Hostname: page-component-6bf8c574d5-rwnhh Total loading time: 0 Render date: 2025-02-23T05:22:54.912Z Has data issue: false hasContentIssue false

Viewing Quantum Charge from the Classical Vantage Point

Published online by Cambridge University Press:  10 June 2022

Marian J. R. Gilton*
Affiliation:
Department of History and Philosophy of Science, University of Pittsburgh, 4200 Fifth Avenue, Pittsburgh, PA 15260, United States
Rights & Permissions [Opens in a new window]

Abstract

This article demonstrates the benefit of studying a classical version of chromodynamics in order to better understand color charge in quantum chromodynamics. Standard presentations of the conservation and confinement of color charge serve to obscure the Lie-algebra-valued character of the conserved Noether charge. This article shows how we can remove these obscuring factors by studying color charge from the vantage point of classical chromodynamics. This key example of color charge illustrates the larger methodological benefit of this classical vantage point: interpreting classical gauge theories helps to delineate the uniquely quantum features of quantum field theories.

Type
Symposia Paper
Copyright
© The Author(s), 2022. Published by Cambridge University Press on behalf of the Philosophy of Science Association

1. Introduction

The quantum field theories of the standard model of particle physics are formulated, with varying degrees of mathematical rigor, by applying quantization procedures to classical versions of these theories. It is the resulting quantum versions that are the successful theories, and it is these theories, therefore, that should be the proper subject of philosophical investigation into what our current best physics has to say about the nature of matter at the smallest levels accessible to empirical study. It may seem, therefore, that the classical versions of these theories are philosophically insignificant, if they are construed as mere formal precursors to the real theory born out of the process of quantization. I argue against this view, defending the methodological value of interpreting classical theories and understanding their quantum counterparts from that classical vantage point.

I demonstrate the value of the classical vantage point by using it to elucidate the Lie-algebra structure of the conserved color charge quantity in quantum chromodynamics (QCD). This $\mathfrak{su}(3)$ character of color charge is obscured in QCD as a result of confinement, but it is made manifest in the unconfined classical version of chromodynamics. The license to interpret quantum charge from the classical vantage point derives from the shared geometric structure (especially as it encodes the relevant gauge symmetry) across the relevant classical and quantum theories.

The notion that our understanding of the world can benefit from scientific theories that are either outright false or else nonfundamental is not new. We can distinguish two categories for how this can be, one for the outright falsehoods and one for the nonfundamental theories. In the first category, many successful model-building strategies make use of a series of false models. Although this view was made famous in the context of the philosophy of biology by Wimsatt (Reference Wimsatt, Nitecki and Hoffman1987), it is no stranger to the philosophy of physics. Li (Reference Li2013), for instance, applies Wimsatt’s account of false models to the development of quantum electrodynamics in the 1930s and 1940s. Furthermore, Cohen (Reference Cohen1983) gives a similar assessment of the value that idealizations (e.g., initially treating planets as point particles and treating ellipses as circles) have in Newton’s Principia. And the more recent literature on idealization (e.g., Batterman Reference Batterman2009) frequently addresses the notion that the strict claims of the idealization employed in a given area of physics are explanatorily ineliminable. The second category concerns the status of nonfundamental theories, which can be viewed as either false or as true depending on one’s stance on reductionism. Thus, one might argue that the special sciences have their own domains of applicability that give them at least an initial claim to be respectable, autonomous sciences separate from fundamental physics (see, e.g., Hendry Reference Hendry, Baird, Robert Scerri and McIntyre2006).

The value of interpreting classical chromodynamics defended here does not fit into either of these categories. This particular classical field theory has no claim to its own domain of applicability, nor does it constitute a specific idealized model within another theory. Rather, classical gauge theories such as classical chromodynamics are best thought of as “classical bones” beneath the “quantum flesh” of current best physics, to borrow Boris Kinber’s famous image for understanding semi-classical mechanics.Footnote 1 By dissecting the “bones” from the “flesh,” we can more easily discern those distinctively quantum features of a given quantum field theory from those features that arise from the shared geometric backbone of both classical and quantum field theory.

2. Color charge in quantum chromodynamics

Charge properties are of philosophical interest for several reasons. Foremost, charge properties are major conceptual linchpins in their respective theories. The electric, color, and weak charges of the standard model (SM) govern the fundamental interaction processes of each theory, categorizing those particles that are eligible for participation in each interaction and bearing a close conceptual connection to the coupling constants that encode the strengths of the interactions for each theory. In addition, as conserved quantities via Noether’s theorem, these charge properties have a distinguished level of physical significance. From the standpoint of metaphysics, these properties are especially relevant for questions concerning the interpretation of fundamental properties in our current best physics of the smallest empirically accessible length scales.

In this section, I focus on two important features of color charge in QCD. First, the conserved Noether current for color charge in QCD is Lie-algebra-valued, taking values in $\mathfrak{su}(3)$ . Second, QCD color charge is confined. I argue that confinement (and related notions) in QCD obscures the Lie-algebra-valued character of color charge. In section 3, I argue that the $\mathfrak{su}(3)$ character of color charge is made manifest in a classical version of chromodynamics.

2.1 Disambiguating confinement from related notions

It is well known that the result of applying Noether’s theorem to gauge symmetries results in a conserved current that takes values in the Lie algebra associated to the symmetry group.Footnote 2 In the case of electrodynamics, this reduces to the usual sense of conserved electric charge given in real numbers. The relevant symmetry group is $U(1)$ , whose Lie algebra is simply $\mathbb{R}$ . Thus, we may not immediately recognize that electric charge is in fact a Lie-algebra-valued quantity. In chromodynamics, the conserved Noether current takes values in the Lie algebra $\mathfrak{su}(3)$ because $\mathfrak{su}(3)$ is the Lie algebra of the symmetry group $SU(3)$ .

For the interpretation of charge in QCD, however, this $\mathfrak{su}(3)$ feature of color charge is obscured by the further recognition that color charge is confined. As is well known, quarks, anti-quarks, and gluons are not observable in isolation. Rather, nature only allows for them to appear in bound states with each other in such a way that the net color charge of the observed state is always neutral. In discussions of the conservation of color charge, most textbook discussions proceed at the level of specific quark and anti-quark color states (r, b, g; $\bar{r}$ , $\bar{b}$ , $\bar{g}$ ). Often in conjunction with Feynman diagrams, physicists speak of the conservation of color in terms of, separately, the conservation of each of these kinds of color at each vertex in a diagram (see figure 1). Because (r, b, g; $\bar{r}$ , $\bar{b}$ , $\bar{g}$ ) are not elements of $\mathfrak{su}(3)$ , this gives a very different way of thinking of the conservation of color charge than the sense of conservation given by Noether’s theorem.Footnote 3

Figure 1. Color states r, b, and g accounted for at each vertex within a Feynman diagram, here given for the process of nucleon scattering via pion exchange. Straight lines with arrows depict quark states, and curly lines with double coloring depict gluon states.

It is valuable to disambiguate several ideas related to confinement. First, there is the as-yet-unproven result of quark/gluon confinement as a consequence of the dynamics of QCD. Confinement is the phenomenon that free quarks and gluons do not exist at low energies.Footnote 4 Second, there is the theoretical stipulation of color neutrality. For example, in the construction of proton states, p, we take the tensor product of the $SU(3)$ carrier spaces for each of the three valance quarks within the proton: $p \in \mathbb{C}^3 \otimes \mathbb{C}^3 \otimes \mathbb{C}^3 \cong \mathbb{C}^{1} \oplus \mathbb{C}^8 \oplus \mathbb{C}^8 \oplus \mathbb{C}^{10}$ . On the right-hand side, we have the four different mathematical possibilities for which representation of $SU(3)$ the proton could be in: $\mathbb{C}^1$ , two copies of $\mathbb{C}^8$ , or $\mathbb{C}^{10}$ . The $\mathbb{C}^1$ option is the only physical possibility, according to the stipulation of color neutrality: all observable states must transform according to the trivial representation (whose carrier space is $\mathbb{C}^1$ ) of the $SU(3)$ color symmetry, that is, be overall color-charge-neutral.

Research into the ultraviolet and infrared behavior of the strong coupling constant $\alpha _s$ in QCD aims to give a dynamical explanation for both quark/gluon confinement and color neutrality. Here, the term constant is something of a misnomer because $\alpha _s$ is a “running” coupling constant whose value depends on the energy scale. There is widespread consensus (although see Seiler [Reference Seiler2003] for a dissenting view) that applying perturbation theory to QCD proves that the theory has ultraviolet asymptotic freedom; that is, the effective QCD coupling goes to zero in the ultraviolet limit as distance goes to zero. QCD appears to have, as demonstrated numerically in lattice QCD, the opposite effect in its infrared behavior: in the infrared regime, the coupling increases as distance increases.Footnote 5 If the coupling in fact increases in this manner, as suggested by lattice QCD, then this would serve to provide a dynamical explanation of quark confinement. The increase of the coupling with increased distance would keep individual quarks and gluons from leaving the confines of their hadronic homes. These numerical results are forcefully suggestive of quark confinement, yet they fall short of an airtight proof. If a proof of this infrared behavior is found, it is expected that it would further show that color neutrality is a dynamical consequence of the theory. Another possibility is that rather than displaying this increasing infrared behavior, the effective coupling “freezes” at a definite value at a given energy and remains frozen for higher energies. This possibility is suggested by recent work within the functional renormalization group framework.Footnote 6 Our primary focus in the next subsection will be the interpretive difficulties posed by the phenomena of quark confinement and color neutrality, setting aside questions of how the ultraviolet and infrared behavior of the strong coupling might explain these phenomena.

2.2 Obscuring ${\mathfrak {su}}(3)$

Because the fundamental entities that are the bearers of color charge are confined within color-neutral hadrons, the metaphysical complexities of how color charge is conserved are far from obvious. Prima facie, conserving a quantity by never allowing it to divert from zero in observable systems is a particularly uninteresting conservation law. Just below the surface, however, there are two ways in which an overall color-neutral state can be found. First, equal amounts of r, b, and g combine to form a “white” or color-neutral state. Second, components of these three colors can be compensated for with equal amounts of their corresponding anti-colors, $\bar{r}$ $\bar{b}$ , and $\bar{g}$ . This suggests (as does figure 1) that what is conserved is red, and separately blue, and separately green. But that is not what Noether’s theorem gives us. By focusing on the (r, b, g; $\bar{r}$ , $\bar{b}$ , $\bar{g}$ ) properties separately, the Feynman-diagram approach to color charge conservation obscures the role of $\mathfrak{su}(3)$ in this conservation law.

I argue in the next section that we can, instead, make sense of the conservation of the $\mathfrak{su}(3)$ Noether charge in QCD by investigating color charge in a classical version of quantum chromodynamics. In that classical theory, the formulation of the Wong force law (the non-Abelian generalization of the Lorentz force law) leads to the ascription of an $\mathfrak{su}(3)$ -valued quantity of charge for matter fields, as well as for the gauge field. This gives us a vantage point from which to see clearly the $\mathfrak{su}(3)$ character of charge for the matter fields themselves and how they therefore contribute to the overall conserved $\mathfrak{su}(3)$ Noether charge. Moreover, by choosing to investigate charge in this classical version of chromodynamics, we have set aside the physically important but conceptually confounding issues of color neutrality, quark/gluon confinement, and ultraviolet and infrared behavior. By looking at this classical version of the theory with unconfined matter fields, we can more clearly see the $\mathfrak{su}(3)$ character of quantum color charge.

3. Color charge in quantum chromodynamics

In this section, I show how the formulation of a classical version of chromodynamics in terms of fiber bundles gives a particular vantage point for understanding the $\mathfrak{su}(3)$ character of color charge in QCD. This classical version of chromodynamics is not the final, correct theory, and it does not enjoy the empirical success that QCD has.Footnote 7 Nevertheless, the geometric “bones” of QCD are more clearly studied in classical chromodynamics than in QCD itself. And these geometric bones give a clear sense in which the color-charged particles have $\mathfrak{su}(3)$ -valued charge. This serves to illustrate the interpretive advantage of the classical vantage point. By isolating and investigating the “classical bones” supporting the “quantum flesh” of our successful field theories, we can separate various conceptual structures and relationships in quantum chromodynamics. In particular, we see here that two key features of color charge picked out in section 2, namely, confinement and the role of $\mathfrak{su}(3)$ , come from two different places in QCD. Confinement must be found in the “quantum flesh,” whereas the Lie-algebra-valued nature of charge is manifest in the “classical bones.”

3.1 The Wong force law

We may formulate the basic features of a classical version of chromodynamics as follows.Footnote 8 We will adopt the abstract index notion developed by Wald (Reference Wald1984) with the further notational conventions of Weatherall (Reference Weatherall2016). Vectors and tensors tangent to M have lowercase Latin indices a, b, c; vectors and tensors tangent to the total space P have lowercase Greek indices; and uppercase Fraktur indices are used for vectors with a Lie-algebra structure. In addition, indices i, j, and so on will be used to label vectors in the carrier space V of a representation of $SU(3)$ used to construct an associated bundle. Using these conventions, fix a relativistic spacetime $(M, g_{ab}$ ) and an $SU(3)$ principal bundle $P \overset{\wp} {\rightarrow}M$ over M. In addition, fix a principal connection $\omega ^{\mathfrak{A}}_{\alpha }$ on P and an inner product $k_{\mathfrak{AB}}$ on the Lie-algebra $\mathfrak{su(3)}$ associated to $SU(3)$ .

The curvature of the principal connection is interpreted as the chromodynamic field strength. This curvature is

(1) $$\Omega _{\;\;\alpha \beta }^{\mathfrak{A}} = {d_\alpha }\omega _{\;\;\beta }^{\mathfrak{A}} + {1 \over 2}[\omega _{\;\;\alpha }^{\mathfrak{A}},\omega _{\;\;\beta }^{\mathfrak{A}}],$$

where $d_\alpha$ is the exterior derivative on P, and the bracket $[\cdot,\cdot ]$ is the Lie bracket on $\mathfrak{su}(3)$ .

The quark and anti-quark matter fields correspond to sections of different associated bundles $P \times _G V \ \overset{\pi \rho} {\rightarrow } M$ , where V is the carrier space for an irreducible representation $\rho$ of $SU(3)$ . Thus, a matter field $\Psi : M \rightarrow P \times _G V$ maps points $x \in M$ to equivalence classes $[p, v^i]$ for some $p \in P$ and $v^i \in V$ . Here $[p, v^i] \cong [p', v^j]$ just in case there exists a $g \in SU(3)$ such that $(pg, \rho (g^{-1})v^i) = (p', v^j)$ . These vectors $v^k$ describe state vectors in the internal charge space for an elementary particle of this kind of matter field (either a quark, anti-quark, or gluon).

The sections $\Psi : M \rightarrow P \times _G V$ are in one-to-one correspondence with V-valued, G-equivariant maps $\psi ^i$ on P. Given a section evaluated at a point x in M, $\Psi (x) = [p, v]$ , we define $\psi ^i : P \rightarrow V$ by $\psi ^i(p) \mapsto \lambda ^{-1}_p(\Psi (\wp (p)))$ , where the map $\lambda _p : V \rightarrow \pi _\rho ^{-1}(\wp (p))$ is defined by $\lambda _p(v) = [p, v]$ . In different contexts, it is sometimes more convenient to use $\psi ^i$ on P rather than $\Psi$ on M. In particular, it is technically simpler to define the action of the covariant derivative induced by the connection $\omega ^{\mathfrak{A}}_{\alpha }$ on $\psi ^i$ than on $\Psi$ . For a field $\psi ^i: P \rightarrow V$ , its covariant derivative is the ordinary exterior derivative $d_\alpha$ on P following the vertical projection. That is, for any vector $\xi ^\alpha$ on P, the covariant derivative induced by the connection is ${\overset {\omega }{D}}{_\alpha }\psi ^i\xi ^\alpha = d_\alpha \psi ^i\xi ^{\alpha |} $ , where $\xi ^{\alpha |} = \omega ^{\mathfrak{A}}\ _{\alpha }\xi ^\alpha$ .

The generalization of the Lorentz force law to non-Abelian gauge theory is known as the Wong force law, first given by Wong (Reference Wong1970). It can be mathematically derived as follows, although the status of this derivation as a physical argument is unclear.Footnote 9 Using the inner product $k_{\mathfrak{AB}}$ , the metric $g_{ab}$ on M, and our principal connection $\omega ^{\mathfrak{A}}_{\alpha }$ , we can construct a metric on the total space known as the bundle metric. First, we take $g_{ab}$ on the base space M, and we pull this back along the projection to get a symmetric rank 2 tensor $(\wp ^*g)_{\alpha \beta }$ on the total space P. That is,

(2) $${({\wp ^*}g)_{\alpha \beta }}{\sigma ^\alpha }{\eta ^\beta } = {g_{ab}}({\wp _*}{\sigma ^\alpha })({\wp _*}{\eta ^\beta })$$

for all vectors $\sigma ^\alpha, \eta ^\alpha$ at a point p in P. There is another symmetric rank 2 tensor on P, denoted $k_{\alpha \beta }$ , defined in terms of the connection $\omega ^{\mathfrak{A}}_{\alpha }$ . It is given by

(3) $$k_{\alpha \beta }\sigma ^\alpha \eta ^\beta = k_{\mathfrak{AB}} \omega ^{\mathfrak{A}}\ _{\alpha }\sigma ^\alpha \omega ^{\mathfrak{B}}\ _{\beta }\eta ^\beta. $$

For any two vectors $\sigma ^\alpha$ and $\eta ^\beta$ at a point p in P, $k_{\alpha \beta }$ gives the inner product of their vertical projections. Adding these two tensors gives us our bundle metric $h_{\alpha \beta } = (\wp ^*g)_{\alpha \beta } + k_{\alpha \beta }$ .

We can use the bundle metric to classify curves through the total space. Let $\gamma : [0, 1] \rightarrow P$ be a geodesic relative to $h_{\alpha \beta }$ with tangent field $\xi ^\alpha (t)$ . It follows that $\omega ^{\mathfrak{A}}\ _{\alpha }\xi ^\alpha (t) = Q^{\mathfrak{A}} \in \mathfrak{su}(3)$ is independent of t (see Bleecker Reference Bleecker2013, theorem 10.1.5). Let $\tilde{\gamma } = \wp \circ \gamma$ be the projection of $\gamma$ down to the base space with tangent field $\xi ^a$ . It follows that, relative to a choice of section $\sigma$ , the acceleration of this curve $\tilde{\gamma }$ on spacetime obeys the Wong force law:

(4) $$\xi ^n\nabla _n\xi ^b = k_{\mathfrak{AB}}g^{cb}Q^{\mathfrak{A}}\Omega ^{\mathfrak{B}} _{\,\,\,\,ac}\xi ^a = Q^{\mathfrak{A}}\Omega _{\mathfrak{A}a}^{\,\,\,\,\,\,b}\xi ^a, $$

where $\Omega ^{\mathfrak{A}}_{\,\,\,ab} = \sigma ^*(\Omega ^{\mathfrak{A}}_{\,\,\,\,\alpha \beta })$ is the field strength (see Bleecker Reference Bleecker2013, theorem 10.1.6). We interpret $\tilde{\gamma }$ as the world line for a particle of mass m and charge $q^{\mathfrak{A}} = Q^{\mathfrak{A}}/m$ . If $G = U(1)$ , this reduces to the familiar Lorentz force law for electromagnetism. In that case, we interpret $q^{\mathfrak{A}} \in \mathbb{R}$ as the amount of electric charge carried by the particle whose world line is $\tilde{\gamma }$ . Thus, in classical chromodynamics, $q^{\mathfrak{A}}$ is the non-Abelian charge property analog of electric charge.

3.2 Interpretation

How may we interpret $q^{\mathfrak{A}}$ in this formulation of classical chromodynamics? We usually predicate the color charge of a particle in the sense of a basis vector in a representation of $SU(3)$ , that is, a specific color state (r, b, g; $\bar{r}$ , $\bar{b}$ , $\bar{g}$ ). Using the electromagnetic interpretation as a guide, one might then expect that $q^{\mathfrak{A}}$ gives the color state of the particle under the influence of the Wong force. And because gluons are the only type of particle with Lie-algebra-valued color states, one might conclude that $\tilde{\gamma }$ is the world line of a gluon.

However, such reasoning is mistaken. The color-charged matter fields subject to this Wong force law couple to the gauge field via the charge-current density. This charge-current density is itself a Lie-algebra-valued 1-form $J^{\mathfrak{A}}_{\alpha }$ on the total space P. The charge $q^{\mathfrak{A}}$ is Lie-algebra-valued because the charge current density is Lie-algebra-valued.

The definition of $J^{\mathfrak{A}}_{a}$ relies on the inner product $k_{\mathfrak{A}\mathfrak{B}}$ on $\mathfrak{su}(3)$ , as well as an inner product $h_{ij}$ on the carrier space V. Fix a basis $\{e^{\mathfrak{A}}\}$ of the Lie algebra $\mathfrak{su}(3)$ . Then, following Bleecker (Reference Bleecker2013, theorem 5.1.2), the current $J^{\mathfrak{A}}_{a}$ is given by

(5) $$J^{\mathfrak{A}}_{\alpha } = k^{\mathfrak{AB}}e_{\mathfrak{B}}h_{ij}\tilde{\psi }^j\overset{\omega }{D}_\alpha \psi ^i, $$

where $\tilde{\psi }^j = \rho _*(e_{\mathfrak{A}})\vartriangleright \psi ^j$ . That is, $\tilde{\psi }^j$ is the result of transforming $\psi ^j$ under the representation $\rho _*$ of $\mathfrak{su}(3)$ on V induced by the representation $\rho$ of G on V.Footnote 10 The definition of the current in equation 5 gives us mathematical reason to acknowledge that the charge values $q^{\mathfrak{A}}$ for charged matter fields that contribute to $J^{\mathfrak{A}}_{\alpha }$ are Lie-algebra-valued.

This general expression for the current associated with a charged matter field $\psi ^i$ in a non-Abelian gauge theory reduces to the more familiar current of electromagnetism as follows. Suppose now that $G = U(1)$ . Further, choose $i = \sqrt{-1}$ as a basis for $U(1)$ ’s associated Lie algebra $\mathfrak{g}= \mathbb{R}$ , and choose $k_{\mathfrak{A}\mathfrak{B}}$ such that $k_{\mathfrak{A}\mathfrak{B}}e^{\mathfrak{A}}e^{\mathfrak{B}} = 1$ . Then the current $J^{\mathfrak{A}}_{\alpha }$ becomes $J^{\mathfrak{A}}_{\alpha } = ih_{ij}(i\psi )^j\overset{\omega }{D}_\alpha \psi ^i.$ Now fix a choice of local section $\sigma : U \rightarrow P$ . The vector potential is $A_a = \sigma ^*(\omega ^{\mathfrak{A}}_{\alpha })$ . Similarly, the complex scalar field on M is $\Psi = \sigma ^*(\psi ^i)$ . With this notation, a local representation of the current on M is

(6) $$ J_a = i((i\Psi )^*\nabla _a\Psi + i\Psi (\nabla _a\Psi )^*) $$
(7) $$ = \Psi ^*\nabla _a\Psi -\Psi (\nabla _a\Psi )^*, $$

where $\nabla _a = \partial _a -iA_a$ is the covariant derivative on M, and the star ${}^*$ indicates complex conjugation.

The foregoing discussion shows how color charge has a distinctly Lie-algebra-valued character as a result of the structure of the field theory that is captured in the geometry and not from the quantum character of QCD. Electric charge in QED is also Lie-algebra-valued, but because the Lie algebra of $U(1)$ is $\mathbb{R}$ , we do not immediately recognize the role of the Lie algebra for electric charge. The field strength (equation 1) is Lie-algebra-valued from the start. This fits with what we expect of the gauge field for chromodynamics in relation to gluons, which transform according to the adjoint representation. Moreover, the charge $q^{\mathfrak{A}} = Q^{\mathfrak{A}}/m$ in equation 4 is Lie-algebra-valued, and it gives the charge of the particle whose trajectory obeys the Wong force law. Thus, charge for quarks and anti-quarks is not only a matter of states within the fundamental representations of $SU(3)$ (wherein we find (r, b, g; $\bar{r}$ , $\bar{b}$ , $\bar{g}$ )) but also of Lie-algebra-valued contributions to the charge current density $J^{\mathfrak{A}}_{a}$ .

Because there is this sense in which charged particles have $\mathfrak{su}(3)$ charge, this gives us another vantage point from which to consider the conversation of the current $J^{\mathfrak{A}}_{\alpha }$ (see Bleecker Reference Bleecker2013, theorem 5.1.5). Rather than thinking of conservation of color charge in terms of (r, b, g; $\bar{r}$ , $\bar{b}$ , $\bar{g}$ ) values in Feynman diagrams, we can think of it in terms of $\mathfrak{su}(3)$ contributions to $J^{\mathfrak{A}}_{\alpha }$ . Although there is merit to the Feynman-diagram viewpoint of charge conservation, the benefit of this classical vantage point is that the Lie-algebra-valued character of $J^{\mathfrak{A}}_{\alpha }$ is manifest.

The point is not that we can only see these features of charge from the classical vantage point or that they are not available from the quantum-field-theory vantage point. Indeed, these features are present in the quantum version of the theory precisely because they are, first of all, present with the quantum theory’s “classical bones.” And by dissecting the “bones” apart from the quantum “flesh,” we are better poised to see the role the bones play in the final, finished quantum body of knowledge.

4. Conclusion

I have argued for the benefit to be gained from viewing quantum field theoretic properties from the vantage point of the classical versions of these field theories. The benefit is not absolute: the ultimate goal is a thoroughgoing understanding of the relevant properties from the standpoint of quantum field theories (and later on, presumably, of their successors). But when we return to renew our efforts at interpreting quantum field theories as applied in particle physics, we will benefit from having first spent some time looking at these fundamental properties from the classical vantage point. I have illustrated the benefit of this approach by showing how the Lie-algebra-valued character of charge properties is made manifest in classical field theories. In the case of chromodynamics, this feature of charge is obscured in the quantum version of the theory as a result of confinement and color neutrality. By turning attention to the unconfined matter fields in classical chromodynamics, we uncover the $\mathfrak{su}(3)$ character of color charge.

Thus, from the classical vantage point, we can more easily see key features of charge properties that are preserved under the process of quantization while also becoming more keenly aware of those features of charge and its associated interpretive issues that are essentially quantum. We should, therefore, not discard classical field theories as irrelevant for philosophical investigation for being “the wrong theory.” Although they do not tell the full story of what current best physics has to say about the subatomic realm, classical field theories are a core part of the skeleton of the SM. Quantization (and many other theoretical and calculational procedures, e.g., renormalization) fill out the rest of the scientific achievement that is the SM. There is much we can learn about the entire body of the SM from investigating the “bones” themselves.

Acknowledgments

I would like to thank Dave Boozer, Mark Mace, John Norton, Jim Weatherall, and David Wallace for helpful discussions on topics discussed in the article. I would also like to thank Benjamin Feintzeig, Joshua Rosaler, and Jeremy Steeger for a stimulating symposium.

Footnotes

*

I would like to thank Dave Boozer, Mark Mace, John Norton, Jim Weatherall, and David Wallace for helpful discussions on topics discussed in the article. I would also like to thank Benjamin Feintzeig, Joshua Rosaler, and Jeremy Steeger for a stimulating symposium.

1 See Bokulich (Reference Bokulich2008), especially chapter 5.

2 See Bañados and Reyes (Reference Bañados and Reyes2016) for a recent review of Noether-type theorems. See also Kosmann-Schwarzbach (Reference Kosmann-Schwarzbach2011) for a comprehensive treatment, and see Olver (Reference Olver1993) for a standard mathematical presentation of the theorem.

3 See Maudlin (Reference Maudlin2007) and Gilton (Reference Gilton2021) for separate interpretive issues regarding the metaphysics (r, b, g; $\bar{r}$ , $\bar{b}$ , $\bar{g}$ ).

4 See Mace (Reference Mace2018, 2). At sufficiently high energies, free quarks and gluons within quark–gluon plasma become possible.

5 For this history of asymptotic freedom and infrared behavior, see Gross (Reference Gross2004). See page 205 therein for a discussion of Weinberg’s and Gross–Wilczek’s competing initial conjectures as to how these behaviors might have confinement as a dynamical consequence.

6 See Deur, Brodsky, and de Teramond (Reference Deur, Brodsky and de Teramond2016, sec. 4.6).

7 For that matter, QCD is not the final correct theory either. But it is the predictions of QCD, rather than any of classical chromodynamics, that currently enjoy empirical success.

8 See Weatherall (Reference Weatherall2016) for more technical details about formulating classical field theories using the machinery of fiber bundles, and see Gilton (Reference Gilton2019) for more on classical chromodynamics in particular.

9 In particular, it is unclear what physical significance we should attribute to geodesics in the total space. The mathematical derivation rehearsed here was first given by Kerner (Reference Kerner1968). It is also given by Bleecker (Reference Bleecker2013, chap. 10, sec. 1). Alternative derivations have been presented by Storchak (Reference Storchak2014), Sternberg (Reference Sternberg1977), and Weinstein (Reference Weinstein1978). Wong (Reference Wong1970) himself arrives at the expression by extracting a classical limit of the equations of motion for the quantum fields.

10 See Hamilton (Reference Hamilton2017, theorem 2.1.12).

References

Bañados, Máximo, and Reyes, Ignacio. 2016. “A Short Review on Noether’s Theorems, Gauge Symmetries and Boundary Terms.” International Journal of Modern Physics D 25 (10):1630021.CrossRefGoogle Scholar
Batterman, Robert W. 2009. “Idealization and Modeling.” Synthese 169 (3):427–46.CrossRefGoogle Scholar
Bleecker, David D. 2013. Gauge Theory and Variational Principles. Minneola, NY: Dover.Google Scholar
Bokulich, Alisa. 2008. Reexamining the Quantum-Classical Relation: Beyond Reductionism and Pluralism. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Cohen, I. Bernard. 1983. The Newtonian Revolution. Cambridge: Cambridge University Press.Google Scholar
Deur, Alexandre, Brodsky, Stanley J., and de Teramond, Guy F.. 2016. “The QCD Running Coupling.” Nuclear Physics 90 (1):195.Google Scholar
Gilton, Marian J. 2021. “Could Charge and Mass Be Universals?Philosophy and Phenomenological Research 102 (3):624–44.CrossRefGoogle Scholar
Gilton, Marian J. R. 2019. “Charge in Classical Gauge Theories.” PhD diss., University of California, Irvine.Google Scholar
Gross, David J. 2004. “Asymptotic Freedom and QCD—a Historical Perspective.” Nuclear Physics B—Proceedings Supplements 135 (1):193211.CrossRefGoogle Scholar
Hamilton, Mark J. 2017. Mathematical Gauge Theory: With Applications to the Standard Model of Particle Physics. New York: Springer.CrossRefGoogle Scholar
Hendry, Robin Findlay. 2006. “Is There Downward Causation in Chemistry?” In Philosophy of Chemistry, edited by Baird, Davis, Robert Scerri, Eric, and McIntyre, Lee, 173–89. New York: Springer.CrossRefGoogle Scholar
Kerner, Ryszard. 1968. “Generalization of the Kaluza-Klein Theory for an Arbitrary Nonabelian Gauge Group.” Annales de l’Institut Henri Poincaré 9 (2):143–52.Google Scholar
Kosmann-Schwarzbach, Yvette. 2011. The Noether Theorems: Invariance and Conservation Laws in the Twentieth Century. New York: Springer.CrossRefGoogle Scholar
Li, Bihui. 2013. “Interpretive Strategies for Deductively Insecure Theories: The Case of Early Quantum Electrodynamics.” Studies in History and Philosophy of Science Part B: Studies in History and Philosophy of Modern Physics 44 (4):395403.CrossRefGoogle Scholar
Mace, Mark. 2018. “Infrared Dynamics of Non-Abelian Gauge Theories Out of Equilibrium.” PhD diss., State University of New York at Stony Brook.Google Scholar
Maudlin, Tim. 2007. The Metaphysics within Physics. Oxford: Oxford University Press.CrossRefGoogle Scholar
Olver, Peter J. 1993. Applications of Lie Groups to Differential Equations. Berlin: Springer Science & Business Media.CrossRefGoogle Scholar
Seiler, Erhard. 2003. “The Case against Asymptotic Freedom.” Paper presented at Seminar on Applications of RG Methods in Mathematical Sciences, Kyoto, Japan.Google Scholar
Sternberg, Shlomo. 1977. “Minimal Coupling and the Symplectic Mechanics of a Classical Particle in the Presence of a Yang-Mills Field.” Proceedings of the National Academy of Sciences 74 (12):5253–54.CrossRefGoogle Scholar
Storchak, Sergey N. 2014. “Wong’s Equations in Yang-Mills Theory.” Central European Journal of Physics 12 (4):233–44.Google Scholar
Wald, Robert M. 1984. General Relativity. Chicago: Chicago University Press.CrossRefGoogle Scholar
Weatherall, J. O. 2016. “Fiber Bundles, Yang–Mills Theory, and General Relativity.” Synthese 193 (8):2389–425.CrossRefGoogle Scholar
Weinstein, Alan. 1978. “A Universal Phase Space for Particles in Yang-Mills Fields.” Letters in Mathematical Physics 2 (5):417–20.CrossRefGoogle Scholar
Wimsatt, William C. 1987. “False Models as Means to Truer Theories.” In Neutral Models in Biology, edited by Nitecki, M. H. and Hoffman, A., 2355. Oxford: Oxford University Press.Google Scholar
Wong, S. 1970. “Field and Particle Equations for the Classical Yang-Mills Field and Particles with Isotopic Spin.” Il Nuovo Cimento A 65 (4):689–94.CrossRefGoogle Scholar
Figure 0

Figure 1. Color states r, b, and g accounted for at each vertex within a Feynman diagram, here given for the process of nucleon scattering via pion exchange. Straight lines with arrows depict quark states, and curly lines with double coloring depict gluon states.