Arxiv:2009.08308V3 [Quant-Ph] 18 Dec 2020 Unveiling a Hidden

Arxiv:2009.08308V3 [Quant-Ph] 18 Dec 2020 Unveiling a Hidden

Unveiling A Hidden Classical-Quantum Link Partha Ghose∗ Tagore Centre for Natural Sciences and Philosophy, Rabindra Tirtha, New Town, Kolkata 700156, India Abstract The conceptual divide between classical physics and quantum mechanics has not been satisfactorily bridged as yet. The purpose of this paper is to show that such a bridge exists naturally in the Green-Wolf complex scalar representation of electromag- netic fields and its extension to massive fields. The quantum mechanical theory of radiation that follows from the the Green-Wolf representation is applied to the cosmic microwave background radiation (CMBR) regarded as a universal medium, and the implications are explored. Keywords: classical electrodynamics, quantum mechanics, hidden link Part I: Electrodynamics 1 Complex Scalar Representation of Classical Electrody- namics Green and Wolf have shown that classical electromagnetic fields in vacuum can be rigorously derived from a single complex scalar potential [1, 2, 3]. The Lagrangian density is arXiv:2009.08308v3 [quant-ph] 18 Dec 2020 µ 1 = ξ (∂ ψ∗(x)∂ ψ(x)) = ξ ψ˙ ∗ψ˙ ~ ψ∗.~ ψ (1) Lγ µ c2 − ∇ ∇ where the fundamental constant ξ = ~clP (with lP the Planck constant) has been introduced 3 to make ψ∗ψ have the dimension L− , and x := (t, ~x). By comparing with the conventional Lagrangian density of free electromagnetic fields in terms of the vector potential A~ which satisfies the subsidiary condition ~ .A~ =0, ∇ 1 1 1 ˙ ˙ = F µνF = A.~ A~ (~ A~).(~ A~) , (2) Lem −4 µν 2 c2 − ∇ × ∇ × ∗[email protected] 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 quantum mechanics 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 photon 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 .

View Full Text

Details

  • File Type
    pdf
  • Upload Time
    -
  • Content Languages
    English
  • Upload User
    Anonymous/Not logged-in
  • File Pages
    22 Page
  • File Size
    -

Download

Channel Download Status
Express Download Enable

Copyright

We respect the copyrights and intellectual property rights of all users. All uploaded documents are either original works of the uploader or authorized works of the rightful owners.

  • Not to be reproduced or distributed without explicit permission.
  • Not used for commercial purposes outside of approved use cases.
  • Not used to infringe on the rights of the original creators.
  • If you believe any content infringes your copyright, please contact us immediately.

Support

For help with questions, suggestions, or problems, please contact us