arXiv:2009.08308v3 [quant-ph] 18 Dec 2020 Keywords eie rmasingle fie a electromagnetic from classical derived that shown have Wolf and Green Electrody- Classical of Representation Scalar Complex 1 aifistesbiir condition subsidiary the satisfies constant fundamental the where arnindniyo reeetoantcfilsi em o terms in fields electromagnetic free of density Lagrangian omake to ∗ [email protected] namics irwv akrudrdain(MR eadda univers a m as quantum explored. regarded are The implications (CMBR) representat radiation Green-Wolf fields. the background massive the microwave to from follows extension that pape its re radiation this scalar and of complex fields Green-Wolf purpose the netic The in naturally yet. exists bridge as bridged satisfactorily been ψ h ocpuldvd ewe lsia hsc n quantu and physics classical between divide conceptual The neln idnCasclQatmLink Classical-Quantum Hidden A Unveiling ∗ ψ lsia lcrdnmc,qatmmcais idnli hidden mechanics, , classical : aetedimension the have aidaTrh,NwTw,Klaa705,India 700156, Kolkata Town, New Tirtha, Rabindra aoeCnr o aua cecsadPhilosophy, and Sciences Natural for Centre Tagore L L em complex γ = = − ξ atI Electrodynamics I: Part ( 4 1 ∂ F µ clrptnil[,2 ] h arnindniyis density Lagrangian The 3]. 2, [1, potential scalar ψ µν ξ ∗ ∇ ~ F ( L = x µν . − ) A ~ 3 ~ ∂ and , = ataGhose Partha cl µ 0 = ψ P 1 2 ( Abstract (with x ,  )= )) x c 1 2 1 =( := A. ~ ˙ l ξ A P ~ ˙ ,~xt,  − h lnkcntn)hsbe introduced been has constant) Planck the c 1 2 ) ( ycmaigwt h conventional the with comparing By . ψ × ∇ ∗ ~ ˙ ∗ ψ ˙ − A ~ ) o sapidt h cosmic the to applied is ion d nvcu a erigorously be can vacuum in lds ∇ h etrpotential vector the f rsnaino electromag- of presentation ~ . ( ψ × ∇ ~ st hwta uha such that show to is r ∗ . ∇ ~ ehnc a not has mechanics m lmdu,adthe and medium, al A ψ caia hoyof theory echanical ~  ) nk  , A ~ which (2) (1) we find the correspondence

1 1 1 ~ ψ˙ ∗ψ˙ = E.~ E,~ E~ = A˙ (3) c2 2 − c 1 ~ ψ∗.~ ψ = B.~ B,~ B~ = ~ A.~ (4) ∇ ∇ b 2 ∇ × µ To make this correspondence more precise,b let nµ satisfying n nµ = 1 be a space-like unit tangent to the wave wavefront at any point, and let −

A = ξ n e(ψ) (5) µ µR µ µ p so that ∂ Aµ = √ξ nµ∂ e(ψ)=0, and let A0 = 0 in a particular frame so that ~ .A~ = 0. Then, R ∇

F = ∂ A ∂ A = 2ξ (∂ n ∂ n ) e(ψ), (6) µν µ ν − ν µ µ ν − ν µ R Ei = F0i = ∂0Ai = 2pξ ∂0 e(ψ)ni, (7) − R B = ǫ ∂ A = 2ξ ǫ ∂ e(ψ)n . (8) i ijk j k p ijk jR k This specifies the correspondence completely.p The variational equation that follows from (1) is

1 ∂2 2 ψ(x)=0 (9) ∇ − c2 ∂t2   which is the classical wave equation of a complex massless potential. This equation is invari- µ µ ν ant under Lorentz transformations x ′ =Λν x , and a plane wave solution can be written in the form ikx i(~k.~x k0x0) ψ(x)= Ake− = Ake − (10) 2 2 2 ~ ~ with k = k0 where k = k.k, k =2π/λ. This is a classical wave of amplitude Ak, and hence need not be normalized. It follows from this form of ψ(x) that

2 2 2 (∂′ ′ )ψ′(x′) = ( ik′ ∂′ ′ )ψ′(x′) (11) 0 −∇~x − 0 0 −∇~x = (∂2 2 )ψ(x) (12) 0 −∇~x = ( ik ∂ 2 )ψ(x)=0. (13) − 0 0 −∇~x

Therefore, using the definition k0 = ω(k)/c, one obtains the Lorentz invariant equation 1 iψ˙ = 2ψ. (14) −ω(k)/c2 ∇

Now, writing the nonrelativistic Schrödinger equation in the form ~ iψ˙ = 2ψ (15) −2m∇

2 and comparing with eqn (14), one finds a surprising ‘correspondence’ between the two: the left hand sides (including the coefficient i) are identical, and the right hand sides differ only by the coefficient of 2. However, note that ∇ 1 ~ ~ ~ω(k) = := , m∗ = (16) 2 ~ 2 2 ω(k)/c ω(k)/c 2m∗ 2c and so eqn (14) has the form ~ iψ˙ = 2ψ (17) −2m∗ ∇ with m∗ as an ‘effective mass’ which transforms like ω under Lorentz transformations. This is therefore a relativistic Schrödinger-like equation for a massless particle with an ‘effective mass’ m∗. The transition from a classical wave function to a quantum wave function with a particle interpretation is brought about by the introduction of the Planck constant. A more detailed exposition of this point will follow, but to arrive at it, let us first apply the remaining time derivative in (14) on ψ to obtain the classical Helmholtz equation

ω2 2 + k2 ψ =0, k2 = (18) ∇ c2  where k is the wave number in vacuum (refractive index n = 1). Most interestingly, this classical equation (18) is derivable from the classical wave equation (9) via the intermediate equation (14) which has the mathematical structure of the Schrödinger equation! To see the crucial difference between equations (14) and (18), consider a general solution of eqn (14) of the form

ψ(x) = ρ(x) exp(iφ(x)), (19) where ρ(x) and φ(x) are real Lorentz scalarp functions. Substituting this in eqn (14) and separating the real and imaginary parts, one obtains the coupled equations

2 ω ∂φ 2 ρ(x) 2 +( φ) = ∇ , (20) c ∂t ∇ pρ(x) ∂ρ + .( φρ(x)) = 0.p (21) ∂t ∇ ∇ Thus, eqn (14) admits such general solutions provided the above conditions are satisfied. For monochromatic waves φ(x)= ~k.~x ωt, and hence ∂φ/∂t = ω, ~ φ = ~k, and one gets − − ∇ 2 2 2 ω ρ(x) k = 2 + ∇ . (22) c pρ(x) However, substitution of the same solutions forpψ in the classical Helmholtz equation (18) and separation of the real and imaginary parts result in the constraint

2 ρ(x) ∇ =0 (23) pρ(x) p 3 from the real part. This shows that the additional x-dependent term in eqn (22), which causes dispersion, vanishes in the classical case, ensuring that classical wave packets

i(~k.~x ω(k)t) 3 ψ(x)= ψ(k)e − d k (24) Z are non-dispersive. But there is no such constraint on a wave function that satisfies eqn (14). This opens up the possibility of a non-classical wave mechanics based on eqn (14) for dispersive wave packets. Since eqn (14) is the same as eqn (17), let us see what the implications are of incorporating a fundamental unit of action ~ into it. Since the action is S = d4x, Lγ scaling the wave function ψ by an arbitrary parameter λ implies that S scales by the factor R λ2. However, if the scale of the action is set by a fundamental constant ~, then it is no longer permissible to scale the wave function ψ arbitrarily, which means it must be normalized. That in turn implies that ψ∗ψ can be interpreted as a probability density. That is, indeed, Born’s rule. The wave function ψ can be normalized by requiring

3 ψ∗ψd x =1. (25) Z Writing φ = S/~, one can rewrite the solution (19) in the form

ψ(x) = ρ(x) exp(iS(x)/~), (26)

where both ρ and S are real functions.p Let us consider the stationary cases for which S(x) = W (~x) Et. Separating the real and imaginary parts and substituting in eqn.(17), one obtains the− coupled equations

∂S(x) ( S(x))2 + ∇ + Q = 0 (27) ∂t 2m∗ with ( S(x))2 ( W (~x))2 H = ∇ + Q = ∇ + Q, (28) 2m∗ 2m∗ ~2 2 ρ(x) Q = ∇ , (29) −2m∗ pρ(x) and p ∂ρ + ~ .(ρ~ W )=0. (30) ∂t ∇ ∇ Eqn. (27) is the Hamilton-Jacobi equation in electrodynamics and Eqn. (30) is a conserva- tion law (essentially the Poynting theorem). Q is known in the literature as the ‘quantum potential’. Eqn (27) shows that the evolution of the phase φ(x, t) = S(x, t)/~ is dependent on the real part of the wave function ρ(x). This is a special feature of absent in classical wave theory in which condition (23) holds, making Q vanish even though p ~ =0. 6 4 Using the relation S/~ = (W Et)/~ = φ = ~k.~x ωt for an eigenstate of energy and momentum, one gets the familiar− quantum mechanical− results

E = ~ω, ~ W = ~p = ~~k. (31) ∇ It now follows from eqns (27) and (28) that

~2k2 E = + Q = pc + Q. (32) 2m∗ This is a scaled version of eqn (22) and is a very significant result which shows that Q, the quantum potential, is the purely quantum mechanical energy which vanishes by condition (23) in classical wave theory, independent of ~. To give a concrete example of Q, one can consider the case of a in a 1D box of length a. The well known solution is ψ(x)= ρ(x) sin kx, k = nπ/a, n =1, 2, . Hence, ··· ~2n2pπ2 n~cπ Qn = 2 = (33) 2m∗a a

The lowest energy level corresponds to n = 1, and Q1 = ~πc/a is the zero-point energy. Instead of a box one can consider other time independent potentials also. It is straightforward 1 2 to see that for a harmonic oscillator potential 2 βx , for example, the zero-point energy is 1 ~ 2 ~ 2 ω0, ω0 = β/m∗ = 2βc / ω. Thus, the zero-point energy depends on the shape of the confining potential. p p Before passing on to the next topic, it would be worthwhile noting that eqn (32) can be written as H = E = pc + Q (34) from which follows the Hamiltonian equations ∂H ∂Q p˙i = = , (35) −∂xi −∂xi ∂H x˙ = = c. (36) ∂p The first of these equations has the form of Bohm’s equation for a nonrelativistic massive particle and may be interpreted as a relativistic generalization of it [4].

Commutation Relations

Now consider the general operator equations

[D , x ]= iδ (37) i j − ij where D = i∂ is the displacement operator, which must hold in classical field theories. If i − i one defines the momentum operators by pi = ~Di, this commutator can be written in the standard quantum mechanical form

[p , x ]= i~δ . (38) i j − ij 5 It is usually argued that this commutator vanishes in the limit ~ 0, the classical limit. The mathematics, however, shows that in the limit ~ 0 what one→ actually gets is 0=0. The underlying non-commutative structure (37) is independent→ of ~. In classical theory the translation operator D = i∂ and x do not commute. In quantum mechanics the i − i i operator ~Di, interpreted as the momentum operator pi, does not commute with the position operator xi. Although the physical interpretations are different, the underlying mathematical structure is the same. The change in physics comes through the Planck constant h which sets a new scale for action missing in the classical theory.

Classical and Quantum Waves Functions

Finally, let us write ψ(x)= x ψ . Let ψ = i ci ψ i where ψ i form a complete basis h | i | i | i | i 2 2 in a Hilbert space. If one defines the operator P = ψ ψ and scale ψ by λ, P ′ = λ P i | iih | | i i i and it cannot be idempotent, i.e. it cannot beP a projection operator. However, if ψ is normalized, c 2 = 1 and P is idempotent. Let us consider the pure state ρ = ψ| iψ . i | i| i | ih | It satisfies the conditions ρ2 = ρ, Trρ = 1. If an observable Oˆ = a P with discrete P i i i and nondegenerate eigenvalues a and projectors P = i i is measured on the system, then i i | ih | according to the Lüders rule [5] the state is updated to P

PkρPk ρ ρk′ = (39) → TrPkρ on the condition that the result ak was obtained. However, if one considers the total state without selection or reading of individual results, the state transforms to

ρˆ = pkρk′ = PkρPk (40) Xk Xk where pk = TrPkρ is the probability weight of the state ρk′ in the full ensemble [6]. This is the von Neumann rule. The Lüders rule clarifies its meaning and applicability. It is clear from these discussions that the fundamental differences between classical and quantum wave functions arise from two features. First, classical wave functions satisfy con- dition (23) but quantum wave functions do not. Second, the Planck constant ~ sets a new scale which requires the quantum wave function to be normalized, giving rise to discreteness in energy and momentum, projection operators and the special nature of projective mea- surements. The classical wave function can be arbitrarily scaled and has no such features. This scaling enables the amplitude and hence the intensity of classical waves to be varied arbitrarily. That freedom is not available to quantum wave functions which are normalized. Interestingly, wave functions that satisfy the quantum mechanical equation (17) are readily derivable from the classical wave equation (9) for monochromatic plane waves which are eigenstates of energy and momentum. In support of this one can cite the well known experiments of Aspect and his group [7] who have shown very clearly that classical light pulses remain classical no matter how weak (low intensity) they are made by inserting neutral density filters—they always produce classically expected coincident counts on a beam splitter. The idea that a sufficiently low intensity light pulse cannot contain more than one photon and hence must be quantum mechanical,

6 is contradicted by experiments. To observe the particle or quantum nature of light, one has to produce single photon light pulses (or squeezed states) which, when of sufficiently low flux, produce ‘anti-coincidence on a beam splitter’, the unambiguous signature of particle- like behaviour. Hence, a state of light is either classical or quantum depending on how it is prepared or produced—there is no transition from one to the other. A coherent state of quantum light is the nearest one can get to classical light, but it is essentially quantum in nature. There is thus a contextuality and complementarity between classical and quantum light: the full nature of light can only be comprehended by taking into account the mutually exclusive methods of preparing these two forms of light. And this is readily understood in terms of the theory outlined above. An important feature of time-independent quantum mechanical wave functions is their single-valuedness. In nonrelativistic quantum mechanics this follows from the ellipticity of the Schrödinger equation and the fact that Euclidean space is simply-connected [8]. Since the Helmholtz equation is also elliptic, classical wave functions that are solutions of this equation must also be single-valued in simply connected spaces.

Helicity

Let us next see how the helicity of electromagnetic radiation is described in the Green- Wolf scalar theory. Following Wolf [2] we write

ψ(~x, t)= ψ+(~x, t)+ ψ (~x, t) (41) − with

0 ∞ ψ+(~x, t) = Ψ(~x, ω)exp( iωt)dω = Ψ(~x, ω)exp(iωt)dω, (42) − 0 − Z−∞ Z ∞ ψ (~x, t) = Ψ(~x, ω)exp( iωt)dω (43) − − Z0 satisfying the relativistic Schrödinger-like equation, but ψ being complex, they are not com- plex conjugates of each other in general. They have been termed partial waves by Wolf who has shown that on time averaging, ψ+ and ψ are incoherent and represent two independent − circularly polarized components of light with helicity 1, i.e. ±

ψ∗ λψˆ = 1 (44) ± ± ± ZV ˆ σ.p where λ = σ p . It follows from this that | || |

ψ∗λψˆ =0 (45) ZV which shows that the convection current i~ ~j = ψ∗ ~ ψ ~ ψ∗ψ (46) − 2 ∇ − ∇ h i 7 does not have any helicity. However, the currents i~ ~j = ψ∗ ~ ψ ~ ψ∗ ψ (47) + − 2 +∇ + − ∇ + + i~ h i ~j = ψ∗ ~ ψ ~ ψ∗ ψ − − 2 −∇ − − ∇ − − h i (48) carry 1 helicities. Finally,± the current ~j satisfies the continuity equation ∂ρ ∂ jµ = + ~ .~j =0 (49) µ ∂t ∇

2 where j0 = ρ = ~k0 ψ > 0. | | µ µ µ Notice that the time component of the conserved current j = ψ∗∂ ψ ∂ ψ∗ψ associate with the classical wave equation (9) is not always positive definite due to− the second time derivative in the equation requiring the choice of two initial conditions. Consequently, ρ = 0 cj = ψ˙ ∗ψ ψ∗ψ˙ cannot be interpreted as a probability density. This was the historical reason for the− Pauli-Weisskopf second quantization of the Klein-Gordon equation [9]. By contrast, the time component of the conserved Schrödinger-like current jµ =(~j, ρ) is positive definite and can be interpreted as a probability density. Hence, the Pauli-Weisskopf second quantization is not mandated.

Entanglement in Classical Optics

Finally, let us consider entanglement. It is by now well known that entanglement occurs in classical optics [10, 11, 12, 13]. The reason is now clear–the classical and quantum mechanical wave functions are mathematically related. Suppose there is a bipartite classical state ψ H H where H and H are two | iAB ∈ A ⊗ B A B Hilbert spaces. Then, according to the Schmidt decomposition theorem (which dates back to 1907 [14] and is pre-quantum) it is always possible to express this state as

d 1 − π = λ i i (50) | iAB i| iA| iB i=0 X where λ are real and strictly positive, λ2 =1, and i , i are orthonormal bases i i i {| iA} {| iB} in H1,H2 respectively. The Schmidt rank d of a bipartite state is equal to the number of P Schmidt coefficients λi in its Schmidt decomposition and satisfies

d min dim(H ), dim(H ) (51) ≤ { A B } If the Schmidt rank d > 1, the state is entangled, i.e. it cannot be written as a product state. In classical optics it is always possible to consider light of unit intensity without implying normalization in the quantum mechanical sense. Hence, the mathematical result (50) is equally applicable to classical and quantum mechanical optics.

8 The two Hilbert spaces H ,H in the Schmidt decomposition have disjoint bases: i A B {| iA}∩ i B = . There is, obviously, nothing in the mathematical theorem that tells us how the disjointness{| i } ∅ is to be physically realized. For intra-system bipartite entanglement (i.e. en- tanglement between two different degrees of freedom of a single system), one can have, for example, path-polarization entanglement in classical optics and path-spin entanglement in quantum mechanics where the choice of paths (strictly speaking, disjoint spatial modes) is restricted to two. For inter-system entanglement (entanglement between two different sys- tems) one can have polarization-polarization entanglement in both classical and quantum optics, the dimension of the Hilbert spaces being 2 in both cases. In this case the spatial wave functions of the two systems remain in product form. So far only intra-system entangle- ment has been experimentally studied in classical optics [13], but extension to inter-system entanglement is possible in principle. It should be clear therefore that the mathematical structure of entanglement is funda- mentally the same for classical and quantum mechanical radiation, though because of the normalization of the quantum states forced by the Planck constant, projective measurements play a role in quantum radiation that has no counterpart in classical radiation. Interaction with Matter Before passing on to the implications, let us briefly consider the interaction of radiation with Dirac particles. The Lagrangian desnity is

µ µ µ = Ψ(¯ iγ ∂ mc/~)Ψ+ ∂ ψ∗∂ ψ + eΨ¯ γ Ψ∂ ψ (52) L µ − µ µ which is invariant (to within total four-divergence terms) under the local gauge transforma- ieθ tions Ψ′ = e Ψ, ψ′ = ψ + θ (θ real) with the restriction θ = 0. Part II: Massive Electrodynamics The Green-Wolf complex scalar representation of electromagnetic fields in vacuum turns out to be crucial in formulating a satisfactory theory of radiation encompassing both its classical and quantum aspects. We will now show that it can be extended to massive fields. Following Part I, let us consider the Lagrangian density of a massive complex scalar field in vacuum, µ 2 = ~cl ∂ ψ∗∂ ψ µ ψ∗ψ (53) L P µ − where µ is an arbitrary constant with the dimension of inverse length. Notice that it is 1 possible to choose µ = lP− . The classical wave equation that follows from it is 1 ∂2 2 µ2 ψ(x)=0. (54) ∇ − c2 ∂t2 −   Let i(~k.~x k0x0) ψ(x)= Ake − (55) 2 2 2 be a monochromatic plane wave solution with k + µ = k0. Following the same procedure as in Part I and applying the operator ∂0 once on ψ results in the equation ∂ψ c µ2c i = 2 + ψ, k = ω(k)2/c2 µ2, (56) ∂t −k ∇ k −  0 0  p 9 which has the mathematical structure of the Schrödinger equation with a k0 dependent potential, and can indeed be written in the form

∂ψ ~ i = 2 + V ψ (57) ∂t −2m ∇ 0  ∗  where 2 ~k0 µ c m∗ = , V0 = . (58) 2c k0 The principal difference from the non-relativistic Schrödinger equation is the occurrence of the effective mass m∗ which transforms like k0 under Lorentz transformations. Writing 2 2 1/2 2 2 m∗ = γm , γ = (1 v /c )− and ignoring terms of (v /c ), one obtains 0 − O ~2 ~2 2 ∂ψ 2 µ 1 2 i~ = + V ′ ψ, V ′ = = m c , (59) ∂t −2m ∇ 0 0 2m 2 0  0  0 which is the nonrelativistic Schrödinger equation with a constant potential. This shows that eqn.(57) is the correct relativistic generalization of the Schrödinger equation. Applying the time derivative on ψ in eqn (56), one obtains the Helmholtz equation

2 + k2 ψ = 0, (60) ∇ ω(k)2 k2 = k2 µ2 = µ2 (61) 0 − c2 − where k is the wave number in vacuum (refractive index n = 1). Most interestingly, this classical equation (60) is derivable from the classical wave equation (54) via the intermediate equation (56) which has the mathematical structure of the Schrödinger equation! Eqn (57) ensures that the convection current

i~ ~j = ψ∗ ~ ψ ~ ψ∗ψ (62) − 2 ∇ − ∇ h i is conserved, ∂ρ + ~ .~j =0, ρ = ~k ψ∗ψ > 0 (63) ∂t ∇ 0 and its time component ρ > 0. It is therefore possible to interpret ψ∗ψ as the position probability density and ~j as the probability current density when ψ is normalized. Hence, eqn (57) can be given a quantum mechanical particle interpretation. The classical wave equation (54) does not have this property because the time component µ µ µ of its conserved current j = ψ∗∂ ψ ∂ ψ∗ψ is not always positive definite due to the second − time derivative in the equation requiring the choice of two initial conditions. Consequently, 0 ρ = cj = ψ˙ ∗ψ ψ∗ψ˙ cannot be interpreted as a probability density. This was the historical − reason for the Pauli-Weisskopf second quantization of the Klein-Gordon equation which is no longer mandated. Now consider a general solution of eqn (56) in the polar form

ψ(x)= ρ(x) exp(iφ(x)), (64) p 10 where ρ(x) and φ(x) are real functions. Substituting this in eqn (56) and separating the real and imaginary parts, one obtains the coupled equations

k ∂φ 2 ρ(x) 0 +( φ)2 + µ2 = ∇ , (65) c ∂t ∇ pρ(x) ∂ρ + ~ .(~ φρ(x)) = 0.p (66) ∂t ∇ ∇ These are therefore the conditions that must hold for general solutions of eqn (56). For monochromatic solutions, φ(x)= ~k.~x k ct, ∂φ/∂t = k c, ~ φ = ~k, and one gets − 0 − 0 ∇ 2 2 2 2 ρ(x) k = k0 µ + ∇ − pρ(x) 2 2 ω(k) p 2 ρ(x) = 2 µ + ∇ (67) c − pρ(x) Substitution of the same solutions in the classical Helmholp tz equation (60) and separation of the real and imaginary parts result in the condition

2 ρ(x) ∇ =0 (68) pρ(x) from the real part. This shows that thep additional x-dependent term in eqn (67), which causes dispersion, vanishes in classical theory, ensuring that classical wave packets are non- dispersive. But, as in the massless case studied in Part I, there is no such restriction on a wave function satisfying eqn (56), which therefore forms the basis of a non-classical wave mechanics with dispersive wave packets. 2 Eqn (57) with m∗ = ~ω/2c is the same as eqn (56) except that the unit of the action S is set to be ~. Hence it is no longer permissible to scale the wave function ψ arbitrarily, which means it must be normalized and can be interpreted as a probability density. Other important consequences of normalization of the wave function have been discussed in Part I. Since ψ describes a massive field, there is a longitudinal component of the polarization vector in this case in addition to two transverse components.

2 Quantum and Classical Particles

Multiplying eqns (65) and (66) by the arbitrary unit of action η and writing ηφ = S, we get

∂S ( S)2 + ∇ + V + = 0, (69) ∂t 2m 0 Q η2 2 ρ(x) ηµ2c = ∇ , V0 = , 2m = ηk0/c. Q −2m pρ(x) k0 p 11 and ∂ρ + ~ .(~ Sρ)=0. (70) ∂t ∇ ∇ Eqn (69) is the Hamilton-Jacobi equation for a massive particle in a potential V + . For 0 Q stationary eigenstates of energy and momentum one can set S = W Et, ~p(= γm~v)= ~S = − ∇ ~ W . Then, ∇ p2 H = + V + (71) 2m 0 Q and hence

∂H pi x˙ i = = , (72) ∂pi m ∂H ∂ p˙i = = Q (73) −∂xi −∂xi Eqn (73) would the relativistic version of Bohm’s equation for a massive particle in a quantum potential Q if one were to identify η with ~. It is the quantum potential that gives rise to interference of quantum particles [15]. Notice that condition (68) prevents dispersion and at the same time causes , the term responsible for quantum mechanical coherence, to vanish. It is therefore a sufficientQ condition for Newton’s equation to hold. Eqn (69) then takes the form ∂S cl + H = 0, (74) ∂t ( W )2 p2 H = ∇ cl + V = + V =0, (75) 2m 0 2m 0 It follows from this that

∂H pi x˙ i = = , (76) ∂pi m ∂H p˙i = =0. (77) −∂xi The absence of interference indicates that there is no fixed phase relationship between dif- ferent points of the wave amplitude. This follows from eqn (74) which shows that the phase

φ = Scl/η is independent of ρ(x). Hence, one cannot write a coherent superposition i ciψi of wave functions describing a classical particle. However, one can still write a density matrix: p P ρˆ = c 2 ψ ψ . (78) | i| | iih | i X A similar situation obtains in the Koopman-von Neumann wave theory of nonrelativistic classical mechanics in which the particle wave function satisfies the Liouville equation [16, 17]. All this shows that eqn (56) for a wave function ψ that satisfies condition (68) is equivalent to Newton’s equation of motion for a free massive particle. Hence, the same equation, namely eqn (56), holds for both quantum and classical mechanics of relativistic particles depending on whether or not the wave function satisfies a certain condition. It describes quantum

12 particles if the wave function is normalized and does not satisfy condition (68), and classical particles if it is not normalized and satisfies condition (68). Eqns (73) and (77) are second order differential equations in time and their solutions require two initial conditions specifying the position and velocity which can be varied inde- pendently. In quantum mechanics this is not permissible and solutions of eqn (73) require special care. There is no such restriction on eqn (77) which is classical. Further, in a theory in which the classical and quantum aspects of a system are intrinsically linked, they have the same ontology, and hence the de Broglie-Bohm type of interpretation [4] is a natural choice.

3 Measurements

Quantum mechanics presumes classical measuring apparatus with which quantum systems interact. This has been a fundamental problem since the inception of quantum mechanics because the two systems appeared so disparate, the quantum system being described by a ray in a Hilbert space and the classical system by a point in phase space. The option of treating the measuring apparatus also as a quantum system gave rise to the measurement problem which refuses to go away. A new option is now available, namely the use of a wave function in a Hilbert space for the classical measuring apparatus. Let us consider the case of an observation designed to measure some observable Pˆ of a stationary quantum system S with wave function ψS(x, t). Let the stationary classical wave function of the apparatus A be ψA(y, t) where y is the coordinate of the ‘pointer’. The initial state is a product state

ΨSA(x ,y , 0) = ψS(x , 0) ψA(y , 0) = ψA(y , 0) c ψS(x , 0) (79) 0 0 0 ⊗ 0 0 p p 0 p X ˆ S x S x where P ψp ( ) = pψp ( ). This is a hybrid wave function. This kind of wave function was first introduced by Sudarshan [18]. Following von Neumann, let us assume that the measurement interaction is impulsive, and that during this impulsive interaction the free evolutions of the quantum particle and the classical apparatus can be ignored because the mass of the particle is very large and the mass of the apparatus (the massive particle) can always be chosen to be sufficiently large. If one sets ~ =1 for convenience, the evolution operator of the system takes the form

U = exp( iΩˆt), (80) − Ωˆ = gPˆDˆ (81) − y where g is a suitable coupling strength and Dˆ y = i∂/∂y is the classical displacement oper- ator corresponding to the coordinate y of the apparatus.− The form (81) of the measurement interaction has been chosen to be of the von Neumann type. Then,

∂ A igpDˆ yt A yp A A Uψ (y, t)= e− ψ (y , 0) = e− ∂y ψ (y , 0) = ψ (y y ), y = gpt (82) 0 0 0 − p p for every p and for t τ, the measurement time which is assumed to be extremely short. For t > τ there is no≤ further displacement of the pointer. Hence, in accordance with (78),

13 the final stationary state is of the form

ρˆSA = c 2 p S p p A p . (83) | p| | i h || i h | p X 2 Each pointer position is correlated with a particular outcome p with probability cp , the correlation being exact in the limits of both g and the number of trials tending to| infinity.| The mixed state ρˆS of the quantum system S alone after the measurement can be obtained by tracing ρˆSA over the apparatus states:

ρˆS = Tr ρˆSA = c 2 p S p (84) A | p| | i h | p X which is formally the same as the standard von Neumann mixed density matrix but does not imply a process of collapse. Thus, we have a unified theory of classical and quantum systems (intrinsically relativis- tic) in which measurement does not occupy any special significance, and the two systems naturally share the same ontology. Part III: Implications

In the previous parts a hidden mathematical link between quantum and classical radiation has been used to develop a relativistic quantum mechanical theory of radiation (as opposed to second quantized quantum electrodynamics). In this part I will explore the possible implications of treating the cosmic microwave background radiation (CMBR) as such a quantum mechanical system and a universal medium.

(ii) Quantum Mechanics of Blackbody Radiation and CMBR

In order to have a proper quantum mechanical theory of blackbody radiation, it is neces- sary to generalize the single photon wave function considered in Part I to the many-photon s case. Consider a state of N placed in A = s A states with occupation num- s s bers (p0,p1, pM ), pr = p , N = N , each photon being placed in one of the states ··· s r max P pj , j =0, i.e. | i 6 x ,P x , , x p ,p ,p , p = ΠN ψ (x ) (85) h 1 2 ··· N | 0 1 2 ··· M i j=1 pj j where xj = (~xj, t) are the coordinates of the particles. The wave function with the correct permutation symmetry is therefore given by 1 ψ(x1, x2, , xN )= P x1, x2, , xN p0,p1,p2, ,pM (86) ··· √W h ··· | ··· i P Λ(p1, ,pM ) ∈ X··· where Λ(p , ,p ) S is the set of all permutations of the p involving different p s. 1 ··· M ∈ M j j This reflects the fact that a photon in a state with a given occupation number pj can come from any of the positions xi. This is the required generalization of the single photon wave function.

14 To derive the Planck formula for blackbody radiation, one can then follow Bose’s method of distributing photons in such quantum states, calculating the macroscopically defined prob- ability W of a state having all types of quanta, the Boltzmann entropy S = klnW , and s s maximising it subject to the constraint that the total energy E = s N hν remains fixed (see the Appendix for details). Now, the cosmic microwave background radiation (CMBR) hasP a Planck spectrum to a high degree of accuracy and can therefore be treated as blackbody radiation. It has been shown in Part I that quantum mechanical radiation has zero-point energy. According s to Bose’s derivation of the Planck law (see Appendix), there are s p0 states in Planck radiation which have no photons, where P s2 s s s hνs/kT s 8πν dν p = A (1 e− ), A = (87) 0 − c3 according to the results (105, 107) in the Appendix. These numbers vanish if hνs =0. They are therefore ‘vacuum states’ with energy. The vacuum energy density of Planckian radiation is therefore

ωc 2 8πω ~ω/kT ρ = ~ω 1 e− dω vac (2π)3c3 − Z0 ~2ω5 ~ω  c , 1 (88) ≃ 5π2c3kT kT ≪

where ωc is a cut-off frequency. Since the CMBR spectral intensity is exponentially damped at high frequencies, its vacuum energy density must also be cut-off beyond some frequency ωc. One must now estimate the cut-off frequency. This can be done, for example, from the anomalous magnetic moment of the on the assumption that it is entirely caused by the CMBR vacuum density.

(iii) Anomalous Magnetic Moment of the Electron and Dark Energy

According to the Dirac theory, the electron has a spin magnetic moment µ = gµ σ /~ = h i Bh i µB where µB = e~/2mec, σ = ~/2 and g = 2. When placed in the CMBR, it interacts with the external electromagnetich i field, and its vertex function is given by iσµν q Γµ = F (q2)γµ + F (q2) ν (89) 1 2 2m 2 µ where q = q qµ is the momentum transfer, and empirically the two form factors are known to satisfy the conditions F1(0) = 1 and F2(0) = ae =(g 2)/2. A non-zero ae is called the ‘anomalous magnetic moment’ of the electron because in− the reigning paradigm of QED the assumption is that F2(0) = 0 in vacuum and in the absence of loop corrections to the vertex function. The non-zero value of ae is then shown to arise from loop corrections to the vertex function which are divergent, but a method exists to extract unambiguous finite results from them. In the one-loop approximation, ae α/2π which is very close to the observed value [21]. ≃ In the relativistic quantum mechanics of radiation developed in the previous paper, the anomalous magnetic moment of the electron can be related to the CMBR vacuum energy

15 in a non-perturbative and phenomenological way. It has also been shown in the previous paper that the quantum potential Q is the source of all forms of quantum mechanical energy of radiation, including its zero-point energy. Now, the term ψ∗. ψ in the Green-Wolf ∇ ∇ Lagrangian density for radiation (1),

µ 1 = ∂ ψ∗∂ ψ = ψ˙ ∗ψ˙ ψ∗. ψ, (90) Lγ µ c2 −∇ ∇ can be written as

2 ψ∗. ψ = ψ ψ∗ + .( ψ ψ∗) ∇ ∇ −∇ ∇ ∇ 2√ρ ω = ∇ ρ + = Qρ + (91) − √ρ ··· ~c2 ···

where is a total 3-divergence term. According to the correspondence (4), this energy ··· density is magnetic. There is therefore a constant effective magnetic field B in the CMBR 1 B B vacuum with energy density 8πµ0 . = ρvac. located within CMBR therefore acquire a Larmor energy ~ωL = aeµBB = ξρvacV (92) where 0 <ξ 1 is an unknown ‘energy transfer efficiency factor’ which we take to be unity ≤ to illustrate the basic physics. Therefore,

~2 5 ρvac V ωc 2mec V ae = = 2 3 µB B 5π c kT e~ B 71 5 3  3    5.5 10− ω erg/cm cm = × c (93) 9.274 10 21 erg/G G × −  

One obtains ae = α/2π 0.0011614 for ωc = 2.87 GHz. For such a value of ωc, ρvac 23 3 ≃ 6 3 ∼ 10− erg/cm which is well within the observed upper bound < 10− erg/cm . In quantum electrodynamics, on the other hand, the vacuum energy per normal mode is ~ω/2 and the result is ~ω4 ρQED = c (94) vac 8π2c3 which diverges because there is no natural cut-off in the theory. Assuming a Planck energy QED 114 3 scale cut-off, ρvac is 10 erg/cm which is some 120 orders of magnitude larger than the observational upper bound∼ [22, 23]. The fact that the vacuum energy of CMBR calculated from the relativistic c-number quantum ‘mechanics’ of radiation developed in this paper is well within the observed upper bound on it, unlike the QED value, is therefore a point in its favour.

(iv) CMBR and Spontaneous Emissions

The presence of CMBR in the universe is ubiquitous, and practically all the light in the universe is spontaneously emitted. Might there be a connection between the two? There might indeed be one. To see how, consider a 2-level atomic system with an energy gap E E = hν, and let a monochromatic beam of radiation of frequency ν and number 2 − 1 16 density N ν of photons be incident on it. The probability of absorption of the incident ν radiation by the system will be proportional to n1N hν where n1 is the number density of the atoms in the ground state. In the absence of CMBR the probability of emission of a ν photon by the atom will be n2N hν where n2 is the number density of the atoms in the excited state. In the presence of CMBR N ν will change to N ν + Aν where Aν =8πν2/c3 is the number density of CMBR states with the same frequency. Hence the emission probability ν ν will change to n2(N + A )hν. Consequently, in thermal equilibrium the condition

ν ν n2(N + A )hν g2 ν = (95) n1N hν g1 must hold, where g1,g2 are the degeneracies (multiplicities) of the two levels [24]. This is, hν/kT in fact, just the Planck law since the Boltzmann distribution law n1g2/n2g1 = e must also hold. The additional term Aν in the numerator causes spontaneous emissions with the correct Einstein coefficient. Contrast this with the corresponding Einstein equilibrium condition

n1B12ρ(ν)= n2 (A21 + B21ρ(ν)) (96)

together with the Boltzmann law, from which the Planck law follows only by imposing the ν ad hoc conditions A21/B21 = A and B12/B21 = g2/g1. Historically, the basic Dirac theory of QED which gave the first explanation of sponta- neous emission dates back to 1927, i.e. much before the discovery of CMBR. In QED the ′ ′ spontaneous emission term arises from the commutation relation [aν† , aν ] = δνν where aν† and aν are creation and annihilation operators of photons of frequency ν, which gives rise to divergences and to an unacceptably large vacuum energy density in the universe. The alternative simple theory presented here is based on the application of a relativistic ‘quantum mechanics’ of radiation to CMBR.

(v) CMBR and the Casimir Effect

A typical example of the Casimir Effect is the tiny attraction between two uncharged conductive plates placed a few nanometers apart in a vacuum [25]. The magnitude of the effect depends on the shape of the plates or the confining region. The effect is believed to be caused by the plates changing the vacuum energy of the electromagnetic field between them [26]. All calculations using QED turn out to be divergent, but there are methods of regularization which can deal with that. The vacuum energy of CMBR may also contribute to this effect. To illustrate this in the simplest case, let us calculate the effect produced by a 1D box of length a. We know 5 from eqn (88) that the vacuum energy of CMBR is ρvac = Cωc where C is a constant, and ωc = πc/a, a being the separation between the plates. Hence

4 2 3 2 17 ∂ρ 5Cπcω ~ π c 1 7.5 10− P = vac = c = × dyne/cm2 (97) ∂a − a2 − kT a6 ≃ − a6 where C = ~2/5π2c3kT and T = 2.7K. The negative sign indicates an attraction between 9 2 5 the plates. A pressure P 10 dyne/cm would thus result for a 4 10− cm. | | ∼ ∼ × 17 4 Concluding Remarks

That the Green-Wolf complex scalar representation of electromagnetic fields would reveal the much needed mathematical and conceptual link between the c-number ‘quantum mechanics’ of radiation and the classical field theory of radiation comes as a surprise. It lays the mathematical and conceptual foundation of wave-particle duality originally discovered by Einstein in the energy fluctuations of Planck radiation. The fundamental role of the Planck constant in forcing the normalization of the wave function and hence the Born rule, becomes transparent. The classical time independent Helmholtz eqn (18) is derivable from the time dependent classical wave equation (9) through the intermediate equation (14) which has an essentially Schrödinger-like structure. 2 √ρ ∇ The function √ρ plays a fundamental role in the theory. Its presence or absence deter- mines whether the theory is quantum mechanical or classical. The Helmholtz equation forces this term to vanish, ensuring dispersion free classical waves in vacuum. Its presence allows non-classical, or quantum mechanical, waves to disperse in vacuum. It also determines the functional form of the quantum potential Q responsible for all quantum mechanical features like quantum coherence and quantized energy levels. It is noteworthy that the quantum potential, a typical feature of nonrelativistic de Broglie-Bohm theory [4], emerges naturally in a relativistic theory. The generalization to massive electrodynamics is straightforward and leads to a theory of classical massive particles, and hence of classical measuring devices, obeying the Schrödinger 2 √ρ ∇ equation with the supplementary condition √ρ = 0 on the wave amplitude, and hence a satisfactory theory of measurement that does not require a collapse postulate. The generalization of the Green-Wolf complex scalar representation to massless classical Yang-Mills fields and their quantum mechanical theory is under investigation. That should be of great importance for the standard model of particle physics as well as for Einstein’s gravitational equations which have close relationships with Yang-Mills equations [19, 20]. In Part III the relativistic quantum mechanical theory of radiation developed in Part I has been applied to CMBR treated as a universal medium. The implications are: (i) a finite vacuum energy of CMBR which is consistent with the observational upper bound on the cosmological constant, (ii) a finite anomalous magnetic moment of the electron immersed in CMBR, consistent with its observed value, (iii) a finite Casimir Effect of the right order of magnitude due to CMBR vacuum energy, and (iv) a natural explanation of spontaneous emis- sion of photons from atomic and molecular systems immersed in CMBR. In second quantized electrodynamics (QED) these effects are due to vacuum fluctuations at zero temperature, and the results are divergent though they can be regularized.

5 Acknowledgement

I am grateful to A. K. Rajagopal and Partha Nandi for many helpful discussions.

Appendix: Bose’s Derivation of the Planck Law

It will be quite instructive to follow Bose’s original derivation of Planck’s law [27] which

18 is quite different from the accounts given in text books and is therefore unknown to most physicists. Let us consider a collection of quantum states with frequency lying between νs and νs +dνs in a hohlraum of volume V . Assuming a spherically symmetric V , one essentially has stationary radiation in a 1D box. As shown by Bose [27], the total number of such states per unit volume in the range dνs is

8πνs2dνs As =2 d3ps/h3 = (98) c3 Z (using ps = hνs/c), the factor 2 being due to helicity 1. This is therefore the number of ± possible arrangements of a single photon. As argued by Bose, all possible arrangements of s s s s the photons in these states will correspond to a state p0,p1,p2, where p0 is the number s |s ···i of empty states, p1 the number of 1-photon states, p2 the number of 2-photon states, etc. Treating states with a given occupation number as identical, the probability of a state is given by s s A ! W = s s s , (99) p0!p1!p2!... s s s s s the number of photons of type ν being N = r rpr and A = r pr. The macroscopically defined probability of a state having all types of quanta is thus P P s s A ! W = ΠsW = Πs s s s , (100) p0!p1!p2!...

s s the total number of photons of all types being N = s N . Taking the pr to be large, one has ln W = As ln As P ps ln ps (101) − r r s s r X X X with s s A = pr. (102) r X This must be maximized subject to the constraint

E = N shνs = constant. (103) s X Carrying out the variations, one gets 1 δps(1 + ln ps)+ hνs rδps =0. (104) r r β r s r s r X X X X It follows from this that rhνs s s β pr = B e− . (105) But, since s s 1 s s rhν s hν − A = B e− β = B 1 e− β , (106) − r X  

19 we have s s s hν B = A 1 e− β . (107) − Further,  

s s s s s hν rhν N = rp = A 1 e− β e− β (108) r − r r   X X hνs s A e− β = hνs , (109) 1 e− β − and using the result (98), one gets

hνs s2 s β 8πν dν s e− E = 3 V hν hνs . (110) c 1 e− β − Now, s E s hν S = k A ln 1 e− β . (111) β − − " s # X   ∂S 1 Hence, using the relation ∂E = T , we have β = kT . Substituting this in the expression for E and dropping the suffix s from νs, we finally get the Planck formula E 8πν2 hν ρ(ν)dν = = 3 hν dν. (112) V c e kT 1 − Notice that only energy conservation plays a role in the derivation of this formula, but not photon number conservation.

References

[1] H. S. Green and E. Wolf, A Scalar Representation of Electromagnetic Fields, Proc. Phys. Soc. A. (1953) 66, 1129-1137.

[2] E. Wolf, A Scalar Represntation of Electromagnetic Fields: II Proc. Phys. Soc. (1959) 74, 269-280.

[3] P. Roman, A Scalar Represntation of Electromagnetic Fields: III Proc. Phys. Soc. (1959) 74, 281-289.

[4] D. Bohm, A Suggested Interpretation of the Quantum Theory in Terms of ”Hidden” Variables, I and II, Phys. Rev. (1952) 85, 166, 180. L. de Broglie, Non-linear wave mechanics, Elsevier, Amsterdam (1960).

[5] Lüders, G., Über die Zustandsänderung durch den Meßprozeß, Annalen der Physik (1951) 8, 322-328. https://doi.org/10.1002/andp.200610207. English translation by K. A. Kirk- patrick: Concerning the state-change due to the measurement process, Ann. Phys. (2006) (Leipzig) 15, 663-670, see quant-ph/0403007v2.

20 [6] Busch, P. and Lahti, P., Lüders Rule. Compendium of Quantum Physics: Concepts, Experiments, History and Philosophy. Eds D. Greenberger, K. Hentschel and F. Weinert. Springer, Berlin, Heidelberg (2009), 356.

[7] P. Grangier, G. Roger and A. Aspect, Experimental Evidence for a Photon Anticorrela- tion Effect on a Beam Splitter: A New Light on Single-Photon Interferences, Europhys. Lett. (1986) 1, 173.

[8] J. Riess, Single-valued and multi-valued Schrödinger wave functions, Helv. Phys. Acta (1972) 45, 1066-1073.

[9] W. Pauli and V. F. Weisskopf, Avoiding Negative Probabilities in Quantum Mechanics, Helvetica Physica Acta 7 (1934), 709-731.

[10] R. J. C. Spreeuw, Classical wave-optics analogy of quantum-information processing, Phys. Rev. A (2001) 63, 062302.

[11] P. Ghose and A. Mukherjee, Entanglement in Classical Optics, Rev. Theor. Sci. (2014) 2, 1-14.

[12] X-F Qian, B. Little, J. C. Howell and J. H. Eberly, Shifting the quantum-classical boundary: theory and experiment for statistically classical optical fields, Optica (2015) 2 No. 7, 611-615.

[13] A. Aiello et al, Quantum-like nonseparable structures in optical beams, New J. Phys. (2015) 17, 043024.

[14] E. Schmidt, Zur Theorie der linearen und nichtlinearen Integralgleichungen Math. Ann. (1907) 63, 433.

[15] C. Philippidis, C. Dewdney and B. J. Hiley, Quantum interference and the quantum potential, Il Nuovo Cimento B (1979) 52(1),15-28.

[16] B. O. Koopman, Hamiltonian Systems and Transformations in Hilbert Space, Proc. Natl. Acad. Sci. U.S.A. (1931) 17, 315.

[17] J. von Neumann, Zur Operatorenmethode in der Klassischen Mechanik, Ann. Math. (1932) 33, 587-642; Zusatze Zur Arbeit "Zur Operatorenmethode...ibid. (1932) 33, 789- 791.

[18] E. C. G. Sudarshan, Interaction between classical and quantum systems and the mea- surement of classical observables, Pramana (1976) 6 (3), 117-126. Sudarshan was the first to write down a measurement interaction between classical and quantum systems by embedding the classical system in a quantum system with a contin- uum of superselection rules that ensure the absence of interference in classical mechanics.

[19] A. Ashtekar, T. Jacobsen and L. Smolin, A new characterization of half-flat solutions to Einstein’s equation, Commun. Math. Phys. (1988) 115, 631-648.

21 [20] L. J. Mason and E. T. Newman, A connection between the Einstein and Yang-Mills equations, Commun. Math. Phys. (1989) 121, 659-668.

[21] J. S. Schwinger, On Quantum-Electrodynamics and the Magnetic Moment of the Elec- tron, Phys. Rev. (1948) 73, 416-417.

[22] S. E. Rugh and H. Zinnkernagel, and the cosmological constant problem, Studies in Hist. and Phil. of Mod. Phys. (2002) 33, 663-705.

[23] P. J. E. Peebles and B. Ratra, The cosmological constant and dark energy Rev. Mod. Phys. (2003) 75, 559-606.

[24] S. N. Bose, Wärmegleichgewicht im Strahlungsfeld bei Anwesenheit von Materie, Zeit. f Phys. (1924) 27, 384-393.

[25] H. B. G. Casimir, On the Attraction Between Two Perfectly Conducting Plates, Kon. Ned. Akad. Wetensch. Proc. (1948) 51, 793.

[26] M. Bordag; G. L. Klimchitskaya, U. Mohideen, and V. M. Mostepanenko, Advances in the Casimir effect, Oxford University Press, Oxford (2009).

[27] S. N. Bose, Plancks Gesetz und Lichtquantenhypothese, Zeit. f Phys. (1924) 26, 168-171.

22