasible (intractable). Rather, I gesture
at a spectrum of manageability, where more tractable problems are easier to handle than
less tractable problems. The sense of “ease” of problem-solving varies from context to
context. However, in particular cases, it can be made precise. In the case of the Wigner-
Eckart theorem, we can measure tractability by the number of integrals used to compute
a set of matrix elements. Ceteris paribus, more tractable approaches require fewer in-
tegrals. Tractability thereby clarifies the sense in which the Wigner-Eckart theorem is
convenient.
To illustrate these methodological advantages in action, I consider electronic tran-
sitions from an atomic d-orbital to a p-orbital. For simplicity, we focus on radiation
caused by the electric dipole operator P, neglecting higher-order perturbations. P is a
rank one irreducible tensor operator with components Pq1 , where q ∈ {−1, 0, 1}. Ne-
glecting electron spin for simplicity, an atomic state in a d-orbital has j = 2, and
a p-orbital state has j 0 = 1. Before applying any selection rules, this scenario has
0

0 + 1)0 × 1(2k
(2j + 1) × (2j + 1) = (5)0 × (3) × (3) = 45 different matrix elements
α , 1, m Pq α, 2, m for fixed α and α . Applying the selection rule for the magnetic
quantum number, we reduce this to (2k + 1) × (2j 0 + 1) = (3) × (3) = 9 matrix elements.3
Thus, using the elementary approach, we must compute nine integrals. If instead we use
the Wigner-Eckart theorem, we will compute one integral, provided we make a fortuitous
choice in step W2. Using this matrix element and its corresponding Clebsch-Gordan
coefficient, we then calculate the reduced matrix element hα0 , 1k P 1 kα, 2i. Multiplying
this by the eight appropriate Clebsch-Gordan coefficients, we compute the remaining
eight matrix elements (W3). Thus, it is clear that repeated application of W3 is more
tractable than repeated application of the elementary approach, provided we have the
Clebsch-Gordan coefficients.
This caveat brings us to the importance of modularization and generality for conve-
nience. Given enough matrix element problems, it is clearly worthwhile to compute once
3
Since j 0 = 1 < j = 2, we must use j 0 to apply the selection rule. Using j fails to exclude matrix
elements with |m0 | > 1. But these necessarily vanish because m0 ∈ {−1, 0, 1}.

22
and for all the Clebsch-Gordan coefficients. This is true even if we disregard the utility
of these coefficients for angular momentum problems more generally. The generality of
these coefficients makes them worth having around. Furthermore, the modularization
implicit in the Wigner-Eckart theorem makes it necessary to calculate these coefficients
only once. A similar moral applies to W2. The generality of the reduced matrix element
makes it necessary to compute only one per fixed parameters j, j 0 , α, and α0 and operator
T k . Thus, modularizing the original problem into two more general sub-problems (W1
and W2) and one sub-problem of increased tractability (W3) makes the Wigner-Eckart
theorem convenient. Modularization, generality, and tractability work together to pro-
vide a more convenient solution procedure.

2.3 The Conceptual Significance of the Wigner-Eckart


Theorem
Having shown how the Wigner-Eckart theorem is convenient, I return to the problem
posed in Section 2.1: what makes the Wigner-Eckart theorem more than convenient? In
other words, what makes the theorem conceptually significant? To answer this question,
I first present a standard interpretation of the Wigner-Eckart theorem in Section 2.3.1,
which I endorse. On the standard interpretation, the Wigner-Eckart theorem is con-
ceptually significant because it provides a separation of degrees of freedom. Then, in
Section 2.3.2, I show how this interpretation aligns with Butterfield’s remarks on in-
formation in classical mechanics. By separating degrees of freedom, the Wigner-Eckart
theorem provides a restraint on solutions to the matrix element problem, and this re-
straint is a type of information. I go on to generalize “information” to a more inclusive
notion of functional role, i.e. a problem-solving function that an expressive means per-
forms. In terms of functional roles, I reinterpret and extend the standard interpretation of
the Wigner-Eckart theorem. I show that two of the methodological advantages discussed
above—modularization and generality—are novel functional roles provided by Clebsch-
Gordan coefficients and reduced matrix elements. These functional roles ultimately stem
from the restraint that the Wigner-Eckart theorem places on matrix elements, which has
special significance as an epistemic gain.
Thus far, I have elided a distinction between two questions that now becomes rele-
vant. The first question is simpler to ask: “why is the Wigner-Eckart theorem concep-
tually significant?” Considered by itself, this question has an easy answer: the theorem
is conceptually significant because it provides a solution to the matrix element problem.
But by the same token, the elementary approach is also conceptually significant. Indeed,
all accounts of conceptual significance should at least agree on this. Thus, as I have used
it, the first question is really shorthand for a second question: “why is the difference
in expressive means between the Wigner-Eckart theorem and the elementary approach
conceptually significant?” It is really this second question that we are after. However, for
convenience, I continue to speak in terms of the first question.

2.3.1 A standard interpretation of the Wigner Eckart theorem


Physicists throughout the literature agree on a standard interpretation of the Wigner-
Eckart theorem (Rose 1957, 85; Shore and Menzel 1968, 288; Biedenharn and Louck

23
1981a, 97). This standard interpretation is straightforward, relying only on the outward
form of the theorem. It states that the Wigner-Eckart theorem separates a matrix ele-
ment into a product of two factors: the “geometry” and the “physics.”
The Clebsch-Gordan coefficients represent the “geometry.” They capture the depen-
dence of matrix elements on the magnetic quantum numbers m, q, and m0 . These quantum
numbers completely specify the rotational degrees of freedom. Since arbitrary rotations
in space do not affect the processes within an isolated system, these degrees of freedom
are interpreted as geometrical in nature, as opposed to physical.
A reduced matrix element, on the other hand, represents the “physics.” For a given
system, the reduced matrix element captures the dependence of a matrix element on all
other parameters. These parameters depend on the particulars of the physical system
at hand. They include the tensor operator T k , the total angular momentum quantum
numbers j, j 0 , and the collective parameters α, α0 . Hence, the reduced matrix element is
interpreted as containing the “physics” of the system of interest. This physics remains
the same under arbitrary rotations in space (changes in geometrical orientation), but
varies between different systems.
This distinction between geometry and physics is tenable within the nonrelativistic
domain of the standard interpretation. Nonrelativistic quantum mechanics models space
as Euclidean, so space forms a fixed background geometry. Of course, this aspect of the
standard interpretation should not be pushed too far. Due to general relativity, a univer-
sal distinction between geometry and physics is impossible. Nevertheless, this evocative
appeal to geometry and physics underscores why the Wigner-Eckart theorem is concep-
tually significant. The theorem separates degrees of freedom that prima facie appear
coupled.

2.3.2 Information, restraints, and functional roles


The standard interpretation of the Wigner-Eckart theorem aligns well with Butterfield’s
account of information in classical mechanics. According to Butterfield (2004a, b), the La-
grangian, Hamiltonian, Hamilton-Jacobi schemes in classical mechanics provide different
information about the structure of a problem’s solution. This difference in information
partially accounts for why these different formulations of classical mechanics are con-
ceptually significant, as opposed to trivial notational variants. Often, this information
arises as a consequence of symmetry. As the chemical physicist Griffith remarks, “In a
fairly literal sense, the irreducible tensor method is able to extract and manipulate at will
all the information in any problem which is implicit in the symmetry of that problem”
(1962, 3). Although “information” is vague, it is easy to make particular instances of
information concrete. For instance, the Wigner-Eckart theorem provides the following
information: a matrix element of an irreducible tensor operator can be factorized into
two components, only one of which depends on the magnetic quantum numbers.
The Wigner-Eckart theorem provides a particular kind of information that I refer to
as a restraint: it restrains the form of a solution to the matrix element problem. In
general, a restraint refines, controls, or confines the form of a solution or solution space.
Any of these verbs—refine, control, or confine—communicate this sense of information
about the form of a solution. I choose “restrain” because it has a convenient noun form,
namely “restraint.” In the case at hand, the Wigner-Eckart theorem provides the follow-
ing restraint: a matrix element of an irreducible tensor operator in a standard angular

24
momentum basis must decompose into a product of a reduced matrix element and a
Clebsch-Gordan coefficient. This restraint provides detailed information about the form
of a solution to the matrix element problem. Although not all kinds of information are
restraints, any adequate account of information should accommodate restraints.
I now generalize information and restraints to a more inclusive notion: functional role.
Although this generalization may initially appear idle, it clarifies much of the preceding
discussion. In particular, functional roles connect the standard interpretation of the
Wigner-Eckart theorem with Section 2.2’s account of methodological advantages. This
notion also extends my account of the Wigner-Eckart theorem to solve the general puzzle
of Section 2.1.
Since functionalism is a well-established family of positions in philosophy of mind, I
begin with a few disclaimers. First, I focus on the functional roles of expressive means
rather than of mental states. I adopt this restriction to avoid additional philosophical
complications that arise for agents and minds. Although we may later extend a theory of
expressive means to the mental states that they support, this is not my present task. Sec-
ond, I do not adopt the causal language used to discuss functionalism in the philosophy
of mind (e.g. Witmer 2003, 198; Moore 2011, 511). I avoid causal language because it
is ill-suited for analyzing how expressive means function in a problem-solving procedure.
For instance, it would be strange to say that the Wigner-Eckart theorem causes matrix
elements to factorize into a Clebsch-Gordan coefficient times a reduced matrix element.
One function of the Wigner-Eckart theorem is to express this fact, but the theorem is
not a physical process. Therefore, it does not function causally in the world.
As I use the term, a functional role is a particular function performed by an expres-
sive means. Two important functions include providing information about a solution and
expressing a restraint on the form of a solution. For example, one functional role of the
Wigner-Eckart theorem is to provide the restraint on matrix elements described above.
Other important functional roles include functions that are not aptly characterized as
restraints or as information. This includes two of the methodological advantages dis-
cussed in Section 2.2: modularization and generality. In this case, modularization is a
functional role supported by Clebsch-Gordan coefficients and reduced matrix elements:
one function of these expressive means is to modularize the matrix element problem into
the sub-problems W1 through W3. Likewise, the Clebsch-Gordan coefficients and re-
duced matrix elements support the functional role of generality, present in steps W1 and
W2. The Clebsch-Gordan coefficients perform this function by carrying the dependence
of matrix elements on magnetic quantum numbers, thereby providing the generality dis-
cussed in Section 2.2.2. A similar remark applies to the generality of the reduced matrix
element, which functions as the bearer of dependence on the parameters α and α0 and
the irreducible tensor operator T k .
Whereas modularization and generality are functional roles, the methodological ad-
vantage of tractability is not a functional role. Simply put, tractability is not a function.
Instead, tractability is a property of a solution procedure that depends on a compari-
son class of alternative solution procedures. For instance, we saw in Section 2.2.2, that
the Wigner-Eckart theorem is more tractable relative to the elementary approach. In
contrast, we can speak of the functional roles of the Wigner-Eckart theorem without ref-
erence to the elementary approach. Unlike tractability, functional roles do not depend
on a comparison class.
Using these functional roles, I now provide a more detailed account of why the Wigner-
Eckart theorem is conceptually significant. Recall—as stated at the start of Section 2.3—

25
that this is shorthand for the following question: why is the difference in expressive means
between the elementary approach and the Wigner-Eckart theorem conceptually signifi-
cant? Naturally, answering this question requires a comparison of the two approaches.
Relative to the elementary approach, the Wigner-Eckart theorem provides three novel
functional roles: (1) a restraint—here, a separation of degrees of freedom; (2) modular-
ization of the matrix element problem into sub-problems; (3) generality of the factors
that comprise a matrix element. In contrast, the elementary approach has no expressive
means that perform these functional roles. First, it is unable to express this separation of
degrees of freedom. Consequently, it is unable to modularize the matrix element problem
into sub-problems that focus separately on the magnetic quantum numbers and the pa-
rameters α, α0 , and T k . Furthermore, by failing to separate these degrees of freedom, the
elementary approach lacks expressive means with the generality of either Clebsch-Gordan
coefficients or reduced matrix elements. In the next section, I generalize this answer to
provide a criterion for when a difference in expressive means is conceptually significant.
According to my account, differences in tractability do not explain why a difference in
expressive means is conceptually significant. As defined in Section 2.2, tractability is fun-
damentally a matter of convenience: both approaches solve the matrix element problem,
but the Wigner-Eckart theorem is more convenient. This difference in convenience arises
from the functional roles that the Wigner-Eckart theorem provides, but it is not itself
a functional role. Furthermore, although there are strong reasons to prefer using the
Wigner-Eckart theorem, it is ultimately idiosyncratic which problem-solving approach
any particular individual prefers. We can imagine a physicist for whom the idea of con-
sulting a table of Clebsch-Gordan coefficients is anathema. Or, perhaps one is working
with relatively simple integrals, or has a fast-enough computer, such that bothering with
Clebsch-Gordan coefficients or reduced matrix elements does not seem worth the effort.
For these reasons, the conceptual significance of the Wigner-Eckart theorem should not
be attributed to its practical convenience. Changes in expressive means do not have to
be convenient in order to be conceptually significant.
Before generalizing these remarks in the next section, there remains one feature of
the Wigner-Eckart theorem that I should emphasize. Implicit in the above discussion
is the central role played by the restraint that the theorem places on matrix elements.
This restraint, i.e. the separation of degrees of freedom, undergirds the functional roles
of the Clebsch-Gordan coefficients and the reduced matrix elements. It is in virtue of
this separation of degrees of freedom that the theorem modularizes the matrix element
problem. Likewise, the Clebsch-Gordan coefficients and reduced matrix elements acquire
their generality in virtue of this separation of degrees of freedom. This restraint is special
because it serves an epistemic function: it tells us that the reduced matrix element is in-
dependent of the magnetic quantum numbers. In the next section, I argue that epistemic
functions like this play a special role in characterizing differences in expressive means.

2.4 A Spectrum of Expressive Means


Having explained why the Wigner-Eckart theorem is conceptually significant, I propose
a tentative generalization that applies to other differences in expressive means. That is:
I propose a principled distinction between trivial and nontrivial notational variants.
A natural generalization is to require that nontrivially different expressive means pro-
vide different restraints. I call this position “(Restraint);” it is the stronger of the two

26
positions that I consider. A weaker generalization is to require that nontrivial differ-
ences in expressive means provide different functional roles. Since restraints are a kind
of functional role, all cases that satisfy the stronger position also satisfy the weaker po-
sition, but not vice versa. I call this weaker position “(Function).” Below, I elaborate
these positions. Using a number of examples, I argue that we should adopt (Function)
as a necessary and sufficient condition for a conceptually significant difference in expres-
sive means. Finally, I show how (Function) organizes theoretically equivalent expressive
means into a spectrum of conceptual significance.

2.4.1 A proposal: (Restraint)


In Section 2.3, I argued that the Wigner-Eckart theorem is conceptually significant be-
cause it provides a novel restraint on solutions to the matrix element problem. This
restraint is novel in the sense that the elementary approach provides no functionally
equivalent restraint. Moreover, recall that our problem is really to describe the concep-
tual significance of the difference in expressive means between the Wigner-Eckart theorem
and the elementary approach. Generalizing this example leads to the following criterion
for conceptual significance:

(Restraint): A difference between two theoretically equivalent expressive means


is conceptually significant if and only if the two expressive means provide dif-
ferent restraints.

This generalization is supported by a number of examples. It captures why theoreti-


cally equivalent formulations of classical mechanics are conceptually significant. As But-
terfield has shown (2004a, 2004b, 2006), Lagrangian, Hamiltonian, and Hamilton-Jacobi
mechanics all provide different restraints on solutions to many of the same problems in
classical mechanics. I would argue that the same holds true of various alternative formu-
lations of quantum mechanics (e.g. as laid out by Styer et al. 2002), although there is
no space here to develop this point.
(Restraint) also matches our intuitions about a number of non-examples of conceptu-
ally significant expressive differences. For instance, the Einstein summation convention
does not provide any novel restraints relative to the Σ-convention for summation. Like-
wise, Dirac’s bra-ket notation does not provide novel restraints relative to other nota-
tions for expressing inner products and matrix elements. Hence, according to (Restraint),
these changes in expressive means are convenient, but they are not conceptually signifi-
cant. Further non-examples come from many technological advances, which can be highly
analogous to alternative expressive means. Consider the same program implemented by
two computers, one of which is faster. Using the faster computer is convenient, but it does
not provide new restraints on solutions to a problem. Thus, according to (Restraint),
the difference between the two computers is not conceptually significant. Similarly, it
is convenient to use an integrated mass spectrometer and gas chromatograph, but the
integrated machine provides no new restraints relative to using the machines separately.
Both provide the same restraints on the identity of chemical compounds. Although these
kinds of technological changes are extremely important in practice, they are conceptually
insignificant: they do not change how we solve problems of interest.
In summary, (Restraint) provides two categories for theoretically equivalent expres-
sive means: those with a difference in restraints vs. those without such a difference.

27
The former are conceptually significant differences, while the latter are at best merely
convenient. However, I claim that between these two extremes, (Restraint) lumps too
much together with merely convenient expressive differences. As I demonstrate in the
next section, alternative expressive means often support new functional roles, although
they do not provide new restraints. These examples seem different in kind than the non-
examples just considered.

2.4.2 My proposal: (Function)


To illustrate this variety of alternative expressive means—namely, expressive differences
in functional roles but not in restraints—I consider various notations for Clebsch-Gordan
coefficients. By 1981, there were at least 22 different published notations in the literature
(Biedenharn and Louck 1981a, 150-151). Many of these notational changes are indeed
trivial: for instance, there is no conceptual difference between using hjkmq| j 0 m0 i to
denote a Clebsch-Gordan coefficient rather than (jkmq|jkj 0 m0 ), which is Condon and
Shortley’s notation (1935, 75). However, there are important differences between an
0
inner product notation like hjkmq| j 0 m0 i and a coefficient notation like Ajkjmqm0 , used by
Tinkham (1964, 121). These two kinds of notation both support the same restraints
on Clebsch-Gordan coefficients. However, the hjkmq| j 0 m0 i-notation makes available the
0
functional roles of inner products, while the Ajkj mqm0 -notation does not. Since this latter
notation does not express Clebsch-Gordan coefficients in terms of bras and kets, inferences
based on inner products are not available. In contrast, using the hjkmq| j 0 m0 i-notation,
we can express the relationship between the uncoupled basis kets |jkmqi and the coupled
basis kets |j 0 m0 i in terms of a resolution of the identity:
j k
X X
0 0
|j m i = |jkmqi hjkmq| j 0 m0 i (2.2)
m=−j q=−k

Using coefficient notation, we express this relationship as follows:


j k
X X 0
0
|j m i =0
|jkmqi Ajkj
mqm0 (2.3)
m=−j q=−k

Both equations express the same restraint on Clebsch-Gordan coefficients: they are
the matrix elements of a unitary transformation connecting the uncoupled basis to the
coupled basis. However, only the inner product notation allows us to express this restraint
using a resolution of the identity:
j k
X X
I= |jkmqi hjkmq| (2.4)
m=−j q=−k

The inner product notation supports the functional role of a resolution of the identity,
while the coefficient notation does not.
A similar moral applies to a third notation for expressing Clebsch-Gordan coefficients:
the Wigner 3-j symbol. The 3-j symbol takes advantage of matrix notation to treat the
angular momenta indices j, k, j 0 and m, q, m0 on equal footing. This leads to greater
symmetry under permutations of these indices. In terms of a Clebsch-Gordan coefficient,
the 3-j symbol is defined via changes in phase and scale (Brink and Satchler 1968, 136):

28
 
 j1 j2 j3  (−1)j1 −j2 −j3
 = √ hj1 j2 m1 m2 |j3 −m3 i (2.5)
  2j3 + 1
m1 m2 m3

By expressing a Clebsch-Gordan coefficient using a matrix, a 3-j symbol supports a


new functional role: we can permute indices by permuting columns. As defined, a 3-j
symbol remains invariant under cyclic permutation of columns:

     
 j1 j2 j3   j2 j3 j1   j3 j1 j2 
 = =  (2.6)
     
m1 m2 m3 m2 m3 m1 m3 m1 m2

Under anti-cyclic permutation, a 3-j symbols changes phase if the sum of the j indices is
not even (Brink and Satchler 1968, 137):
     
 j2 j1 j3   j1 j3 j2   j3 j2 j1 
 = =  (2.7)
     
m2 m1 m3 m1 m3 m2 m3 m2 m1
 
 j1 j2 j3 
= (−1)j1 +j2 +j3 



m1 m2 m3

0
Compared with the hjkmq| j 0 m0 i and Ajkj mqm0 notations, the 3-j symbol does not ex-
press any new restraints on Clebsch-Gordan coefficients. Any relation that the 3-j symbol
expresses, the hjkmq| j 0 m0 i-notation can express as well by picking up a change in phase
and scale. However, as shown above, the 3-j symbol does support novel functional roles
relative to the hjkmq| j 0 m0 i-notation. Examples like this occupy a middle ground be-
tween merely convenient notational variants and conceptually significant differences in
restraints. Similar remarks apply to graphical notations for Clebsch-Gordan coefficients
(Brink and Satchler 1968, 112ff.)
Generalizing, this example suggests the following criterion for conceptual significance:

(Function): A difference between two theoretically equivalent expressive means


is conceptually significant if and only if the two expressive means perform dif-
ferent functional roles.

Notice first that any expressive difference that satisfies (Restraint) necessarily satisfies
(Function) because providing a restraint is a functional role. Likewise, any expressive
difference that fails to satisfy (Function) necessarily fails to satisfy (Restraint) as well: if
different expressive means do not perform different functional roles, then they certainly
do not provide different restraints. Cases that fail to satisfy (Restraint) thus break up
into two classes, those that satisfy (Function) and those that don’t. The former class
includes the above differences between notations for Clebsch-Gordan coefficients. As I

29
now argue, the latter class includes the earlier examples of conceptually trivial notational
variants. Not only do the Einstein summation convention and Dirac bra-ket notation fail
to support novel restraints relative to other conventions, but also they fail to provide any
new functional roles. Similarly, neither a faster computer nor an integrated spectrometer
provide novel functional roles relative to a slower computer implementing the same pro-
gram or separated spectrometers. Thus, (Function) preserves the spectrum established
by (Restraint) while distinguishing among expressive differences that provide the same
restraints. Some of these expressive differences exhibit different functional roles, a fact
that (Restraint) neglects.
For these reasons, I propose adopting (Function) as a necessary and sufficient condi-
tion for a conceptually significant difference in expressive means. Appropriately, (Func-
tion) is compatible with a spectrum of conceptual significance. Expressive differences
that provide different restraints are more conceptually significant than those that just
provide different functional roles. For instance, adopting the Wigner-Eckart theorem to
compute matrix elements is more conceptually significant than using Wigner 3-j notation
to compute Clebsch-Gordan coefficients.
Furthermore, (Function) leaves room for further practical distinctions between con-
ceptually insignificant differences in expressive means. As discussed above, many con-
ceptually insignificant changes—such as an integrated spectrometer—are nevertheless
practically important. This practical dimension leads to another grade of distinction:
namely, between practically trivial and practically nontrivial differences in expressive
means. A trivial symbolic substitution exemplifies a practically trivial difference, such as
substituting the symbol “y” for “x.” Differences in expressive means that are practically
nontrivial but conceptually insignificant are what we might call merely convenient.
Collectively, these three attributes—convenience, functional roles, and restraints—
lead to a spectrum of theoretically equivalent expressive means. I summarize this spec-
trum in Table 2.1, which lists examples I have discussed and a few others. For each
Class—0. through 3.—I list a paradigmatic example in bold.

30
Table 2.1: A Spectrum of Theoretically Equivalent Expressive Means

Class Attributes Examples

0.
Same methodological advantages. Symbol substitution.
Practically
and Left vs. right-side operator multiplication.
conceptually Same functional roles. Left vs. right-handed coordinate systems.
trivial

1. Einstein summation convention.


Practically
nontrivial Methodological advantage(s). Dirac bra-ket notation.
but Same functional roles. Faster computer, same program.
conceptually
Integrated vs. separate spectrometers.
trivial
0
hjkmq| j0 m0 i vs. Ajkj
mqm0 notations.

Clebsch-Gordan coefficient
vs.
Wigner 3-j symbol.
2.
Methodological advantages. Group-theoretic vs. algebraic derivations
Practically
of traditional Wigner-Eckart theorem.
and Different functional role(s).
conceptually Modern character notation
nontrivial Same restraints.
vs.
Dirichlet’s original notation
(cf. e.g. Morris and Avigad 2014).

Roman numerals vs. Arabic numerals


(cf. e.g. Colyvan 2012, 133-134).
3. Wigner-Eckart vs. elementary approach.
Practically
Methodological advantages. Various formulations of classical mechanics.
and
conceptually Different restraint(s). Various formulations of quantum mechanics
nontrivial (cf. e.g. Styer et. al. 2002).

To further support my proposed criterion, (Function), many examples in the third


column of Table 2.1 need elaboration. I believe this can be done along similar lines to
my analysis of the Wigner-Eckart theorem. However, there is no space to further develop
these examples here.4
4
For simplicity, I neglect a fifth class consisting of expressive differences that are practically trivial but
conceptually nontrivial. This class would sit between Classes 1 and 2 in Table 2.1. One example may be
Cayley’s theorem—that every group is a subgroup of a permutation group. This theorem is conceptually
significant but not very useful in practice. This fifth class plausibly has few examples: generally, novel
functional roles and restraints provide methodological advantages.

31
Finally, I of course admit that examples in each class lie on smaller spectra of practical
and conceptual significance. For instance, a computer that computes ten times faster is
more convenient than one that only computes five times faster. Similarly, differences in
functional role will seem more or less significant insofar as the functional roles themselves
seem more or less significant. Differences in functional role ultimately lead to different
ways of structuring a problem-solving approach. The spectrum of conceptual significance
parallels how significant these different structurings are, i.e. how severe a change is in-
volved. Some changes lead to a novel epistemic gain, e.g. a new restraint, which in
general can be expected to change a problem-solving procedure more dramatically than
a mere change in functional role.

2.5 Summary
I hope to have presented a convincing case for further philosophical examination of
both the Wigner-Eckart theorem and, more generally, theoretically equivalent expres-
sive means. As shown in Section 1.2.2, the Wigner-Eckart theorem provides a cluster
of applications to atoms, molecules, and nuclei where methodological issues come to the
fore. In all of these domains, the theorem provides important simplifications that aid
problem-solving. My analysis in Section 2.2 has shown how this simplification comes
about. Related symmetry arguments in atomic, nuclear, and molecular spectroscopy
can lead to a fuller account of this simplification process. Developing these examples
would contribute significantly to philosophical accounts of why symmetry arguments are
methodologically important.
As developed in this essay, the philosophical issues at the heart of the Wigner-Eckart
theorem go beyond the topic of symmetries. One of the most puzzling aspects of the
Wigner-Eckart theorem is its lack of practical necessity, despite its supreme usefulness.
Although physicists and chemists apply the theorem because it is convenient, they could
get by with more elementary means. This feature raised some central questions: is there
a conceptually significant difference between the elementary approach and the Wigner-
Eckart theorem? If the latter, what constitutes this difference? To generalize these ques-
tions, I developed—in Section 2.1—Manders’ notion of expressive means, which char-
acterizes the notational, mathematical, and physical resources for treating a problem.
Moreover, I argued that the Wigner-Eckart theorem and the elementary approach are
theoretically equivalent because they are committed to the same physical ontology and
provide identical solutions to the matrix element problem. Generalizing this case study
led to the following question: when is a difference between theoretically equivalent ex-
pressive means conceptually significant rather than merely convenient?
To answer this general question, I first—in Section 2.3—developed an answer in the
specific case of the Wigner-Eckart theorem. Compared to the elementary approach, the
Wigner-Eckart theorem provides a novel restraint on the matrix elements of irreducible
tensor operators. Hence, the difference in expressive means between the elementary
approach and the Wigner-Eckart theorem is not merely a matter of convenience: the
Wigner-Eckart theorem provides additional information about matrix elements. Gener-
alizing this answer, I argued in Section 2.4.1 that a difference in restraints is too strong
a criterion for a conceptually significant difference in expressive means. Instead, in Sec-
tion 2.4.2, I proposed that a difference in expressive means is conceptually significant if

32
and only if there is a concomitant difference in functional roles. Although providing a
restraint is a kind of functional role, there are many other functional roles that are not
restraints, such as modularization and generality.
As a necessary and sufficient condition for conceptual significance, differences in func-
tional roles undergird a spectrum of expressive means, illustrated by Table 2.1. This spec-
trum distinguishes between various kinds of theoretically equivalent expressive means. It
separates merely convenient differences from those that are conceptually significant, while
further distinguishing cases within these classes. This spectrum is ripe for further de-
velopment. By conducting similar analyses, we can classify more examples while further
refining these classes. These studies will clarify the significance of mathematical, nota-
tional, and diagrammatic re-expression in both science and mathematics.

33
References
Armstrong, D. M. (1983). What is a law of nature? Cambridge: University Press
Avigad, J. (Forthcoming). Mathematics and language. In E. Davis & P. Davis (Eds.),
Mathematics, substance and surmise: views on the meaning and ontology of
mathematics: Springer.
Avigad, J., & Morris, R. (2014). The concept of "character" in Dirichlet's theorem on primes in
an arithmetic progression. Archive for History of Exact Sciences, 68, 265-326.
Bangu, S. (2013). Symmetry. In R. Batterman (Ed.), The Oxford Handbook of Philosophy of
Physics (pp. 287-317). Oxford: Oxford University Press.
Batterman, R. W. (2002). The devil in the details: asymptotic reasoning in
explanation, reduction, and emergence. Oxford: Oxford University Press.
Batterman, R. W. (2010). On the explanatory role of mathematics in empirical science.
British Journal for the Philosophy of Science, 61, 1-25.
Batterman, R. W., & Rice, C. C. (2014). Minimal model explanations. Philosophy of Science,
81(3), 349-376.
Bell, J. S. (1987). Six possible worlds of quantum mechanics. Speakable and unspeakable in
quantum mechanics (pp. 181-195). Cambridge: Cambridge University Press.
Biedenharn, L. C., & Louck, J. D. (1981a). Angular momentum in quantum physics: theory
and application. Reading, Mass.: Addison-Wesley
Biedenharn, L. C., & Louck, J. D. (1981b). The Racah-Wigner algebra in quantum
theory. Reading, Mass.: Addison-Wesley.
Biedenharn, L. C., & van Dam, H. (1965). Quantum theory of angular momentum: a collection
of reprints and original papers. New York: Academic Press.
Brink, D. M., & Satchler, G. R. (1968). Angular momentum (2nd ed.). Oxford: Clarendon Press.
Brussaard, P. J., & Glaudemans, P. W. M. (1977). Shell-model applications in nuclear
spectroscopy. Amsterdam: North-Holland.
Butterfield, J. N. (2004a). Between laws and models: some philosophical morals of Lagrangian
mechanics. arXiv.org.
Butterfield, J. N. (2004b). On Hamilton-Jacobi theory as a classical root of quantum theory. In
A. C. Elitzur, S. Dolev & N. Kolenda (Eds.), Quo Vadis Quantum Mechanics?
Possible Developments in Quantum Theory in the 21st Century. New York: Springer.
Butterfield, J. N. (2006). On symmetry and conserved quantities in classical mechanics. In W.
Demopoulos & I. Pitowsky (Eds.), Physical theory and its interpretation: essays in honor
of Jeffrey Bub (pp. 43-99). Dordrecht: Springer.
Chang, H. (2009). Ontological principles and the intelligibility of epistemic activities. Scientific
understanding: philosophical perspectives (pp. 64-82). Pittsburgh: University of
Pittsburgh Press.
Colyvan, M. (2012). An introduction to the philosophy of mathematics. Cambridge: Cambridge
University Press.
Condon, E. U., & Shortley, G. (1935). The theory of atomic spectra. Cambridge: Cambridge
University Press.
Cornwell, J. F. (1984). Group theory in physics. London: Academic Press.
de-Shalit, A., & Talmi, I. (1963). Nuclear shell theory. New York: Acad. Press.
de Regt, H. (2009). Understanding and scientific explanation. Scientific understanding:
philosophical perspectives (pp. 21-42). Pittsburgh: University of Pittsburgh Press.
de Regt, H., & Dieks, D. (2005). A contextual approach to scientific understanding. Synthese,
144, 137-170.
Earman, J. (2004). Laws, symmetry, and symmetry breaking: invariance, conservation
principles, and objectivity. Philosophy of Science, 71, 1227-1241.
Eckart, C. (1930). The application of group theory to the quantum dynamics of monatomic
systems. Revs. Mod. Phys., 2, 305-380.
Frank, A., & Van Isacker, P. (1994). Algebraic methods in molecular and nuclear structure
physics. New York: Wiley.
Griffith, J. S. (1962). The irreducible tensor method for molecular symmetry groups. Englewood
Cliffs, N.J.: Prentice-Hall.
Hamermesh, M. (1962). Group theory and its application to physical problems. Reading, Mass.:
Addison-Wesley.
Hempel, C. G., & Oppenheim, P. (1965 [1948]). Studies in the logic of explanation. Aspects of
scientific explanation, and other essays in the philosophy of science (pp. 245-290). New
York: Free Press.
Iachello, F., & Levine, R. D. (1995). Algebraic theory of molecules. New York: Oxford
University Press.
Judd, B. R. (1963). Operator techniques in atomic spectroscopy. New York: McGraw-Hill.
Judd, B. R. (1985). History of complex spectra. Reports on progress in physics, 48, 907-954.
Khalifa, K. (2012). Inaugurating understanding or repackaging explanation? Philosophy of
Science, 79(1), 15-37.
Kibler, M. R. (2011). The impact of Giulio Racah on crystal and ligand-field theories. Paper
presented at the international conference in commemoration of the centenary of the birth
of G. Racah (1909-1965), Zaragoza, Spain. arXiv.org.
Kitcher, P. (1989). Explanatory unification and the causal structure of the world. In P. Kitcher &
W. Salmon (Eds.), Scientific Explanation (Vol. 13, pp. 410-505). Minneapolis: University
of Minnesota.
Lewis, D. K. (1973). Counterfactuals. Cambridge: Harvard University Press.
Lipton, P. (2004). Inference to the best explanation (2nd ed.). London: Routledge.
Mackey, G. W. (1981). Introduction. In L. C. Biedenharn & J. D. Louck (Eds.), The Racah-
Wigner algebra in quantum theory (pp. xxix-lxxxviii). Reading, Mass.: Addison-Wesley.
Manders, K. (2008a). Diagram-based geometric practice. In P. Mancosu (Ed.), The
philosophy of mathematical practice (pp. 65-79). Oxford: Oxford University Press.
Manders, K. (2008b). The Euclidean diagram (1995). In P. Mancosu (Ed.), The philosophy
of mathematical practice (pp. 80-133). Oxford: Oxford University Press.
Manders, K. (unpublished). Expressive Means and Mathematical Understanding.
Moore, D. (2011). Role functionalism and epiphenominalism. Philosophia, 39, 511-525.
Morrison, M. (2000). Unifying scientific theories: physical concepts and mathematical
structures. Cambridge: Cambridge University Press.
Norton, J. D. (2012). Approximation and idealization: why the difference matters. Philosophy of
Science, 79(2), 207-232.
Racah, G. (1942). Theory of Complex Spectra. I. Phys. Rev., 61, 186-197.
Racah, G. (1949). Theory of Complex Spectra. IV. Phys. Rev., 76, 1352-1365.
Rose, M. E. (1957). Elementary theory of angular momentum. New York: Wiley.
Shore, B. W., & Menzel, D. H. (1967). Principles of atomic spectra. New York: Wiley.
Smith, P. R., & Wybourne, B. G. (1967). Selection rules and the decomposition of the Kronecker
square of irreducible representations. Journal of mathematical physics, 8, 2434.
Styer, D. F., Balkin, M. S., Becker, K. M., et. al. (2002). Nine formulations of quantum
mechanics. American Journal of Physics, 70(3), 288-297.
Tinkham, M. (1964). Group theory and quantum mechanics. New York: McGraw-Hill.
van Fraassen, B. C. (1989). Laws and symmetry. Oxford: Oxford University Press.
Wigner, E. (1927). Einige Folgerungen aus der Schrödingerschen Theorie für die
Termstrukturen. Z. Physik 67, 624-652.
Wigner, E. P. (1931). Gruppentheorie und ihre Anwendung auf die Quantenmechanik der
Atomspektren. Braunschweig: F. Vieweg & Sohn.
Wigner, E. P. (1959). Group theory and its application to the quantum mechanics of atomic
spectra (J. J. Griffin, Trans.). New York: Academic Press.
Witmer, D. G. (2003). Functionalism and causal exclusion. Pacific Philosophical Quarterly, 84,
198-214.
Wulfman, C. (2011). Dynamical symmetry. Hackensack, N.J.: World Scientific.
Wybourne, B. G. (1970). Symmetry principles and atomic spectroscopy. New York: Wiley-
Interscience.
Wybourne, B. G. (1972). Ligand field theory and continuous groups. Chemical Physics Letters,
16(2), 365-368.
Wybourne, B. G. (1973a). The "gruppen pest" yesterday, today, and tomorrow. International
Journal of Quantum Chemistry, 7, 35-43.
Wybourne, B. G. (1973b). Lie algebras in quantum chemistry: symmetrized orbitals.
International Journal of Quantum Chemistry, 7, 1117-1137.
Wybourne, B. G. (1974). Classical groups for physicists. New York: Wiley.
Zeldes, N. (2009). Giulio Racah and theoretical physics in Jerusalem. Archive for History of
Exact Sciences, 63, 289-323.

You might also like