1. Introduction
The history of magnetism is very long and rich, with some of its earliest scientific discussions attributed to the Greek Thales of Miletus during the 6th century BC. And yet, even today, we do not have a full understanding of why there are two poles in magnets, and it is not even clear whether isolated magnetic poles (or magnetic monopoles) exist, despite much theoretical and experimental work on these questions [
1,
2,
3,
4,
5]. In 1269, the French scientist Pierre le Pèlerin de Maricourt, or Petrus Peregrinus of Maricourt, wrote his
Epistola de magnete, in which he described his observations on properties of magnets, such as the presence of two opposite poles in magnets (Maricourt introduced the term
polus), the repulsion between two like poles, and the fact that cutting a magnet in two halves results in two halves with two further opposite poles [
6]. Another breakthrough came in 1864 with James Clerk Maxwell’s theory of electrodynamics, based on the discoveries by many precursors such as Ampère and Faraday, and in which Maxwell simply assumed that magnetic charges did not exist, but did not forbid them either [
7]. At that time, and for a few years thereafter, the general opinion among physicists was that magnetic monopoles were not elementary particles that existed in reality. In 1894, however, Pierre Curie made the, then bold, suggestion that monopoles could exist in nature, even though none had been observed [
8].
It is common to credit Paul Dirac for the first modern investigation of magnetic monopoles, in 1931. For more references pre-dating Dirac’s contribution, see the reprint book [
9]. Not only did Dirac show that monopoles are compatible with quantum mechanics, but he also demonstrated that the existence of monopoles would explain the observation that electric charge is discrete (or ‘quantized’, as it is more commonly referred to); that is, a magnetic charge
g would imply that electric charges are given by [
10]
Seventeen years later, Dirac showed that a Lorentz-invariant Lagrangian containing monopoles should be non-local [
11]. About twenty years after that, Zwanziger wrote a local Lagrangian with an additional, ‘magnetic photon’ field, by expanding on former ideas by Cabibbo and Ferrari [
12]. However, he also needed a constraint in order to reduce the degrees of freedom to that of the on-shell photon, thereby requiring a space-like Lorentz-symmetry-violating four-vector associated with the direction of the Dirac string, itself an unphysical artifact [
1,
13,
14,
15]. (Hereafter, we shall not discuss ‘dyons’, hypothetical objects with both electric and magnetic charges, investigated in Refs. [
13,
16], and proposed as a phenomenological alternative to quarks in Ref. [
17]. Let us just mention that the Dirac quantization condition is generalized, given two dyons with respective electric and magnetic charges
and
, to
; limits on their charge and mass were recently established after the first search for dyons at the Large Hadron Colliders [
18].) An excellent review on current monopole searches is Ref. [
5].
At the time of writing, neither elementary magnetic monopoles nor dyons have been observed yet. An analysis of 13 TeV proton–proton collisions by the MoEDAL Collaboration, during the 2015–2017 Run 2 at the CERN Large Hadron Collider (LHC) led to mass limits in the 1500–3750 GeV range for magnetic charges, and up to 5
(where
is the minimum magnetic charge, called the ‘Dirac magnetic charge’, allowed by Equation (
1)), for monopoles with spin 0, 1/2, and 1 [
19]. The MoEDAL Collaboration also performed a search for dyons based on a Drell–Yan production and excluded dyons with a magnetic charge up to 5
and an electric charge up to 200
e for mass limits in the range 870–3120 GeV, and monopoles with magnetic charge up to 5
with mass limits in the range 870–2040 GeV [
18]. More recently, the ATLAS Collaboration (CERN) also reported on a search for magnetic monopoles and high-electric-charge objects during the LHC’s Run 2 and found no highly ionizing particle candidate [
20]. Further limits, if not a discovery, should be obtained after the LHC’s ongoing Run 3, which should be completed in 2026.
We can write the fully symmetric Maxwell equations with magnetic monopoles as follows:
For the sake of simplicity, hereafter we utilize units such that
. These equations can be written as
which are invariant under the electric–magnetic duality transformation, defined as the following complex phase rotations:
In this context of classical electromagnetism, this duality transformation was first observed by Heaviside [
21]. Afterwards, the concept of duality was extended and extensively investigated in a wide array of the physics literature. The appearance of imaginary numbers in the previous expressions might be seen as a hint for exploiting the analytical extension of U(1) hereafter.
The second expression of Equation (
2) appears to be incompatible with
if
. However, we can recover Equation (
2) by utilizing a
singular Dirac potential,
as a function of the position
, with the unit constant vector
, which is parallel to the Dirac string. Other, more suitable, expressions of this potential are discussed in Ref. [
1], along with discussions of incidental subtleties.
The continuity equations for electric and magnetic densities and currents are given, respectively, by
or, in tensor notation,
. In 1948, Dirac attempted to generalize electrodynamics in terms of these four currents,
and
, as well as the field strength tensor
and its dual
. However, despite these equations being elegant and symmetric, their expressions in terms of potentials require modified expressions of the field strength tensor in terms of the potentials. These, in turn, involve complications such as the ‘Dirac veto’, which states that the trajectory of electric charges in the presence of magnetic charges must not intersect the Dirac string. The duality, that is, the interchange between electric and magnetic objects, such that the possibility of attaching a Dirac string to an electric charge as well as a magnetic charge, led to dual-invariant electrodynamics Lagrangians in terms of two potentials: the four-vector
and the pseudovector
[
12,
13,
14]. Without going into detail, let us mention the field strength tensor and its dual from Ref. [
12],
which, whereas they do not require a Dirac string, pose the difficulty of introducing new degrees of freedom, which can be dealt with in various ways. Further discussions on the Cabibbo–Ferrari two potentials are given in Refs. [
22,
23,
24]. Other constructions of models with two potentials are based on a local Lagrangian formulation by Zwanziger, where the electromagnetic field strength tensor and its dual can be written as [
13,
14]
where the vector
is spatially parallel to the Dirac string but with the physics being independent of
.
The central objective of this paper is to provide a symmetry argument which leads naturally to the addition of a second potential which will account for the existence of magnetic charges, rather than imposing that second potential by hand. The resulting approach is covariant, valid in any curved spacetime, and should accommodate two-potential Lagrangians of monopoles; note, however, that models that break Lorentz invariance, such as Zwanziger’s Lagrangian with its Dirac string vector, might involve slight modifications to the formalism. Hereafter, we apply an analytical extension of the parameter space of electrodynamics’ U(1) gauge transformations and obtain thereof two potentials which describe elementary particles with electric charge and magnetic charge, the resulting gauge group being effectively the product of the compact Lie group U(1) with the non-compact Lie group of dilations
. Although the literature on duality is abundant, to the best of our knowledge this is the first attempt to explain the appearance of the dual gauge field by modifying the gauge symmetry group. We thus build gauge theories of magnetic monopoles from first principles: along with the U(1)
symmetry, so that in addition to the usual (electric) gauge field
, there appears a magnetic gauge field
, as in Refs. [
23,
24]; we shall also require conformal invariance. Although this may suggest a similarity with the Weyl approach to gravity, we shall explain how our approach differs from the Weyl theory.
We examine the gauge invariance by means of Utiyama’s methodology [
25,
26,
27,
28,
29]. This approach considers infinitesimal gauge transformations and leads to a set of equations under the assumption of independence of the gauge parameters and their derivatives. The solutions of these equations determine the relevant objects that should be considered in a deductive way, such as the covariant derivative and field strength. We illustrate this approach with a massless scalar field as well as with a spinor field. We comment on the two-potential monopole phenomenology at the end of
Section 4.2. We observe that for spinor models a non-symmetrized Lagrangian enables the existence of magnetic monopoles, but this is not the case when we consider a symmetrized Lagrangian. These subtleties do not affect the scalar case, where the magnetic monopole’s appearance is clear.
2. Analytical Extension of the U(1) Gauge Symmetry
Hereafter, we examine and motivate from symmetry arguments the following expressions of
and
in terms of vector and scalar potentials,
,
,
, and
:
Clearly, their symmetry allows for point-like magnetically charged particles, just as we usually have with the potentials
and
A, only for electrically charged particles. In
Appendix A, we briefly discuss the gauge invariance conditions behind Equation (
3), and we write these field equations in terms of the gauge fields.
As we shall explain hereafter, the essence of this section lies in the generalization of the U(1) symmetry by applying an analytical extension,
so that, for instance, a scalar field would transform
globally as
This amounts to the product of a usual (compact) U(1) phase transformation,
, of the field, by a (non-compact) dilation,
, and, as we shall describe, the invariance of the Lagrangian also requires specific conformal transformations related to this dilation factor. As alluded to in the Introduction, this may suggest to some that we are actually constructing Weyl’s invariant theories of gravity, but in
Appendix B, we explain how our approach is different.
In
Section 2.1 and
Section 2.2, we examine the effects of the analytical extension of the U(1) gauge transformations on the matter field, with their corresponding invariant Lagrangians, and the analogue for the gauge fields, along with their related field strength tensors.
2.1. Matter Fields
In order to describe the appearance of the magnetic gauge field
, let us consider the Lagrangian of a massless complex scalar field,
, with a quartic potential:
The quartic interaction is a well-studied non-trivial addition to the free theory terms; it follows naturally the quadratic massive term (which vanishes here, in order to maintain global conformal invariance). But first, let us describe its invariance under the global U(1) transformation with analytical extension. We recover the invariance of
L if, in addition to the transformation on
,
, we also perform a transformation on the metric tensor; that is, we also perform a rigid conformal transformation on the metric tensor:
where
and
are constant, and
n depends on the Lagrangian. Note that whereas the general theory should be developed in a non-Minkowskian spacetime, the Minkowski case can be seen as a gauge fixing of the metric tensor. For instance, if we start with a non-flat metric
that is conformal to
, then the parameter
can be chosen in such a way that
. (We note that the conformal invariance of Maxwell equations in the presence of magnetic monopoles has been verified in Ref. [
30].) In this way, we see that
For the scalar field described by Equation (
5), it is clear that we recover the invariance of the Lagrangian by choosing
, and then,
.
In the late 1970s, Kyriakopoulos showed the invariance of the Maxwell equations with monopoles under conformal transformations [
30]. That work inspired us to study the role played by conformal transformations in the candidate theories for magnetic monopoles. In particular, one might ponder about its influence at the level of a fundamental symmetry, such as gauge symmetry. One could wonder if the duality transformation,
, would also imply an analogous relation for the potentials involving complex numbers, e.g.,
. If that was the case, we could argue that the U(1) symmetry was thereby extended to a symmetry containing a real-phase factor and a complex phase. Moreover, this somewhat curious idea of extending the parameter space with an imaginary factor appears to explain in a natural way the otherwise ad hoc appearance of the additional, so-called ‘magnetic’, potential. In this way, one could see the gauge transformation parameter as being generalized to produce an extended symmetry group such as
. Magnetic monopoles have not been observed yet in the forms proposed in the literature. Hereafter, we suggest a formulation from first principles (gauge theory) aiming to predict monopoles and to offer a possible explanation for the reason why these particles remain elusive to detection.
Now, let us extend the U(1) gauge theory for this type of transformation and see how the magnetic potential results from the analytical extension in Equation (
4). We begin with a Lagrangian of
N scalar fields
, with
:
(Were we working with other tensor fields, the standard derivative should be replaced by a spacetime covariant derivative.) We apply the transformation
where G is the gauge group,
are real spacetime functions, and
are representation matrices of
, the generators of G, with commutation relations
In order to avoid confusion, the group indices
a are displayed between parentheses, as in Equation (
6). The Lagrangian is assumed to be invariant under transformations of the fields:
Henceforth, we will work with an abelian group; that is, such that
, with
(where the
i factor is incorporated within the Lie algebra representation of
) so that the transformed fields read
in which
now denotes the representation matrices of the corresponding Lie algebra element on the
field. This should not cause confusion in the remainder of this paper.
Here, again, the spacetime metric also transforms as
where
n is determined according to the form of the Lagrangian. (For instance, for the massless complex scalar field Lagrangian
and the Lagrangian will be invariant if
. Similarly, if
then the Lagrangian will be invariant if
.)
Next, if we take the group parameters,
and
, to be spacetime dependent, we necessitate gauge fields in order to preserve gauge invariance. Since we have two independent parameters, we shall introduce two gauge fields,
with the transformation law:
for
. As usual, these additional fields restore the invariance of the Lagrangian lost due to the derivatives
.
We insert these fields within the initial Lagrangian,
, which results in an invariant Lagrangian:
The variation under infinitesimal transformations is written as [
25]
If
is invariant under these transformations with
,
,
, and
considered as independent parameters, then we obtain the symmetry equations:
The last equation is satisfied as long as the dependence of the Lagrangian with derivatives of the fields
and the gauge field is expressed via covariant derivatives,
which implies that the gauge-invariant Lagrangian should be extended as
The transformation law of this object is described by
which ensures that this object is covariant under the symmetry group with generators
, with
dim(G).
Naturally, a question that occurs at this point is whether the resulting models reflect the expected magnetic–electric duality. We respond in the affirmative in
Appendix C, for any curved spacetime, by making use of the Levi–Civita form defined in terms of the well-known Levi–Civita symbols. In addition, while recovering the duality property we will encounter another motivation for exploiting the analytical extension of the parameter space, Equation (
4), which underlies our approach.
2.2. Gauge Fields
Here, we analyze the Lagrangian
of the free gauge fields. For the functional dependence of
, as usual, we expect it to depend on the gauge fields and their first derivatives. In addition, since we are dealing with vector fields, the presence of the metric tensor is almost mandatory, otherwise we would not be able to build scalar quantities out of vector fields. Nonetheless, since the metric is not flat a priori, the derivative of the gauge fields should take into account the curvature of the spacetime; in other words, the derivative of the gauge fields should actually appear in
by means of the spacetime covariant derivative. In summary, we propose the functional dependence of
as
where
, with
being the Christoffel symbols. As a matter of consistency with the remainder of the theory, this Lagrangian should be invariant under the gauge transformation of the metric and the gauge fields. Accordingly,
The transformation law for the Christoffel symbols can be straightforwardly evaluated and results in
When we impose invariance of the Lagrangian,
, and assume the independence of the parameters
and
and their derivatives, we find the equations
Note that the last three equations consider the symmetry of the lower indices of the Levi–Civita connection and the symmetry of the second derivatives, i.e.,
. By substituting the last equation into the previous two equations, we can simplify these three equations as follows:
The solutions to the last equations show us that
where
are the field strengths for both
and
. In other words,
does not depend explicitly on the gauge fields, but depends on the metric and on the field strengths
. But there is also an extra condition for the Lagrangian of the free gauge field. The first of the equations can be rewritten in terms of a new object
, leading to
This expression tells us that the energy–momentum tensor
associated with
has to be traceless.
5. Concluding Remarks
This paper pursues a well-known attempt to circumvent the non-local nature of Dirac-type Lorentz-invariant Lagrangian descriptions of magnetic charges. That idea, originally due to D. Zwanziger, involves not one but two gauge potentials, the second one being ascribed to the magnetic charge (of the monopole or dyon). This leads to
local Lorentz-invariant Lagrangian descriptions of magnetic charge. Such models were examined quite recently in Ref. [
34]: its authors investigated the existence of spurious poles and showed, among other things, that the amplitude for single-photon production of magnetically charged particles by electrically charged particles is equal to zero.
The purpose of the present paper was to propose an elegant motivation, based on symmetry arguments, for the existence of the second ‘magnetic’ gauge potential. As far as we know, that is not found in the literature. We proposed that the magnetic potential appears via an analytical extension of the usual U(1) group underlying standard electromagnetism with electric charges only, resulting in a non-compact two-dimensional group. We implemented this extension via the Utiyama methodology, which considers infinitesimal gauge transformations and provides conditions due to the independent gauge parameters and their derivatives. These conditions allow us to deduce relevant objects such as the covariant derivative and field strength.
As examples of this approach, we examined a model with a massless scalar field and a model with a massless spinor field. For spinor models, we pointed out that a non-symmetrized Lagrangian enables the existence of magnetic monopoles, but that this is not the case with a symmetrized Lagrangian. These subtleties do not occur with the scalar field. Also, we examine massless spinor fields, and if we add a mass term, in order to describe real leptons and quarks, we find that our model is no longer invariant, since we cannot find a value for n that makes both the kinetic and massive terms invariant. The masslessness of the fermions is not necessarily a problem since the electroweak interaction associated with the SU(2) × U(1) gauge symmetry also demands the same condition before spontaneous symmetry breaking, through which the generation of mass for quarks and leptons is provided by the Higgs mechanism. The same could be applied in the present case.
Another example that could eventually be considered is the spin-one field. However, the standard kinetic term of the Lagrangian, is not invariant under our extended gauge group. This means that the magnetic monopoles will not interact with spin-one fields. Physically, charged spin-one fields could be associated with the vector bosons . From this perspective, no experiment involving these particles would be useful to detect magnetic monopoles in the context of our proposal.
To sum up, the detection of magnetic monopoles in the context of the present work is more likely to be achieved with experiments involving scalar fields, which essentially means that they should consider a flux of Higgs particles. As stated before, this may eventually be achieved with a new generation of accelerators. The investigation of how this measurement could be made in experiments, such as MoEDAL and others, is something to be implemented in future works. In this regard, an extension to non-abelian groups accommodating magnetic monopoles could also be attempted. We intend to implement this in a separate paper.
When dealing with particles of the standard model, one usually starts with a gauge group in order to deal with the electroweak interactions (or eventually with if we also include the strong interaction), where the particles have no masses. This massless character is essential in order to preserve the symmetry. The particles then acquire mass through the Higgs mechanism, which generates mass for both the leptons and the gauge bosons. In the literature, several models predict the existence of mass for the magnetic monopole. This may eventually happen in our model as well. However, in order to properly address this point, we should extend our analysis to unify our extended group with the group. We expect that the use of the Higgs mechanism may eventually result in a mass for the gauge boson associated with the magnetic monopole. This analysis lies beyond the scope of this paper, but at this point it is not possible to know if the magnetic boson will acquire mass or not and even less to estimate an order of magnitude for it. Some experimental constraints establish a lower limit of a few TeVs for the mass of the monopole. It would be interesting to check if this will be the case for our model, and also if we will face problems with the mass hierarchy or some fine tuning. We intend to consider this elsewhere.