<<

arXiv:gr-qc/0405046v2 19 May 2004 h oko hc hsppri ae a enpope yarec a by Li prompted the of been form generalized has a how based showed he is which in paper Newman this Ezra which on work The Introduction 1 hterirwr oei hsdrcincudb ofimdan confirmed be could direction this a in be done work electr would actual earlier particle of that properties a known all such essentially has Since co which which particle [1]. charged moment compl a magnetic-dipole of into world-lines extended the equations to correspond Maxwell’s to solution fields, hrfr opoiealn ewe lmnaypril phy particle elementary theory. between element relativity link other a of provide currents to interaction therefore the yield to interpretatieralized field-theoretical a Neverth given be interpretation. can geometric operations of simple a whi have world-line, to the seem on not operation another by obtained is the tron On characte singularity current. magnetic-dipole interaction electric-monopole pion-proton may pseudo-scalar it a that so as im it, to into component singularity magnetic-dimonopole translat a electric-monopole adding the bare that a show of spaceti We world-line complex into electrodynamics. extended Lanczos-Newman field Maxwell complexified the edsustepyia aueo lmnaysingularities elementary of nature physical the discuss We iglrte nLanczos-Newman in singularities ntepyia nepeainof interpretation physical the On o 0 H11 eea1,Switzerland Geneva-12, CH-1211 30, Box needn cetfi eerhInstitute Research Scientific Independent SI0-42 coe 6 2007 16, October ISRI-04-04.20 electrodynamics -al [email protected] e-mail: nr Gsponer Andre Abstract 1 nwihcnb gen- be can which on itco ia elec- Dirac a of ristic n,i so ra interest great of is it ons, r atce,and particles, ary hhwvrdoes however ch gnr pc is space aginary isadgeneral and sics l aea intrinsic an have uld xadd[,3 4]. 3, [2, expanded d ls,bt type both eless, te ad the hand, other xsaeie may spacetime, ex einterpreted be e .. in i.e., me, o fthe of ion enard-Wiechert ´ ia electron, Dirac rsn in arising n ril of article ent However, as stressed by Newman, there are great difficulties, both techni- cal because of the complicated inherited from the general relativity considerations which gave rise to his new approach to the problem of electricity in curved spacetime, and physical because the interpretation of the singularities associated with arbitrary world-lines in complex spacetime is not trivial, even in flat . In this paper we focus on the second class of difficulties, mainly because they are closestto theauthor’sdomainof expertise, i.e., elementary particle physics, and because the problem of embedding electrodynamics and particle-physics into gen- eral relativity is a longstanding unsolved problem, to which Newman’s approach might provided a solution if its physical interpretation is fully clarified. This paper is therefore of a cross-disciplinary kind, implying that since its goal is to interpret Newman’s world-lines in the perspective of their possible relevance to elementary particle physics, we use a language and a methodology different from his. In particular, instead of congruences of “world-lines” in complexi- fied Minkowski space, we deal with “poles,” i.e., “point-like” singularity of the complexified Maxwell field,1 which should be, at least in the simplest cases, in one-to-one correspondence with the world-lines by means of special or general Lorentz transformations. Therefore, as a first step, we mostly work in the non- relativistic limit, which should retain the distinctive qualitative features of the singularities arising in relation to Maxwell’s and Dirac’s equations expressed over complexified spacetime, and thus be sufficient to interpret and classify them. Inotherwords, weassumethattheelementaryparticlesknown from experiment could correspond to poles in complexified spacetime, and that at least the simplest such particles — namely the electrons, nucleons, and pions — could in some appropriate low-energy limit2 correspond to poles of the Maxwell and Dirac fields expressed over complexified spacetime.3 1Such poles might be simple spherically-symmetric distributions arising from the Coulomb- like potential of electric or magnetic monopoles, or from the Yukawa potential of a pion; or more complicated distributions arising, for example, from the axially-symmetric intrinsic magnetic- dipole of electrons; as well as combinations of such poles, e.g, the electric-monopole magnetic- dipole combination characteristic of a Dirac electron. 2By low-energy we mean energies E that are small in comparison to the mass of the heaviest particles under consideration, i.e., the nucleons. Therefore E ≪ 939MeV ≈ 1GeV, i.e., energies that are very low in comparison to current elementary particle research, which is done using accelerators with energies on the order of 1000 GeV. For this reason, the “parton aspect” of elementary particles (i.e., their modeling in terms of quarks, gluons, etc.), which is the main focus of present-day high-energy physics, is not essential in the energy domain considered in this paper. 3As is well known, the assumption that in the low-energy limit the nucleons and pions are “whole particles,” just like electrons and photons, is in good agreement with experiment, and gives a satisfactory description of essentially all of atomic and nuclear physics. In particular,

2 Fortunately, this idea is not new, and was systematically investigated for the first time by Cornelius Lanczos in 1919 in his PhD dissertation[7, 8]. In thisthesis he postulated that electrons and protons could be singularities of the Maxwell field expressed in theformalism of complex quaternions,and showed that this formalism is mostappropriate to study such singularitieswhen they are analytically continued into complexified Minkowski space. In particular, to deal with the problem of infinities which automatically arise when integrating quantities associated with even the most simple singularities, Lanczos had the idea of shifting the origin of the Coulomb potential pole by a small imaginary amount into complex 3-space, so that the singularity from point-like would become a circle. This “circle electron,” whichasobservedbyLanczoshasthenicepropertyofyielding afiniteHamiltonian function, will be rediscovered by others, for instance by Newman who saw in it a possibility to endow a bare Coulomb monopole with a magnetic moment [2, 3, 4].

However, since the Coulomb potential ϕmonopole = e/r is just a scalar in the rest-frame, the only effect of shifting its origin by an imaginary amount is to add an imaginary part to it, and therefore to make it complex instead of real. While this complexified scalar potential leads to an electromagnetic field having some of the characteristics of a magnetic dipolar field, it is nevertheless unsuitable to yield the correct field of an intrinsic magnetic dipole, because such a field derives from a vector potential A~dipole, not from a scalar potential. Hence, the interpretation of the singularities obtained by displacing a world-line into complexified spacetime is more complicated than anticipated by Newman. As will be shown in Section 12, what is actually generated by means of Lanczos- Newman’s infinitesimal “imaginary scalar translation” is a pseudo-scalar field which could possibly be associated with the pion field of a nucleon, while the magnetic dipole field characteristic of a Dirac electron is obtained by another type of transformation: an infinitesimal4 “imaginary quaternion operation” which is adding a vector part to Coulomb’s scalar potential so that it becomes the 4- potential of the electromagnetic field of a Dirac electron which, in its rest-frame, is the e ~r × ~µ A = ϕ − iA~ = − i . (1.1) D monopole dipole r r3

In order for such operations to be possible and meaningful it is necessary to contemporary theory of nuclear forces is still essentially that of the 1950s, with the difference that the constraints coming from its high-energy limit (in which quark and gluon effects, i.e., quantum chromodynamics, QCD, dominate) is now better understood [5, 6]. 4The magnitudes of the transformations considered in this paper are infinitesimal in the sense that they correspond to lengths characteristic of elementary particles, i.e., to their classical elec- tromagnetic radius, e2/mc2, or to their Compton wave-length, ~c/mc2, which are on the order of 10−15 m.

3 write Maxwell’s field as a complex field over complex spacetime, and to correctly associate the poles and sources of that field with elementary particles. This requires a theory in which both Maxwell’s and Dirac’s equations are expressed over the same space, i.e., everyday’s Minkowski spacetime extended to complexified spacetime by analytic continuation, and a common formalism whichcan handlethe photon, electron, and possibly other particle’s fields consistently. Such aformalism has been developed by Cornelius Lanczos in 1919 for Maxwell’s equations [7, 8], and in 1929 for Dirac’s equations [9, 10, 11, 12].5 Lanczos’s formalism is briefly reviewed in Sections 2 and 3. It is based on Hamilton’s biquaternion (i.e., complexified quaternion) algebra B, which is defined by the non-commutative product6

(a + A~)(b + B~ )= ab − A~ · B~ + aB~ + bA~ + A~ × B~ . (1.2) where a, b ∈ C are complex scalars and A,~ B~ ∈ C3 are complex vectors.7 Using this unique multiplication rule, which provides a very straightforward and gen- eral formulation of for both integer and half-integer spin (i.e., bosonic and fermionic) fields, it is possible to formulate not just Maxwell’s and Dirac’s theories, but the whole of relativistic quantum field theory in a concise and consistent way which avoids many of the complications required by the standard formalism [14, 15]. Moreover, while it is well known that basic 4-vectors such a the 4-position X = ict+~x, or the 4-momentum P = E +ic~p, can mostnaturallybe written as , it is important for the purpose of this paper that it appears that the fields and singularities associated with all known elementary particles can also be expressed as biquaternions, e.g., (1.1). Sections 4 to 10 are concerned with a simplified discussion of the elementary singularities which appear in a biquaternion field theory, namely a discussion of the singularities which pertain to real or bireal quaternions8 written in spherical coordinates, and whichare essentiallythe same as, or combinationsof those, which arise in the theory of real scalar and vector functions over R3. This part of the paper, which could be the subject of a separate article, is mostly mathematical and was written because the author was unable to find a paper in which this question is 5In these papers Lanczos also derived the correct equation for a massive spin-1 particle, which would be rediscovered in 1936 by Albert Proca, see [9, eq.98], [11, eq.23], and [12]. In 1958 Feza Gursey¨ showed that Lanczos’s formalism is directly applicable to strong interactions as well [13]. 6The quaternion product is a direct generalization of the complex number product, which was defined by Hamilton as (a + iA)(b + iB)= ab − AB + iaB + ibA. 7For clarity we conform as much as possible in the sequel to the “no-hidden-i” rule, and therefore use the arrow symbol exclusively for real vectors. See, e.g., (2.1). 8I.e., such that Q = Q∗ or Q = Q+, so that they are of the form s + ~v or s + i~v where s and ~v are real.

4 dealt with in an explicit and systematic manner, even though everything seems to be well known to and physicists who are well acquainted with the theory of distributions.9 For this reason, in contrast to other sections of the paper, this part may appear as rather detailed and low-level to some readers. Nevertheless, the fact that a magnetic dimonopole (a scalar-potential singularity) produces the same magnetic field as a magnetic dipole (a vector-potential singularity), except at the origin where the two singularities produce different delta-function-like fields, suggests that a somewhat starchy discussion could be useful, especially since this has the advantage of making the paper essentially self-contained with regards to the use of distribution theory in the present context. The final part of the paper, Sections 11 to 14, deals with the physical inter- pretation. In Section 11 we focus on the physical characteristics of the Dirac electron, showing in particular how the singularity (1.1) appears in Dirac the- ory. In Section 12 we show that the Lanczos-Newman translation leading to the “circle electron” singularity yields a pseudo-scalar-potential rather than the vector- potential of Dirac’s electron magnetic-dipole field, and derive the operation which yields the correct potential for this field. In Section 13 we give a field-theoretical interpretation of the scalar and quaternion operations which lead from the bare electric-monopole current of classical electrodynamics to those of the proton-pion and electron-photon interactions of quantum theory. In the concluding section we return to Newman’s papers to discuss the relevance and possible generalization of the results presented here, and to stress the amount of work which remains to be done in order to fully assess the value of the Lanczos-Newman approach to the problem of the essence of elementary particles in special and general relativity.

2 Lanczos’s formulation of Maxwell’s equations

The fundamental quantity in Lanczos’s biquaternion10 formulation of Maxwell’s equations is the Silberstein-Conway electromagnetic field bivector B which com- bines the electric and magnetic fields in a single complex vector (which transforms as a 6-vector) B := E~ + iH~ . (2.1) 9After these sectionswere written I was informedthat somewhat similar questions and solutions have recently been discussed in the context of classical electrodynamics [18, 19]. 10Hamilton introduced the prefix “bi-” to indicate that some quantity is complexified — which is why Hamilton called a complex number a “biscalar.”

5 This bivector should not be confused with the Faraday-Kilmister electromagnetic field tensor, which in biquaternions corresponds to the linear function

1 + F ()= 2 ([ ]B − B [ ]) . (2.2)

Maxwell’s equations are then [7, 14]

∇ ∧ A = B , ∇B = −C , (2.3) where A is the 4-potential (which transforms as a 4-vector)

A := ϕ − iA~ , (2.4) and C the non-rationalized source 4-current density, which is related to the usual rationalized charge-current density

J := ρ − i~j (2.5) by the equation C := 4πJ . (2.6)

In equation (2.3) the operator ∇ is the four-dimensional generalization of Hamilton’s differential operator ∇~ , i.e., ∂ ∇()= []+ ∇~ [ ] , (2.7) ∂ict and the operator ∧ means that the scalar part is discarded so that B is a bivector as it should be according to (2.1). By keeping all parameters real, and interpreting i as a device to get Minkowski’s metric with Hamilton’s real quaternion product, equation (2.3) can be seen as just another way of rewriting Maxwell’s equations in a form which is equivalent to the more standard tensor, vector-algebra, or differential-form formulations. However, if we accept the complexification of the potential and source 4-vectors, as well as any kind of transformation on or between their scalar and vector parts, including (possibly complex) transformations of the coordinates, we get generalized types of electric and magnetic fields which are considered as “unphysical” in the standard interpretation of Maxwell’s field as a real field over real Minkowski space. In this paper we intentionally retain these generalized fields, and interpret them in terms of experimentally known elementary particle fields. The reason why such an interpretation makes sense is that, as discovered by Lanczos [9, 10, 11], there is a very close relationship between the complexified Maxwell field, and the usual

6 Maxwell, Dirac, Proca, and Yukawa fields, which can all be interpreted as distinct subfields of this generalized field [12, 17]. Fortunately, despite the inherent complexity of this relationship, much of the discussion can be simplified by considering stationary fields in the rest frame of the particles. This is why very little of the quaternion formalism is actually used in this paper, and why in most of it we just work in the standard vector algebra formalism of classical electrodynamics.

3 Lanczos’s formulation of Dirac’s equation

SoonafterthepublicationofDirac’sfamouspaperinwhichthequantumfieldofthe electron was expressed in an abstract space using four-dimensional matrix algebra, Lanczos showed that Dirac’s equation could be expressed in everyday’s spacetime just like Maxwell’s equations, and that working solely with vector algebra in its original form, namely biquaternions, one could dispense of any sophisticated formalism such as the cumbersome “Dirac matrices.” In fact, as the titles of his three papers suggest, The tensor analytical relationships of Dirac’s equation [9], The covariant formulation of Dirac’s equation [10], and The conservation laws in the field theoretical representation of Dirac’s theory [11], Lanczos was motivated by showing that Dirac’s equation had a perfectly sensible and straightforward tensor interpretation, just like Maxwell’s, Einstein’s, and all other fundamental equations of physics. Moreover, Lanczos showed that the Maxwell and Dirac fields were just two distinct11 subfields of a more general field, which he could not interpret. But, with todays hindsight, we recognize now that the full Lanczos field is exactly what is needed to express the physics of photons and electrons together with their weak- isospin partners, namely the massive weak-interaction intermediate-bosons and the massless neutrinos, as well as their interactions, as given by today’s “Standard Model” of electro-weak interactions [17]. Moreover, as shown by Feza Gursey,¨ Lanczos’s field can easily accommodate strong interactions, at least in the low- energy limitin whichtheseinteractions are givenby thecharge-independent theory based on pseudo-scalar pion fields [13]. In Lanczos’s representation the Dirac field is a biquaternion D ∈ B satisfying 11For a concise discussion of the similarities and essential differences between Maxwell’s and Dirac’s fields see [16].

7 the Dirac-Lanczos equation [9, eq.63] 1 α ∇D = D∗i~ν + ADi~ν . (3.1) ℓ e Here A is the electromagnetic 4-potential (2.4), α = e2/~c the electromagnetic coupling constant, ℓ = ~/mc the Compton wave-length associated with the mass m, and ~ν an arbitrary unit vector.12 While it is fully equivalent to other forms of Dirac’s equation, the main value of Lanczos’s form is the presence of the complex conjugation on the right hand side, which explicitly shows that Dirac’s and Maxwell’s equations are fundamentally different becauseit implies that Dirac’s field is fermionic, and therefore obeying Pauli’s exclusion principle, rather than bosonic as Maxwell’s [16]. For the purpose of this paper we do not need more than Dirac’s bilinear covariant quantities,13 that is, Dirac’s conserved 4-current14 C and spin pseudo-4- vector Σ C = eDD+ , Σ= D~νD+ , (3.2) Dirac’s invariant scalar s and spin 6-vector S s = DD, S = D~νD , (3.3) as well as the so-called Gordon decomposition of C, which is a trivial algebraic identity deriving from (3.1) and (3.2) that we write in the general form given by Gursey¨ [20, p.50-52], i.e.,15 1 1 C = sαA − ε(∇) − M(∇) . (3.4) ℓ 2 In this decomposition the first two terms correspond to the convection current, which would exist even if the electron had no spin, and the last one to the polar- ization current, which is due to spin. The function ε()= e[ ]hDi~νDi (3.5) 12This unit vector defines a axis relative to which quantities such as the spin are quantized. Its arbitrariness correspondsto the freedomin the choice of a set of anticommuting matrices in Dirac’s representation. The presence of this non-scalar factor in the Dirac-Lanczos equation insures that 1 the field D is of spin- 2 exclusively, and therefore cannot be of spin-1 as Maxwell’s for example. 13In Lanczos’s formalism the 16 bilinear covariant quantities of Dirac’s theory (two 4-vectors, one invariant biscalar, and one 6-vector) can immediately be written down in explicit form without the need of operators as in Dirac’s matrix formalism. 14In this paper we write Dirac’s conserved probability density 4-vector as an electromagnetic current density, which is why we include the electric charge e in its definition. Charge-current density conservation is then expressed by the identity ∇ · C =0. 15The Gordon decomposition in its traditional form is restricting the terms on the right-hand- side to their bireal (i.e., “Hermitian”) part, which has the effect of specializing the Dirac equation to electromagnetic interactions, even though Dirac particles such as the proton may also have non-electromagnetic interactions.

8 is the electric convection scalar,16 and the function M()= eDi~νD[ ] (3.6) the electromagnetic moment bivector. Note that we use the same symbol C for Dirac’s current (3.2) as for the source current in Maxwell’s equations (2.3). This is to suggest that this is precisely how the Dirac and Maxwell equations are coupled, namely by having common A and C. Therefore, in the absence of an external field, the potential Aself solution of (2.3) for the current C is the self-field A = Aself which in the Dirac equation (3.1) leads to that current. In the case where there is a non-vanishing external field Aext, the potentialA in (3.1) is the sum A = Aself + Aext.

4 Distributions in polar coordinates

In this section we present a general method which uses distribution theory17 to deal with functions that are singular at the origin of a spherical polar coordinates system. This method is based on Tangherlini’s for differentiating at the origin in polar coordinates [21, p.511–513]. While it is equivalent to those used in standard text books for dealing with simple cases such as the Newton or Coulomb fields, and is occasionally used for solving problems such as finding the Green’s function of electrostatics [23, p.51], this method allows a straight-forward and consistent treatment of the more complicated poles which are studied here. We consider differentiable scalar- or vector-valued function over R3, i.e., in general, quaternion-valued functions18 ~x −→ F (~x)= s(~x)+ ~v(~x) ∈ H . (4.1)

We suppose that F is everywhere regular except possibly at one point ~x0 which 19 we take as the origin of a polar coordinate system: ~x − ~x0 = ~r(r,θ,φ). We now interpret the function F (~x − ~x0) as a distribution in the variable ~r. We therefore write F (~x − ~x0) := F (~r)Υ(r) , (4.2) 16The underline specifies the range of action of the operator inserted in the square brackets. 17For an introduction see [22, p.766–798], and for a modern presentation [23]. 18The use of quaternions,i.e., the original form of vector calculus, is most appropriate here since it allows a simultaneous treatment of both scalar and vector functions. However, the generalization to biquaternions (i.e., complex quaternions), or even to functions over C3, are not trivial since the structure of singularities can be substantially more complicated. 19We assume r = |~r| and the usual parametrization such that θ ∈ [0, π] and φ ∈ [0, 2π], so that r ≥ 0, contrary to Tangherlini who allowed for r< 0.

9 where F (~r) ∈ C∞ over R3−{0}, while F (~r)Υ(r) is assumed to be infinitly differentiable in the sense of distributions over R3, as well as “weakly convergent” which means that for any “test function” T (~r) we have20

d3Ω F (~r)Υ(r)T (~r) ∈ H . (4.3) ZZZ R3

The scalar distribution Υ(r) has been introduced in (4.2) to specify how to differ- entiate at the origin.21 It is postulated to have the properties

Υ(r)=1 , ∀r> 0 , (4.4) and d3Ω F (~r)Υ(r)= d3Ω F (~r) , ∀F (~r) ∈ H , (4.5) ZZZ ZZZ which due to (4.3) imply that F (~r)Υ(r) and F (~r) are equivalent with regards to weak convergence and therefore correspond to the same distribution. However, the value Υ(0) does not need to be known, or even to exist, as long as (4.2) is weakly convergent. In fact, the only property that matters is that Υ(r) is discontinuous at r =0. In the theory of distributions this means that its derivative is of the form22 d Υ(r)= a δ(r) (4.6) dr where a ∈ R is a constant that we expect to be finite, and δ(r) the Dirac δ-function defined as the distribution 1 ε δ(x) := lim (4.7) ε→0 π ε2 + x2 which is normalized so that

+∞ dx δ(x)=1 . (4.8) Z−∞ In order to determine a we require that the Gauss theorem in its various forms should be true for any distributionsuch as (4.2), i.e., that for any simply-connected 3-volume Ω bounded by the 2-surface Σ= ∂Ω we have

d2Σ F (~r)Υ(r) ≡ d2Σ F (~r)= d3Ω ∇~ F (~r)Υ(r) . (4.9) ZZ ZZ ZZZ 20A test function is an infinitly differentiable bounded function with compact support. 21Note that we write Υ(r) rather than Υ(~r). This is an implicit postulate that a distribution Υ(r) which is symmetrical about the origin is adequate to deal with point-like singularities. 22Strictly speaking, this derivative could be a finite linear combination of δ(r) and its derivatives.

10 We therefore take the most prominent example of a physical distribution, the Newton or the Coulomb field, which we write ~x − ~x ~r 0 (4.10) F (~x)= 3 = 3 Υ(r) . |~x − ~x0| r Then, taking for Ω a sphere of radius R, and using (4.5), equation (4.9) becomes ~r ~r R ~r ~r 2 2 2 2 (4.11) d ωr · 3 = d ω dr r · 3 aδ(r) ZZ r r ZZ Z0 r r which yields the condition

R 4π =4π dr aδ(r) . (4.12) Z0 However, since the δ-function is even, Dirac’s normalization (4.7) implies that R 1 dx δ(x)= , ∀R> 0 (4.13) Z0 2 which, in accord with Courant and Hilbert, is the stipulation to be used to express multi-dimensional distributions in polar coordinates [22, p.791]. Thus, a =2, so that finally d Υ(r)=2δ(r) . (4.14) dr Therefore, as found out by Tangherlini by another method, Υ(r) is not the Heaviside step function H(x), but the so-called “sign step function” which can be represented by the sequence23 2 x Υ(x) := lim arctan( ) . (4.15) ε→0 π ε

In conclusion, by means of the definition (4.2), where the Υ-function stipulates how to differentiate at the origin, it is possible to make all calculations with distributions in polar coordinates in a systematic and consistent manner.24 In particular, as shown in the Appendix, it appears that the Gauss theorem is always true, even if there are singularities at the origin, provided they are consistent with (4.2) and (4.5). 23 ∞ Note that instead of (4.8) one could formally adopt the normalization 0 dx δ(x)=1. In 1 this case one would avoid the factors “ 2 ” in (4.13) and “2” in (4.14). However,R this is inconsistent if the same symbol “δ” is used for both δ(r) and Dirac’s δ(x). In order to remind that we keep Dirac’s normalization(4.8) and use (4.14) in polar coordinates, we will always write mnemonically “2δ(r)” for the derivative of Υ(r). 24Using a more heuristic method, the same conclusion was reached in [18, 19].

11 5 Monopole singularity

Starting from the Coulomb potential of a point-charge, 1 ϕ (~r) := e Υ(r) , (5.1) m r which may be either an electric-monopole of charge e, or a magnetic-monopole if e is replaced by ie, we get the field ~r ~r E~ (~r)= −∇~ ϕ = e Υ(r) − e 2δ(r) (5.2) m m r3 r2 and the rationalized source charge distribution 2 −4πρ (~r)= −∇·~ E~ = e δ(r) . (5.3) m m r2 There are two differences between these distributions and the usual “text book” expressions for the Coulomb field and its charge distribution. First, there is the δ-function term on the right of (5.2). This is a solution of the homogeneous equation ∇~ Em(~r)=0 which therefore can always be added to any solution of the inhomogeneous equation. Second, there is the factor of 2 on the right of (5.3), which when integrating over r is taken care of by the normalization (4.13). In particular, the charge of the source is

q = d3Ω ρ (~r)= e . (5.4) ZZZ m

6 Yukawa singularity

It is possible to generalize the monopole distribution by multiplying it by a scalar function f(r). However, from a physical point of view, the most interesting generalization of this type is the Yukawa distribution of low-energy pion-exchange strong-interaction. Its potential, field, and “current” are25 1 ϕ (~r) := g e−mr Υ(r) , (6.1) y r ~r ~r E~ (~r)= g(1 + mr) e−mr Υ(r) − g e−mr 2δ(r) , (6.2) y r3 r2 1 1 2 −mr −mr (6.3) −4πρy(~r)= gm e Υ(r) + g(1+2mr) 2 e 2δ(r) . r  r 25When g is imaginary the Yukawa interaction is “pseudo-scalar,” i.e., changingsign in a 3-space inversion, and “scalar,” i.e., invariant, when g is real.

12 The main difference with the monopolesingularity is the presenceofthe“mass 2 term” on the right of (6.3), i.e., the expression m ϕy(~r) which is proportional to m2 and characteristic of a massive particle. The term containing the δ-function is then interpreted as the “source,” which has the charge

q = d3Ω ρ (~r) = g . (6.4) ZZZ y source

7 Dipole singularity

The dipole singularity, which through extensive experimental verification is found to very precisely characterize the intrinsic magnetic dipole moment of elementary particles such as the electron, is given by the vector potential ~µ × ~r A~ (~r) := Υ(r) (7.1) d r3 where |~µ | has the dimension of a charge times a length. The calculation of the magnetic field is straightforward. We get ~r ~µ ~r × (~µ × ~r) H~ d(~r)= 3 (~µ · ~r) − Υ(r)+ 2δ(r) . (7.2)  r5 r3  r4 The first term in this expression is well-known, but the one with a δ-function is rarely mentioned in text books. However, when integrated over 3-space, this second term gives the contribution [24, p.184] ~r × (~µ × ~r) 8π d3Ω2δ(r) = ~µ (7.3) ZZZ r4 3 which is essential in calculating the hyperfine splitting of atomic states [25]. We can now calculate the sources. As expected, the magnetic charge density is zero −4πρd(~r)= ∇·~ H~ d(~r)=0 , (7.4) while the rationalized current density is ~r × ~µ −4π~j (~r)= ∇~ × H~ (~r)= −3 2δ(r) . (7.5) d d r4 Using this current density we can now calculate the magnetic moment by means of the standard expression [24, p.181] to get 1 ~m = d3Ω ~r × ~j (~r)= ~µ. (7.6) 2 ZZZ d Therefore, although there are actually no “circulating currents” in the point-like distribution (7.5), the magnetic moment calculated with the formula derived for a localized current distribution gives the correct answer.

13 8 Dimonopole singularity

The dimonopole singularity corresponds to the field produced by two electric (or magnetic) monopoles of opposite charge separated by an infinitesimal distance |~λ|. The potential for such a field is therefore e e ϕdm(~r) := Υ(r) − Υ(r) . (8.1) |~r + ~λ| |~r |

At large distance, or at vanishinglysmallseparation ~λ, wecantakeforthispotential the first term of the Taylor development, i.e., 1 1 ϕ (~r) ≈ +e(~λ · ∇~ ) = −e(~λ · ~r) . (8.2) dm r r3 Fromthereonitispossibletocalculatethefieldandthesource by either recursively applying the gradient operator on (8.2), or by applying the operator (~λ · ∇~ )() on the field (5.2) and the source (5.3) of a point charge. Either way, we get for the field the expression

~r ~µ ~r(~µ · ~r) H~ dm(~r)= 3 (~µ · ~r) − Υ(r) − 2δ(r) , (8.3)  r5 r3  r4 where we have defined ~µ = e~λ . (8.4) Expression (8.3) is remarkably similar to the corresponding expression (7.2) for an intrinsic dipole, and it can be seen that the difference between a dipole and a dimonopole field is entirely contained in the point-like singularity at the origin, i.e.,

~r × (~µ × ~r) ~r · (~µ · ~r) ~µ H~ (~r) − H~ (~r)= 2δ(r)+ 2δ(r)= 2δ(r) . (8.5) d dm r4 r4 r2 As a result, when integrated over 3-space, the dimonopolar δ-singular term gives the contribution [24, p.141]

~r(~µ · ~r) 4π − d3Ω 2δ(r)= − ~µ (8.6) ZZZ r4 3 which differs in sign and in magnitude from the corresponding expression (7.3) for an intrinsic dipole. It is this difference which enables to conclude that the dipolar fields from distant stars are produced by magnetic dipoles, rather than by magnetic dimonopoles [25].

14 We can now calculate the sources. As expected, the current density is zero

−4π~jdm(~r)= ∇~ × H~ dm(~r)=0 , (8.7) while the rationalized charge density is

~r · ~µ −4πρ (~r)= ∇·~ H~ (~r)= −3 2δ(r) , (8.8) dm dm r4 i.e., a distribution that is odd in ~r so that the total charge is zero, as it should be for a dimonopole. We can finally calculate the first moment of this charge density by means of the standard expression for a charge distribution [24, p.137]. This gives

d~ = d3Ω ~rρ (~r)= ~µ = q~λ , (8.9) ZZZ dm a result which illustrates again that despite the great similarity of their fields at a distance from the origin, the dipole and dimonopole singularities are in fact very different.

9 Vectorpole singularity

Inordertocompleteoursystematicstudyofthemostsimplepoles we nowconsider the “vectorpole.” In order that it has the dimensions of a potential we define this singularity as ~λ ~µ A~ (~r) := e Υ(r) ≡ Υ(r) . (9.1) v r2 r2 Since this is a vector potential, we get the magnetic field

~r × ~µ ~r × ~µ H~ (~r)= ∇~ × A~ (~r)= −e 2Υ(r)+ e 2δ(r) . (9.2) v v r4 r3

Because H~ v is a rotational, there are no magnetic charges

−4πρv(~r)= ∇·~ H~ v(~r)=0 , (9.3) while the rationalized current distribution is ~µ ~r(~µ · ~r) ~µ ~r(~µ · ~r) −4π~jv(~r)= − 4 − 8 Υ(r)+ 4 − 6 2δ(r) . (9.4)  r4 r6   r3 r5  In contrast to the previously considered poles, this source distribution is relatively complicated and difficult to interpret. Compared to (6.3) the Υ-term does not even

15 have the correct sign to be a “mass term,” and the δ-term is not simply the potential multiplied by a constant times r−1 as it normally is for a source. Moreover, since (9.4) is even in ~r, the magnetic moment (7.6) zero, as is the charge, so that

qv = ~mv = d~v =0 . (9.5)

Nevertheless, considering the relative simplicity of the potential (9.1) and field (9.2), one should not discard this pole in the perspective of a still to be discovered fundamental theory of elementary particles. In particular, if we define the extent ~̺ of a current distribution (whose norm ̺ = |~̺| is a measure of its “radius”) by

~̺ = d3Ω r~j(~r) (9.6) ZZZ we get for the source term in (9.4) a finite result

~̺ =2~µ, (9.7) in contrast to previously encountered sources which always had a zero extent. The vectorpole singularity may therefore in some way be related to an extended structure such as the “bag” which characterizes hadronic particles like the nucleons.

10 Systematics of elementary poles

We have now studied a number of elementary singularities of gradual complexity. Their main properties are summarized in Table 1 where, for clarity, we have simplified the expression of the sources by removing the numerical factors in front of them. For each singularity we give their charge q, magnetic moment ~m, electric moment d~, and extent ~̺, calculated from the full expression of their respective sources according to the definitions (5.4), (7.6), (8.9), and (9.6). Obviously, the elementary poles in Table 1 do not constitute an exhaustive list. In particular, it is possible to multiply them by various scalar functions, and to obtain more complicated vector structures by using more parameters than just one vector ~µ. However, this list seems fairly complete considering that it contains all elementary poles which correspond to the following criteria:

1. The potential should have the dimensions [charge]×[length]−1. 2. The potential should be regular everywhere except at the origin (r =0). 3. The poles should by simple combinations of r−n, ~r, and a constant vector ~µ.

16 4. Each pole may be multiplied by a factor e−mr to yield a mass term. 5. The Laplacian of the potential should contain a δ-like source term. 6. Thepolesshouldhave(apartfromapossiblemass)atleast one finite attribute such as charge q, magnetic moment ~m, electric moment d~, or extent ~̺, and none of these attributes should be infinite.

Moreover, we have included in the table two exceptional poles which do not satisfy the last two criteria. First, the “photon26 potential” which is simply the point-like solution of Maxwell’s equations in a charge vacuum (i.e., a source-free region where there are no charges)

ϕp(~r) := eδ(r) , (10.1) and the “vacuum potential,” or “abon27 potential,” ~r A~ (~r) := e Υ(r) (10.2) a r2 which, because ∇×~ A~a =0, does not have a field and therefore has only vanishing attributes. The abon is included in the table for the sake of completeness with regards to the third criterion, and because it is the simplest possible singularity in a total vacuum, i.e., a region where there are no charges and no photons, thus no electromagnetic field at all. The second criteria is excluding potentials with a line-like singularity such as the Dirac or the Schwinger magnetic monopoles, i.e., g ~r × ~u g (~r · ~u)(~r × ~u) A~ = , A~ = , (10.3) D r r − ~r · ~u S r r2 − (~r · ~u)2 where ~u is a unit vector [27].28 As other examples of potentials excluded by our criteria, we have those con- taining high-order singularities in r−n and ~r/rn+1 with n > 2 because they lead to infinite attributes. In conclusion, while the approach taken in defining the above criteria is heuris- tic, it nevertheless leads to all major singularities which play an essential role in 26By photon we do not refer here to a single quantum-mechanical photon, but to a superposition of such fields with a well-defined position, which is a classical object. 27We suggest this name because this potential is a point-like singularity which may give rise to a quantum mechanical effect similar to the Aharonov-Bohm effect. 28Note that these potentials are obtained by multiplying the abon potential (10.2) by a dimen- sionless vector function of the type ~uf(r, ~r · ~u).

17 elementary particle physics. In particular we see that contrary to classical elec- trodynamics (where the potential is less fundamental than the fields) Lanczos’s electrodynamics emphasizes (just like quantum theory) the 4-potential A and the source 4-current C, which are related by the second order equation

∇(∇ ∧ A)= −C . (10.4)

From Table 1, which lists singularities at rest, we see that this differential equation yields (for all poles except the vectorpole) the following algebraic relation for the currents 1 ∇~ 2A = C = κ δ(r)A , κ ∈ Z . (10.5) r

11 Electrons in classical and quantum physics

We know from experiment that all electromagnetic properties of electrons derive from the assumption that they are point-like particles characterized by a scalar, their electric charge e, and a vector, their intrinsic dipole moment ~µ.29 Therefore, by merging these two quantities into a single quaternion, the source singularity which yields all known properties of electrons becomes a single whole entity, i.e., in the electrons’s rest frame, the 4-current density biquaternion

e ~r × ~µ B (11.1) Ce = 2 − 3i 4 2δ(r) ∈ , r r  which, in a moving frame defined by a Lorentz boost B,30 generalizes to

′ Ce = BCeB . (11.2)

However, while these expressions are wholy satisfactory in classical electrody- namics where the magnitudeof the magnetic moment |~µ| is taken from experiment, it would be desirable to have a “first principles” derivation starting from the Dirac equation showing that their form is also correct in quantum electrodynamics, and that e~ µ = |~µ| = . (11.3) 2mc Unfortunately, this is not possible in a simple way, mainly because the standard interpretation of quantum mechanics is associating electrons with plane-wave 29By “all” we mean all classical and first order quantum mechanical properties. 30In the biquaternion formulation of special relativity the Lorentz boost factor B is simply the “biquaternion square root” of the 4-velocity U, i.e., U = BB. For more details, see [14].

18 solutions of Dirac’s equation — which are of infinite extent — so that the singular 4-current (11.1) emerges only as the result of a complicated calculation which involves the whole apparatus of quantum field theory. Nevertheless, it is possible to give a heuristic derivation of (11.1) using Gordon’s decomposition (3.4), which will also give some further insight that is useful for this paper. Let us therefore consider the general plane-wave solution of Dirac-Lanczos’s equation (3.1) for a free electron, Aext =0, i~ν D = aL exp( U · X ) (11.4) ℓ where a ∈ C is a normalization constant.31The four covariant quantities appearing in the Gordon decomposition are then

C = eDD+ = e|a|2U , s = DD = |a|2 , (11.5)

i~ν U·X − i~ν U·X ε()= e[ ]hDi~νDi = e[ ]|a|2hB e+ ℓ i~ν e ℓ Bi , (11.6)

M()= eDi~νD[ ] = e|a|2Bi~νB[ ] . (11.7) Since the only variable in the plane-wave solution (11.4) is the 4-position X , the only non-zero contribution comes from the electric convection current (11.6), and the Gordon decomposition reduced to the trivial equality e|a|2U = e|a|2U. In particular, the electromagnetic moment term (11.7) does not contribute since for a plane-wave it is a constant independent of X . Moreover, because of the Gordon identity, the self-field A = Aself is forced to be zero. In the field-theoretical interpretation of the Dirac equation this null result is avoided by considering that electrons are plane-wave solutions restricted to finite regions of spacetime. Heuristically, this interpretation can be implemented by replacing the normalization factor |a|2 by an inverse volume element which upon integration over 3-space yields a finite constant. In spherical symmetry, this is achieved by means of the substitution 1 |a|2 → 2δ(r) . (11.8) r2 Then, by keeping this factor within the range of the operators as suggested by the position of |a|2 in the expressions (11.6) and (11.7), both the electric convec- tion and magnetic moment terms yield non-trivial contributions, and the Gordon 31All calculations in this section are elementary and based on the differentiation rule ∇(U ·X )= U, and the identity U = BB which leads to expressions such as B∗ = UB because UU =1, etc.

19 decomposition becomes

1 1 e 1 e 3i~ν~r C = αA 2δ(r) − U 2δ(r) − B B2δ(r) . (11.9) ℓ self r2 ℓ r2 2 r4 Multiplying both sides by ℓ and grouping the electric convection and magnetic moment terms we get

1 e eℓ ~ν~r (11.10) C = ℓαAself 2 2δ(r) − B 2 +3i 4 B2δ(r) . r r 2 r  where, upon developing the quaternion product ~ν~r = −~ν ·~r −~r ×~ν, we recognize the electron current density (11.1), as well as the correct magnitude of the magnetic ~ eℓ moment because ℓ = mc implies that 2 = µ as given by (11.3). However, at the same time as we got the sought-after electric-monopole magnetic-dipole current density (11.1), we got an extra term, arising from the scalar part of the ~ν~r quaternion product, yielding a magnetic dimonopole contri- bution ~r · ~µ C = iρ = −3i 2δ(r) ∈ C . (11.11) dm m r4 Had we restricted the Gordon decomposition to its standard “Hermitian” part this term would automatically have been discarded, and we would have remained within the usual, “purely electromagnetic,” interpretation of the Dirac equation. By keeping it, we have found a term that can be interpreted as a source current for a strongly interacting field, such as the one which successfully describes the low energy interactions between pions and nucleons, see, e.g., [28, p.3-10] or [29, p.434-439]. In conclusion, by using the Gordon decomposition in the general form given by Gursey,¨ (3.4), we have been able to show in a heuristic32 way that both the magnetic dipole and the magnetic dimonopole source distributions arise on equal 1 footing within the context of Dirac’s equation for spin 2 particles. However, while the dipole distribution corresponds to the usual electromagnetic interpretation of this equation, i.e., to “electrons in interaction with photons,” the dimonopole distribution (11.11) corresponds to a hadronic interpretation, i.e., to “protons in interaction with pions.” 32The heuristic character of our derivation is illustrated by the form of the self-potential A = 3 e ~ν~r Aself = −i 2 α B r2 B which is needed to balance the Gordon decomposition (11.9). Since this potential cannot be eliminated by a gauge transformation, it requires a non-trivial interpretation which is beyond the scope of this paper. In particular, it features the notorious muon to electron 3 1 mass ratio, i.e., 2 α ≈ 205, so that its interpretation may have to do with the existence of muons.

20 12 Physical particles in Lanczos electrodynamics

The first singularity which comes to mind when considering classical electrody- namics is the electric monopole, i.e., in the non relativistic limit, the Coulomb potential33 e ϕ = , (12.1) C r and its relativistic generalization, the Lienard-Wiechert´ potential e A = U (12.2) LW ξ where the radial distance r is replaced by the retarded distance ξ and the 4-velocity U corresponds to the world-line of the singularity. However, neither of these potentials correspond to a physical particle since, as known from experiment and shown in the previous section, a physical electron is endowed with a magnetic dipole moment, so that its source current is given by (11.1) which corresponds to the pole-dipole potential associated with a Dirac particle, i.e., e ~r × ~µ A = ϕ − iA~ = − i . (12.3) D monopole dipole r r3

In fact, the potentials (12.1) and (12.2) also have other problems: they lead to infinite self-energy, and to an infinite Hamiltonian function. It is to find a cure to this second problem that in his doctoral dissertation of 1919 Lanczos proposed to shift the singularity at the origin of (12.1) by a small imaginary translation into complex spacetime [7, p.55]. In this case the singularity is no more point-like but has the form of a circle, hence the name given by Lanczos to this shifted singularity: the “circle electron.” Mathematically, if for simplicitywe confine our discussion to the non-relativis- tic case — equation (12.1) — and to infinitesimal translations which should be enough to get the essential features of the physical differences between the point electron and the circle electron, we can represent the shifting of the origin by the imaginary vector i~λ by the scalar operator34

Ωsca()=1[] − (i~λ · ∇~ )[ ] . (12.4) Then, applying this operator to the Coulomb potential we get 1 1 Ω : ϕ → ϕ = e − e(i~λ · ~r) . (12.5) sca C r r3 33In this section, to simplify the notation, we keep implicit the Υ(r) factors multiplying the distributions. 34The minus sign is used in order to fit the quaternion formulation of 4-vectors such as (12.3).

21 i.e., the monopole-dimonopole potential. Similarly, since this operator commutes with ∇~ , it can also be applied to the Coulomb field and its source current, so that one gets all the corresponding distributions in a consistent way.

In other words, by means of Ωsca(), we have added to the Coulomb potential (12.1) an imaginary dimonopole potential of the form (8.2). Since in Lanczos- Newman electrodynamics ∇~ ϕ = −E~ − iH~ , the effect of such an imaginary scalar potential is to produce a magnetic field which — away from the origin — has a dipolar form, Newman was led to conclude that the effect of the imaginary translation (12.4) is to induce a magnetic dipole moment [2, 4]. However, as seen in sections 7 and 8, where the differences between the dipole and the dimonopole singularities were discussed, this is not the case. In particular, the magnetic dipole field derives from a vector potential, while the only effect of the operator (12.4) is to make the scalar potential (12.1) complex instead of real. Moreover, in the standard interpretation of Maxwell’s field, imaginary contributions to the scalar potential correspond to “magnetic poles” which are normally excluded by Maxwell’s inhomogeneous equations. What is then the physical interpretation of the monopole-dimonopole singu- larity (12.5) ? Could there be a relation between this complex scalar potential and the physics of low-energy pion-nucleon interactions which are precisely described by a complex scalar field ? Let us therefore consider the theory of pseudo-scalar pion-nucleon interactions, in which nucleons (i.e., protons and neutrons) are de- scribed by Dirac’s equation, and neutral pions by the scalar field ϕ solution of the second order equation, [28, p.9] or [29, p.436],

−2 ~ ∇∇ + ℓπ ϕ = gℓπ(i~ν · ∇)ρ(r) . (12.6)  Here ρ(r) is the nuclear matter distribution source of the pion field, ~ν the nu- ~c cleon spin axis as in Dirac-Lanczos’s equation (3.1), 2 the Compton ℓπ = mπc wave-length associated with the pion mass, and g the strong interaction charge. Assuming that the nuclear distribution is reduced to a point-like nucleon, i.e., 1 ρ(r)= 4πr2 2δ(r), this equation admits a static solution g 1 g i~ν · ~r ϕ(r)= ℓ (i~ν · ∇~ ) e−r/ℓπ = − (ℓ + r) e−r/ℓπ (12.7) 4π π r 4π π r3 which in the limit of a mass-less pion, mπ → 0, reduces to g i~ν · ~r ϕ π (r)= − ℓ , (12.8) (m =0) 4π π r3 i.e., to the dimonopole term on the right hand side of (12.5). Therefore, the effect ~ g of translating the Coulomb singularity into complex space by eiλ = 4π ℓπi~ν is

22 equivalent to inducing a pseudoscalar interaction potential, formally similar to the oneleadingtothesourcecurrent (11.11)thatwefoundinthe previous section when discussing the generalized Gordon decomposition of the conserved current of a Dirac particle, but which we can here interpret as corresponding to the well-known pion-nucleon interaction at low-energy.35 Having shown that the imaginary scalar operator (12.5) is generating a strong interaction potential, it is natural to ask whether the magnetic interaction potential which corresponds to the vector part of the electron potential (12.3) could also be derived from the Coulomb potential by an operator of a similar type. The answer is immediate, and it is easy to verify that (12.3) can be derived from (12.1) by means of the imaginary quaternion operator

Ωqua()=1[] − (i~λ × ∇~ )[ ] . (12.9) i.e., 1 1 Ω : ϕ → e − e(i~λ × ~r) = A , (12.10) qua C r r3 D provided e~λ = ~µ. In summary, we have found that starting from the Coulomb potential — which corresponds to a hypothetical charged particle with only “electric” interactions — we can obtain both the magnetic interaction potential needed to have a physical electron, and the pseudo-scalar pion-nucleon interaction potential characteristic of low energy strong interactions, simply by applying the operators Ωsca and Ωqua on the world-line of a Coulomb singularity moving in biquaternionic spacetime. For instance, combining our previous results, it is trivial to write down the 4-potential which in Lanczos’s electrodynamics corresponds to a proton, i.e.,36 1 i~ν · ~r i~ν × ~r A = e − g(ℓ + r) e−r/ℓπ −µ , (12.11) P r π r3 P r3 and from (6.1) the associated potential of the neutral pion field 1 ϕ (~r)= gi e−r/ℓπ . (12.12) π r These results were obtained in the non-relativistic limit, but there should be no reasonwhytheycouldnotbegeneralizedtotherelativistic case and arbitrary world- lines. The recent work of Ezra Newman [1], which was motivatedbythepossibility 35Of course, the action of the operator (12.4) on the Coulomb field is to lead to the pseudo-scalar field (12.5) which corresponds to a massless pion. However, the generalization to a finite mass pion field is straightforward. 36This simplified expression containing only neutral pions could easily be generalizedto include charged pions, as well as other strongly interacting particles such as vector mesons, whose effect quickly become non-negligible as energy is increased [5, 6].

23 that the translation Ωsca could yield essential properties of elementary particles [2, 4], is therefore of great interest. In particular, as is stressed by Newman, it is important that the displacement Ωsca is not an ordinary Poincare´ translation (under which all physical laws are invariant) but an imaginary translation which by complexifying a scalar potential is leading to “new” physics, what is indeed the case since starting from a pure Maxwell field we get strong interactions. Similarly, 37 the operation Ωqua is not an ordinary rotation or boost (which would leave the physics invariant) but a biquaternion operation which is transforming the real scalar Coulomb potential into the biquaternion potential (12.3). In other words, it is by these “non-orthodox” operations that we are possibly gaining some new insight into the nature of spacetime and elementary particles, something that was anticipated by Cornelius Lanczos in 1919 already, as well as in another context by Albert Proca in a largely forgotten line of research that he initiated in 1947, see [30].

13 Field-theoretical interpretation

The main results of this paper are the operations Ωsca and Ωqua, defined by equa- tions (12.5) and (12.9), which lead from the potential, field, and source of a bare monopole singularity (which cannot be associated with any known physical par- ticle) to the corresponding distributions which characterize the physically well established pion-nucleon and photon-electron interactions. While the scalar oper- ator Ωsca may correspond to an infinitesimal imaginary translation which has the effect of adding an imaginary part to a real monopole singularity, the interpretation of the biquaternion operator Ωqua, which is adding an imaginary vector component to this monopole singularity, does not seem to have such a simple geometric inter- pretation. In fact, as will now be shown, it may be that the interpretation of Ωsca as an imaginary translation could be a mere coincidence, and that in fact the two operations Ωsca and Ωqua have a completely different interpretation. Suppose that we start from a solution B(X ) of Maxwell’s equations (2.3) such that for X 6=0 we have ∇B =0 . (13.1) If we exclude fields such that ∇B =0 for X =0, this solution will therefore have a δ-like source, as is the case, for example, for the monopole field. Then, for any constant quaternion Q which transforms as a 4-vector in a Lorentz transformation

37 Since there is some resemblance between Ωqua and a rotation by an imaginary angle iθ about an axis ~ν, we recall the infinitesimal form of such an operation acting on the coordinates, which is the scalar operator Ωrot()=1[] −~r · (i~λ × ∇~ )[ ] where ~λ is now the dimensionless vector ~λ = θ~ν.

24 we can write all sort of expressions such as, for example,

Q∇∇B =0 , ∇Q∇B =0 , (13.2) which are trivial covariant identities from which we can derive new solutions of Maxwell’s equations in the generalized form (2.3). In particular, by adding or subtracting the previous identities we get

(Q · ∇)∇B =0 , (∇ ∧ Q)∇B =0 , (13.3) which since Q is a constant can be rewritten as

∇(Q · ∇)B =0 , ∇(Q ∧ ∇)B =0 , (13.4) so that starting from the solution (13.1) we have derived the new solutions

Bsca =(Q · ∇)B , Bqua =(Q ∧ ∇)B . (13.5)

In the special case where Q reduces to the vector ~µ in the rest frame, and if we restrict ourselves to the stationary case, we recognize the non-trivial parts of the operators Ωsca and Ωqua. Thus, we have just proved that these operators lead from a given solution of Maxwell’s equations to new solutions, which in the case where B is the Coulomb/Lienard-Wiechert´ field are simply the pion-nucleon and photon-electron interaction fields.38 Moreover, since a suitably normalized Dirac field D can be interpreted as defining a special Lorentz transformation L = BR, the quaternion Q can formally be related to Dirac’s pseudo-vector Σ and scalar s given by (3.2) and (3.3). This enables to write the covariant expressions

gℓ Ω ()=1[]+ (Σ · ∇)[ ] , (13.6) sca 2s eℓ Ω ()=1[]+ (Σ ∧ ∇)[ ] , (13.7) qua 2s where we have normalized Ωqua so that it yields the Dirac-electron potential (1.1) eℓ with ~µ = 2 ~ν as in Section 11, and left the normalization of Ωsca open by leaving the coupling constant g undefined. Therefore, the covariant operators (13.6) and (13.7) can be interpreted in two ways: first, as operators where Σ is related to an arbitary special Lorentz transformation D = BR with B any boost and R any rotation (in which case s = 1); second, as operators where Σ and s are related 38As in Section 11 where we dealt with the Gordon decomposition of the Dirac current, we see that a fully covariant treatment necessarily leads to both of these interactions on an equal footing.

25 to a solution of Dirac’s equation interpreted as defining an arbitrary world-line in spacetime [20, 31]. Obviously, the methodology developed in this section can easily be generalized to yield potentials, fields, and sources which in Lanczos’s electrodynamics may correspond to other type of particles and their interactions. Theproblem, ofcourse, is to properly interpret these distributions, and to investigate whether or not they correspond to physical particles.39

14 Conclusion

To conclude let us return to the references [1, 2, 3, 4] and make a few comments:

• In his paper of 1973 Ezra Newman rediscovered Lanczos’s “circle electron” and the possibility that the magnetic field which derives from it in a com- plexified Maxwell theory could correspond to the intrinsic magnetic dipole moment of an electron [2]. In fact, as we have shown in this paper, this field does not correspond to a magnetic dipole, but rather to a magnetic dimonopole, as should be the case since it derives from a scalar rather than from a vector potential.

• In 2001 Gerald Kaiser reviewed the possibility that the magnetic field of the Lanczos-Newman “circle electron” could be that of an intrinsic magnetic dipole. Using acomplexified spacetimemethodology based on the definition of a complexdistance, he apparently succeeded in showingthatthisisindeed the case. For instance, he was able to show that the source-current associated with the “circle electron” is a 4-vector having a non-zero vector part in the rest-frame, while the complex potential from which he started from was a scalar. This is clearly impossible since the two should be connected by the Laplace operator ∆ = −∇~ ∇~ , which is a scalar. Nevertheless, the method used by Kaiser is very interesting because it confirms that the intrinsic difference between the dipolar and dimonopolar fields is to be found in the δ-like singularity at the origin, when the dimonopole field is studied in the point-like limit as in the present paper, or on the “singular circle” in the more sophisticated approach of Kaiser which keeps the circular structure 39Note that while the real Coulomb field of classical electrodynamics contains just the electric charge e (which combined with the mass yields the Thomson charge radius e2/mc2 as length scale), the complexified fields obtained by means of the operators Ωsca and Ωqua contain the Planck constant ~ (which combined with the mass yields the Compton wave-length ~c/mc2 as length scale), so that these operators provide a link between classical and quantum theory.

26 of the field near the origin. Then, by surrounding the singularity with a closed surface (i.e., respectively, a sphere or a torus) and studying the field external to this surface while letting its dimensions tend to zero, one is effectively removing a part of the singularity and reducing it to that of a true dipole. In particular, while the method of Kaiser is perfectly correct in ordinary electrodynamics (where neither “magnetic monopoles” nor “circle electrons” exist), Kaiser’s result may be due to the fact that his method has the effect of “filtering out” that part of the dimonopole singularity which corresponds to an intrinsic magnetic dipole. Therefore, Kaiser’s result has an important consequence: it could be that what he did may apply to nucleons and other elementary particles. In effect, while the measured magnetic dipole moments of baryons appear to exactly conform to those of intrinsic magnetic dipoles, they cannot be fully consistently derived from the magnetic moments of their constituents, i.e., the quarks, which in some respect behave more like magnetic dimonopoles confined within a “bag,” than as regular Dirac dipoles.

• In his paper of 2002 Ezra Newman revived another consequence of em- bedding the real Maxwell field and the real linearized Einstein equations in complex Minkowski space: the possibility of interpreting magnetic moment and spin angular momentum as arising from a charge and mass monopole source moving along a complex world-line in the complex Minkowski space [4]. Thisleadhimtoshowthatinthe propercircumstancesthe gyromagnetic ratio of the singularity would by g =2, i.e., that of the Dirac electron. Since according to the present work the world-line obtained by simply translating it by an imaginary amount into complex space is that of a dimonopole, the conclusion of Newman should be rephrased accordingly. Then, since there is a growing consensus that, in fact, there is nothing special about the Dirac electron in this respect [32, p.79], and that g = 2 is the correct gyromag- netic ratio for all elementary particles [33, 34], Newman’s result for the dimonopole singularity is perfectly consistent with current knowledge.

• Finally, in his paper of 2004 Ezra Newman considerably increased the value of his earlier considerations about world-lines displaced into complex space by showinghow such world-lines, and their association to the relativistic and causal generalization of the Coulomb field — the Lienard-Wiechert´ fields — naturally ariseas solutions ofMaxwell’s equations in complex spacetime [1]. It would therefore be of great interest to make a direct connection between these findings and Maxwell’s equations written in Lanczos’s formalism. In particular, using the biquaternion spinor techniques developed by Paul Weiss [35], it could be that much of the complicated calculations leading

27 to these results could be simplified. Moreover, such a transcription could help finding how to substitute the imaginary quaternion operation (12.9), which leads from a bare electric-monopole to the intrinsic magnetic-dipole moment of a Dirac electron, for the imaginary scalar operation (12.5), which leads instead to a magnetic-dimonopole moment. If that can be done, it is likely that Newman’s conclusions, which apply to the Lienard-Wiechert´ fields of an electric-monopole magnetic-dimonopole world-line, could be generalized to the world-line of a genuine Dirac electron, and possibly to those of other particles, therefore opening a new approach to the problem of elementary particles in special and general relativity.

In conclusion, the possibility that Lanczos’s electrodynamics and Lanczos’s formulation of Dirac’s theory could provide a connection between elementary particle physics and general relativity theory appears as very promising.40 In particular, it is striking that in the context of a simple extension a classical elec- trodynamics into complex space (which leads to the Lienard´ Wiechert fields in a direct way) a rather trivial imaginary translation operation could lead to the correct form of low-energy pseudo-scalar pion-nucleon strong interaction. If confirmed, this could mean that the well-known formal similarities between magnetodynam- ics and low-energy strong interactions, see [28, p.3-10] and [29, p.434-439], are not just accidental but of a fundamental nature [15]. Much work remains to be done, but considering the persistent difficulties with the physical interpretation of the “Standard Model,” it is certainly worth doing it.

15 Acknowledgments

This paper is dedicated to Prof. J.D. Jackson whose enlightening work on the nature of intrinsic magnetic dipole moments has greatly influenced the results presented here. I am also very much indebted to Dr. J.-P. Hurni for many stimulating discussions and considerable assistance in many aspects of this paper. 40In this perspectiveit is importantto stress that there are a number of approaches closely related to Lanczos’s and Newman’s which lead to similar conclusions. See, in particular, [36, 37, 38].

28 16 Appendix: Gauss theorem in polar coordinates

In this appendix we give an elementary, and possibly incomplete, proof of our conjecture that the Gauss theorem (4.9), which we rewrite here as

d2Σ F (~r)= d3Ω ∇~ F (~r)Υ(r) , (16.1) ZZ ZZZ ∂Ω Ω is true for any weakly convergent quaternion-valued distribution F (~r)Υ(r) which may have a singularity at r = 0. Developing the right hand side, and using weak convergence (4.5), we get ~r d2Σ F (~r)= d3Ω ∇~ F (~r)+ d3Ω F (~r)2δ(r) , (16.2) ZZ ZZZ ZZZ r ∂Ω Ω Ω which is the standard form of the Gauss theorem for the function F (~r), except for the extra term on the right which takes into account the singularity at the origin. As F (~r) is quaternion valued, equation (16.2) contains all forms of the Gauss theorem, namely the so-called “gradient-,” “divergence-,” and “rotational- theorems” at once [26, vol.II, p.439]. For the sake of completeness, we recall the definition of the quaternion product

(a + A~)(b + B~ )= ab − A~ · B~ + aB~ + bA~ + A~ × B~ . (16.3)

Since F (~r) is supposed to be finite and differentiable everywhere except at r = 0, the theorem is true by any standard proof for any volume which does not contain the origin. We then proceed as in any standard analysis of point-like singularities, namely we surround the origin by a small sphere and verify the theorem for the leading singularity. However, for simplicity, we postpone to a more complete proof the cases where the origin in on the surface of the volume Ω, and the one where Ω is multiply connected. From now one the volume Ω is therefore an ordinary 3-ball of radius R, ∂Ω its surface, and the function F (~r) the leading singularity. If F (~r) is simply a scalar function f(r), which does not depend on the angular variables, the theorem reduces after angular integration to the identity 0=0. This remains so after multiplying both sides of (16.2) by a constant vector, so that to deal with non trivial singularities we must consider expressions containing the vector ~r. The most simple such expression is ~r F (~r)= (16.4) rn 29 which is singular at r =0 for n ≥ 1. For the 3-ball we have ~r d2Σ= dωr2 and d3Ω= dωr2dr . (16.5) r Thus, after some elementary calculation, the three terms appearing in (16.2) be- come ~r ~r d2Σ F (~r)=4πr2 = −4πR3−n (16.6) ZZ r rn ∂Ω R R d3Ω ∇~ F (~r)=4π dr (3 − n)r(2−n) = −4π r3−n (16.7) → ZZZ Z0 r 0 Ω

~r R ~r ~r r→0 3 2 3−n (16.8) d Ω F (~r)2δ(r)=8π dr r n δ(r)= −4π r ZZZ r Z0 r r Ω where the minus signs comes from the definition (16.3), i.e., ~r~r = −r2. Therefore, by comparing (16.7) and (16.8), we see that the second term on the right hand side of (16.2) has the remarkable effect of removing the divergent term at r → 0 which should not be there according to (16.6), and, moreover, inthecase n =3, of correcting equation (16.7) so that instead of zero it gives the same value as (16.6). In other words, this second term “repairs” the usual formulation of the Gauss theorem in such a way that it becomes true for all functions of the type (16.4). We have therefore proved the identity ~r ~r d2Σ = d3Ω ∇~ Υ(r) (16.9) ZZ rn ZZZ rn ∂Ω Ω which is stillvalidif we multiplyit by a constantvector ~c. Since ~r~c = −~r·~c+~r×~c, thismeansthatthe Gausstheorem isalsotrue for thefunctions ~r·~c/rn and ~r×~c/rn, which correspond to generic cases of the “divergence-” and “rotational-” theorems. Moreover, since these scalar and vector products enable to project the radius vector on any axis or plane of the polar coordinate system, they can be combined with suitable scalar functions to yield any possible function of r, θ, and φ, which may be singular for r = 0, so that the Gauss theorem in the forms (16.1) or (16.2) should be true for any such function.

References

[1] E.T. Newman, Maxwell Fields and Shear-Free Null Geodesic Con- gruences, (12 February 2004) 39pp. arXiv:gr-qc/0402056 available at http://arXiv.org/abs/gr-qc/0402056.

30 [2] E.T.Newman, Maxwell’s equations and complex Minkowski space, J.Math. Phys. 14 (1973) 102–103.

[3] G. Kaiser, Distributional sources for Newman’s holomorphic field, (22 August 2001) 15pp. arXiv:gr-qc/0108041 available at http://arXiv.org/abs/gr-qc/0108041.

[4] E.T. Newman, Classical, geometric origin of magnetic mo- ments, spin-angular momentum, and the Dirac gyromagnetic ratio, Phys. Rev. D65 2002) 104005-1/8. arXiv:gr-qc/0201055 available at http://arXiv.org/abs/gr-qc/0201055.

[5] B.D. Serot and J.D. Walecka, Recent progress in quantum hadrodynamics, Int. J. Mod. Phys. E6 (1997) 515–631. arXiv:nucl-th/9701058 available at http://arXiv.org/abs/nucl-th/9701058.

[6] R. Machleidt and D.R. Entem, The nuclear force problem: Are we seeing the end of the tunnel ? Lead talk presented at the 17th International Confer- ence on Few-Body Problems in Physics (Duke University, June 2003) 5pp. arXiv:nucl-th/0309026 available at http://arXiv.org/abs/nucl-th/0309026.

[7] Kornel Lanczos, Die Funktionentheoretischen Beziehungen der Maxwell- schen Aethergleichungen — Ein Beitrag zur Relativitats-¨ und Elektronen- theorie (Verlagsbuchhandlung Josef Nemeth,´ , 1919) 80 pp. Lanczos’s dissertation is reprinted in the Appendix of W.R. Davis et al., eds., Cornelius Lanczos Collected Published Papers With Commentaries, 6 (North Carolina State University, Raleigh, 1998) A-1 to A-82.

[8] A. Gsponer and J.-P. Hurni, Lanczos’s functional theory of electrodynam- ics — A commentary on Lanczos’s PhD dissertation, in W.R. Davis et al., eds., Cornelius Lanczos Collected Published Papers With Commen- taries, 1 (North Carolina State University, Raleigh, 1998) 2-15 to 2-23. arXiv:math-ph/0402012 available at http://arXiv.org/abs/math-ph/0402012.

[9] C. Lanczos, The tensor analytical relationships of Dirac’s equation, Zeits. f. Phys. 57 (1929) 447–473. Reprinted and translated in W.R. Davis et al., eds., Cornelius Lanczos Collected Published Papers With Commentaries (North Carolina State University, Raleigh, 1998) pages 2-1132 to 2-1185. Soon to be available as arXiv:quant-ph/040xxxx at http://arXiv.org/abs/quant-ph/040xxxx.

[10] C. Lanczos, The covariant formulation of Dirac’s equation, Zeits. f. Phys. 57 (1929) 474–483. Reprinted and translated in W.R. Davis et

31 al., eds., Cornelius Lanczos Collected Published Papers With Com- mentaries (North Carolina State University, Raleigh, 1998) pages 2- 1186 to 2-1205. Soon to be available as arXiv:quant-ph/040xxxx at http://arXiv.org/abs/quant-ph/040xxxx. [11] C. Lanczos, The conservation laws in the field theoretical representation of Dirac’s theory, Zeits. f. Phys. 57 (1929) 484–493. Reprinted and translated in W.R. Davis et al., eds., Cornelius Lanczos Collected Published Pa- pers With Commentaries (North Carolina State University, Raleigh, 1998) pages 2-1206 to 2-1225. Soon to be available as arXiv:quant-ph/040xxxx at http://arXiv.org/abs/quant-ph/040xxxx. [12] A. Gsponer and J.-P. Hurni, Lanczos-Einstein-Petiau: From Dirac’s equa- tion to non-linear wave mechanics, in W.R. Davis et al., eds., Cornelius Lanczos Collected Published Papers With Commentaries, 3 (North Carolina State University, Raleigh, 1998) 2-1248 to 2-1277. Soon to be available as arXiv:quant-ph/040xxxx at http://arXiv.org/abs/quant-ph/040xxxx. [13] F. Gürsey, Relation of charge independence and baryon conservation to Pauli’s transformation, Nuovo Cim. 7 (1958) 411–415. [14] A. Gsponer and J.-P. Hurni, The Physical Heritage of W.R. Hamilton. Lec- ture at the conference “The Mathematical Heritage of Sir William Rowan Hamilton,” 17-20 August, 1993, Dublin, Ireland. arXiv:math-ph/0201058 available at http://arXiv.org/abs/math-ph/0201058. [15] A. Gsponer and J.-P. Hurni, Lanczos’s equation to replace Dirac’s equa- tion ? in J.D. Brown et al., eds, Proc. Int. Cornelius Lanczos Conf., Raleigh, NC, USA (SIAM Publ., 1994) 509–512. There are a number of typographical errors in this paper. Please refer to the corrected version arXiv:hep-ph/0112317 available at http://arXiv.org/abs/hep-ph/0112317. [16] A. Gsponer, On the “equivalence” of the Maxwell and Dirac equations, Int. J. Theor. Phys. 41 (2002) 689–694. arXiv:math-ph/0201053 available at http://arXiv.org/abs/math-ph/0201053. [17] A. Gsponer, Comment on Formulating and Generalizing Dirac’s, Proca’s, and Maxwell’s Equations with Biquaternions or Clifford Numbers, Found. Phys. Lett. 14 (2001) 77–85. arXiv:math-ph/0201049 available at http://arXiv.org/abs/math-ph/0201049. [18] S.M. Blinder, Delta functions in spherical coordinates and how to avoid losing them: Fields of point charges and dipoles, Am. J. Phys. 71 (2003) 816–818.

32 [19] B. Yu-Kuang Hu, Comment on “Delta functions in spherical coordinates and how to avoid losing them: Fields of point charges and dipoles” by S.M. Blinder, Am. J. Phys. 72 (2004) 409–410.

[20] F. Gürsey, Correspondence between quaternions and four-spinors, Rev. Fac. Sci. Istanbul A 21 (1956) 33–54.

[21] F.R. Tangherlini, General relativistic approach to the Poincare´ compensat- ing stresses for the classical point electron, NuovoCim. 26 (1962) 497–524.

[22] R.CourantandD.Hilbert, MethodsofMathematicalPhysics 2 (Interscience Publ., New York, 1962) 830pp.

[23] T. Schucker,¨ Distributions, Fourier transforms, and some of their applica- tions to physics (World Scientific, Singapore, 1991) 167pp.

[24] J.D. Jackson, Classical electrodynamics (J. Wiley & Sons, New York, sec- ond edition, 1975) 848 pp.

[25] J.D. Jackson, On the nature of intrinsic magnetic dipole moments, CERN report 77-17 (CERN, Geneva, 1 Sept. 1977) 18 pp. Reprinted in V. Stefan and V.F. Weisskopf, eds., Physics and Society: Essays in Honor of Victor Frederick Weisskopf (AIP Press, New York, Springer, Berlin, 1998) 236pp.

[26] W.R. Hamilton, Elements of Quaternions, Vol I et II (First edition 1866; second edition edited and expanded by C.J. Joly 1899-1901; reprinted by Chelsea Publishing, New York, 1969) 1185pp.

[27] Ya.M. Shnir and E.A. Tolkachev, New non-singular description of the Abelian monopole, Phys. Lett. A 183 (1993) 37–40.

[28] J.D. Jackson, The Physics of Elementary Particles (Princeton University Press, Princeton, 1958) 135pp.

[29] S. DeBenedetti, Nuclear Interactions (J. Wiley & Sons, New York, 1967) 636pp.

[30] A. Proca, New possible equations for fundamental particles, Phys. Soc. Cambridge Conf. Rep. (1947) 180–181.

[31] F. Gürsey, Connection between Dirac’s electron and a classical spinning particle, Phys. Rev. 97 (1955) 1712–1713.

[32] J.J. Sakurai, Advanced Quantum Mechanics (Addison-Wesley, 1967, Eleventh printing, 1987) 336pp.

33 [33] A. Ferrara, M. Porrati, and V.L. Telegdi, g=2 as the natural value of the tree-level gyromagnetic ratio of elementary particles, Phys. Rev. D 46 (1992) 3529–3537.

[34] R. Jackiw, g=2 as a gauge condition, Phys. Rev. D57 (1998) 2635–2636.

[35] P. Weiss, On some applications of quaternions to restricted relativity and classical radiation theory, Proc. Roy. Irish. Acad., 46 (1941) 129-168.

[36] V.V. Kassandrov, General solution of the complex 4-eikonal equa- tion and the ‘algebrodynamical’ field theory, Gravitation & Cosmol- ogy 8, Suppl.2 (2002) 57–62. arXiv:math-ph/0311006 available at http://arXiv.org/abs/math-ph/0311006.

[37] A.Ya. Burinskii, Microgeons with spin, Sov. Phys. JETP 39 (1974) 193–195.

[38] A.Ya. Burinskii, Complex Kerr geometry and nonstationary Kerr solu- tions, Phys. Rev. D 67 (2003) 124024. arXiv:gr-qc/0212048 available at http://arXiv.org/abs/gr-qc/0212048.

34 Elementary poles

Potential Field Source q m ~m d~ ~̺

δ(r) e photon 0 0 0 0 0 0 Υ(r) r−1e monopole δ(r) r−2e e 0 0 0 0 Υ(r) r−1g e−mr Yukawa δ(r) r−2g e−mr g m 0 0 0 Υ(r) r−3~µ × ~r dipole δ(r) r−4~µ × ~r 0 0 ~µ 0 0 Υ(r) r−3~µ · ~r dimonopole δ(r) r−4~µ · ~r 0 0 0 ~µ 0 Υ(r) r−2~µ vectorpole Equation (9.4) 0 0 0 0 2~µ Υ(r) r−2~r abon 0 0 0 0 0 0

Table 1: Elementary poles

35