This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

The Klein-Gordon equation with relativistic mass: a relativistic Schrödinger equation

P.-A. Gourdain Contact author: gourdain@pas.rochester.edu Physics and Astronomy Department,
Laboratory for Laser Energetics,
University of Rochester, New York 14627, USA
Abstract

The Klein-Gordon equation describes the wave-like behavior of spinless particles since it is Lorentz invariant. While it seemed initially ripe for explaining the electronic structure of the hydrogen atom, the lack of a unconditional positive probability density really limited its applications. Yet, it is intimately connected with fermions. Any solution to the Dirac equation is automatically a solution to the Klein-Gordon equation. What is even more surprising, the Klein-Gordon equation for a free particle turns into the Schrödinger equation in the non-relativistic limit. In this work we show that these problems disappear when we use the relativistic mass instead of the rest mass. While the Klein-Gordon equation losses its Lorentz invariance because of this transformation, it gains most of the features present the Schrödinger equation, including the unconditional positivity of probability density, while keeping most of its relativistic characteristics intact, including the matter-wave dispersion relation. What is even more surprising, the non-relativistic, quasi-static limit of the Klein-Gordon equation with relativistic mass is simply the Schrödinger equation under all possible conditions. So, it can be argued that this Klein-Gordon equation is a sort of relativistic Schrödinger equation.

ntroduction

The Klein-Gordon equation [1, 2] was initially seen by Schrödinger as a possible means to explain the electronic structure of the hydrogen atom. While Schrödinger had discovered this equation before Klein and Gordon, the lack of agreement between theory and experiment led him to devise his namesake equation [3] instead. It was found later that the Klein-Gordon equation actually describes spinless particles since it is Lorentz invariant[4].

Yet, any solution to the Dirac equation [5] is also a solution to the Klein-Gordon equation. Further, its dispersion relation in vacuum matches the matter-wave theory of de Broglie [6]. It is also recognized as a direct relativistic generalization of the free-particle Schrödinger equation [7]. Finally, it has a strong connection with condensed matter physics [8] and lattice dynamics [9]. It is possible to solve this equation when using different type of potentials [10, 11, 12], or adding nonlinearities [13] or time fractional derivatives [14, 15] using new numerical algorithms [16, 14, 17].

As a result, this equation should be at the heart of quantum mechanics. However, it is not the case, mostly because the density ρ\rho is not always positive and cannot be interpreted as an actual probability density, a cornerstone of the Schrödinger equation. In this paper, we show that this major problem disappears as soon as the relativistic mass is used instead of the rest mass. This simple transformation allows the density to remain positive and this density can now be interpreted as a probability density. The only requirement for this probability density to be conserved is energy conservation.

After this introduction, the paper lays the foundations necessary to construct the Klein-Gordon equation with relativistic mass, then derives its continuity equation and verifies that the density is conserved and always positive. Finally, a parallel with electromagnetism highlights a key feature behind the space-time coupling.

he Klein-Gordon equation with relativistic mass

We start with the total relativistic energy EE,

E2=p2c2+m02c4,E^{2}=\textbf{p}^{2}c^{2}+m_{0}^{2}c^{4}, (1)

of a particle with charge qq, a mass at rest m0m_{0}, and moving with a velocity v. Here p=mv\textbf{p}=m\textbf{v} is the relativistic momentum and mm the relativistic mass, i.e. m=γm0m=\gamma m_{0}, where γ\gamma is the Lorentz factor. We want to remove the rest mass from both sides of this equation to get a quantum equivalent equation where only the relativistic mass is present. To do so, we need to isolate the rest mass from the total relativistic energy, and then define all quantities as a function of the relativistic mass mm.

Isolation of the rest mass inside the relativistic energy

Using the relativistic Lagrangian LRL_{R} defined by

LR=m0c2γ,L_{R}=-\frac{m_{0}c^{2}}{\gamma}, (2)

it can be shown that the total relativistic energy of Eq. (1) is a Hamiltonian of motion when energy is conserved, since E=r˙LRr˙LRE=\partial_{\dot{\textbf{r}}}L_{R}\cdot\dot{\textbf{r}}-L_{R}. Further, since any Lagrangian LRL_{R} is indefinite with respect to addition of a constant kinetic energy [18], we can define a new energy UU simply by adding m0c2m_{0}c^{2} to LRL_{R}, i.e. U=r˙(LR+m0c2)r˙(LR+m0c2) or U=Em0c2U=\partial_{\dot{\textbf{r}}}(L_{R}+m_{0}c^{2})\cdot\dot{\textbf{r}}-(L_{R}+m_{0}c^{2})\text{ or }U=E-m_{0}c^{2}. Using this new energy we get [19]

(m0c2+U)2=p2c2+m02c4.\left(m_{0}c^{2}+U\right)^{2}=\textbf{p}^{2}c^{2}+m_{0}^{2}c^{4}. (3)

Mass for a charged particle inside an electromagnetic potential

When a particle with rest mass m0m_{0} and charge qq is moving inside an electromagnetic field with scalar potential ϕ\phi and vector potential A, the Lagrangian LPL_{P} of the particle is given by [20],

LP=LRqϕ+qAr˙,L_{P}=L_{R}-q\phi+q\textbf{A}\cdot\dot{\textbf{r}}, (4)

When the potentials ϕ\phi and A depend explicitly neither on time nor particle velocities, the total energy of the particle,

P=r˙LPr˙LP=mc2+qϕ,\mathscr{E}_{P}=\partial_{\dot{\textbf{r}}}L_{P}\cdot\dot{\textbf{r}}-L_{P}=mc^{2}+q\phi,

is conserved. First, we suppose the particle to be at rest and located at infinity, where ϕ0=0\phi_{0}=0. In this case, the total energy of the particle P\mathscr{E}_{P} is

P=m0c2\mathscr{E}_{P}=m_{0}c^{2}

Now, for this particle to move inside the potentials ϕ\phi and A, we need qϕ<0q\phi<0. This requirement can be deduced explicitly from the fact that mm0m\geqslant m_{0} since γ1\gamma\geqslant 1. In this case

P=mc2+qϕ\mathscr{E}_{P}=mc^{2}+q\phi

Because the energy of the particle is conserved, we can link the initial (at infinity) and final states, yielding

m0c2=mc2+qϕ.m_{0}c^{2}=mc^{2}+q\phi. (5)

Now, we may want to work with an electromagnetic field that varies in time. In this case, this is not the particle energy P\mathscr{E}_{P} is not conserved but this is the total energy T\mathscr{E}_{T} of the particle at rest and the electromagnetic field that is conserved. As before, we first compute the total energy at infinity

T=m0c2+12[ε0E02+B02μ0]𝑑V.\displaystyle\mathscr{E}_{T}=m_{0}c^{2}+\frac{1}{2}\int\left[\varepsilon_{0}\textbf{E}_{0}^{2}+\frac{\textbf{B}_{0}^{2}}{\mu_{0}}\right]dV.

Here E0(t,x,y,z)\textbf{E}_{0}(t,x,y,z), B0(t,x,y,z)\textbf{B}_{0}(t,x,y,z) are the total electric and magnetic fields with the particle at infinity. When the particle moves inside ϕ\phi and A we have

T=mc2+qϕ+12[ε0E2+B2μ0]𝑑V,\displaystyle\mathscr{E}_{T}=mc^{2}+q\phi+\frac{1}{2}\int\left[\varepsilon_{0}\textbf{E}^{2}+\frac{\textbf{B}^{2}}{\mu_{0}}\right]dV,

where E(t,x,y,z)\textbf{E}(t,x,y,z) and B(t,x,y,z)\textbf{B}(t,x,y,z) are the total electric and magnetic fields when the particle is moving inside the potentials ϕ\phi and A. Note that the electric and magnetic fields of the particle are included in all these electric and magnetic fields. We now define an effective mass

m~=m+12c2[ε0(E2E02)+B2B02μ0]𝑑V,\tilde{m}=m+\frac{1}{2c^{2}}\int\left[\varepsilon_{0}\left(\textbf{E}^{2}-\textbf{E}_{0}^{2}\right)+\frac{\textbf{B}^{2}-\textbf{B}_{0}^{2}}{\mu_{0}}\right]dV, (6)

which encapsulates these non-conservative processes. If we suppose that there is no outgoing energy flux at infinity (full system isolation hypothesis), the Poynting theorem guarantees the conservation of T\mathscr{E}_{T}, and we have

m0c2=m~c2+qϕ.m_{0}c^{2}=\tilde{m}c^{2}+q\phi. (7)

Despite the existence of non conservative effects, it is interesting to note that the equation above yields

tm~=qc2tϕ.\partial_{t}\tilde{m}=-\frac{q}{c^{2}}\partial_{t}\phi. (8)

We can turn any equation with non-conservative effects to an equation with conservative effects simply by setting m~m\tilde{m}\rightarrow m. However, conservative effects also require ϕ\phi and A to be time independent. To avoid any inconsistencies between time dependent and time-independent gauges, we chose to use the Lorenz gauge [21]

c2tϕ+A=0c^{-2}\partial_{t}\phi+\nabla\cdot\textbf{A}=0 (9)

throughout, which turns into the Coulomb gauge

A=0\nabla\cdot\textbf{A}=0

for steady state fields. Further, if we use Eqs. (8) and (9) we get

tm~=qA,\partial_{t}\tilde{m}=q\nabla\cdot\textbf{A}, (10)

which again is consistent when we transition to time independent phenomena. Further, we can obtain a non-relativistic equation by setting m~m0\tilde{m}\rightarrow m_{0}, and a quasi static version by setting c+c\rightarrow+\infty.

Derivation of the Klein-Gordon equation with relativistic mass

Using minimal coupling, we can replace the particle energy with the quantum energy operator (i.e. UU^qϕU\rightarrow\hat{U}-q\phi), its momentum with the momentum operator (i.e. pp^qA\textbf{p}\rightarrow\hat{\textbf{p}}-q\textbf{A}), and its rest mass with the effective mass from Eq. (7) inside Eq. (3), and we get

[(m~c2+U^)2(p^qA)2c2]ψ=[m~c2+qϕ]2ψ.\left[\left(\tilde{m}c^{2}+\hat{U}\right)^{2}-\left(\hat{\textbf{p}}-q\textbf{A}\right)^{2}c^{2}\right]\psi=\left[\tilde{m}c^{2}+q\phi\right]^{2}\psi. (11)

Here the quantum energy operator U^\hat{U} is iti\hbar\partial_{t}, p^\hat{\textbf{p}} is the quantum momentum operator p^=i\hat{\textbf{p}}=-i\hbar\nabla, and ψ\psi is the function that capture the wave behavior of the particle. Since the mass depends on time explicitly, U^\hat{U} and m~\tilde{m} do not commute and we must rewrite Eq. (11) as

[im~t22c2tt12(p^qA)2]ψ=[m~qϕ+12(q2ϕ2c2itm~)]ψ.\begin{split}\left[i\tilde{m}\hbar\partial_{t}-\frac{\hbar^{2}}{2c^{2}}\partial_{tt}-\frac{1}{2}(\hat{\textbf{p}}-q\textbf{A})^{2}\right]\psi=\\ \left[\tilde{m}q\phi+\frac{1}{2}\left(\frac{q^{2}\phi^{2}}{c^{2}}-i\hbar\partial_{t}\tilde{m}\right)\right]\psi.\end{split} (12)

We can expand the momentum operator as

(p^qA)2ψ=22ψ+2iqAψ+iqψA+q2A2ψ(\hat{\textbf{p}}-q\textbf{A})^{2}\psi=-\hbar^{2}\nabla^{2}\psi+2i\hbar q\textbf{A}\cdot\nabla\psi+i\hbar q\psi\nabla\cdot\textbf{A}+q^{2}\textbf{A}^{2}\psi

since (ψA)=Aψ+ψA\nabla\cdot(\psi\textbf{A})=\textbf{A}\cdot\nabla\psi+\psi\nabla\cdot\textbf{A}. Now, using Eq. (10), we can replace qAq\nabla\cdot\textbf{A} with tm~\partial_{t}\tilde{m} in the equation above. Once this equation is injected into Eq. (12), we get the Klein-Gordon equation for non-conservative effects

im~tψ+22ψiqAψ=[m~qϕ+q22(ϕ2c2+A2)]ψ,\begin{split}i\hbar\tilde{m}\partial_{t}\psi+\frac{\hbar^{2}}{2}\Box\psi-i\hbar q\textbf{A}\cdot\nabla\psi=&\\ \left[\tilde{m}q\phi+\frac{q^{2}}{2}\left(\frac{\phi^{2}}{c^{2}}+\textbf{A}^{2}\right)\right]&\psi,\end{split} (13)

where =2c2tt\Box=\nabla^{2}-c^{-2}\partial_{tt} is the d’Alembertian operator. When only conservative effects are considered, m~m\tilde{m}\rightarrow m, and Eq. (13) turns into the Klein-Gordon equation with relativistic mass.

onservation of the probability density

Conservation equation

We construct the probability current density using Noether’s framework [22, 23] rather than using the usual method [19] involving ψ\psi and its complex conjugate inside Eq. (13), as the mass is not constant. We start with the first order Euler-Lagrange equation given by

δδψiνδδνψi=0i{1,,N},\frac{\delta\mathcal{L}}{\delta\psi_{i}}-\partial_{\nu}\frac{\delta\mathcal{L}}{\delta\partial_{\nu}\psi_{i}}=0\,\,\,\forall i\in\{1,\dots,N\}, (14)

where \mathcal{L} is the Lagrangian density, and we used summation for repeated Greek index ν{t,x,y,z}\nu\in\{t,x,y,z\}. Here N=2N=2, since we will use only the function ψ\psi and its complex conjugate ψ\psi^{*}. We can verify that

=im~2(ψtψψtψ)+22(ψψ1c2tψtψ)+iq2(ψAψψAψ)+[m~qϕ+q22(ϕ2c2+A2)]ψψ,\begin{split}\mathcal{L}=&-i\frac{\hbar\tilde{m}}{2}(\psi^{*}\partial_{t}\psi-\psi\partial_{t}\psi^{*})\\ &+\frac{\hbar^{2}}{2}(\nabla\psi^{*}\cdot\nabla\psi-\frac{1}{c^{2}}\partial_{t}\psi\partial_{t}\psi^{*})\\ &+i\frac{q\hbar}{2}(\psi^{*}\textbf{A}\cdot\nabla\psi-\psi\textbf{A}\cdot\nabla\psi^{*})\\ &+\left[\tilde{m}q\phi+\frac{q^{2}}{2}\left(\frac{\phi^{2}}{c^{2}}+\textbf{A}^{2}\right)\right]\psi^{*}\psi,\end{split} (15)

used in Eq. (14) gives Eq.(13) using the Lorenz gauge. With the first order Euler-Lagrange equation, there is a corresponding Noether’s current density for μ,ν{t,x,y,z}\mu,\nu\in\{t,x,y,z\}

jμ=iδδμψiδψi,j_{\mu}=\sum_{i}\frac{\delta\mathcal{L}}{\delta\partial_{\mu}\psi_{i}}\delta\psi_{i}, (16)

where δψ\delta\psi and δψ\delta\psi^{*} correspond to infinitesimal variations under a U(1)U(1) phase transformation, i.e. ψeiθψ\psi\rightarrow e^{i\theta}\psi and ψeiθψ\psi^{*}\rightarrow e^{-i\theta}\psi^{*}. In this case, δψiθψ\delta\psi\rightarrow i\theta\psi, δψiθψ\delta\psi^{*}\rightarrow-i\theta\psi^{*}, νδψ=iθνψ\partial_{\nu}\delta\psi=i\theta\partial_{\nu}\psi, and νδψ=iθνψ\partial_{\nu}\delta\psi^{*}=-i\theta\partial_{\nu}\psi^{*}. Computing the four currents we get

Jt=θ(m~ψψiψtψψtψ2c2)Jμ=θ(iψμψψμψ2qAμψψ)}\left.\begin{split}J_{t}&=\hbar\theta\left(\tilde{m}\psi\psi^{*}-i\hbar\frac{\psi\partial_{t}\psi^{*}-\psi^{*}\partial_{t}\psi}{2c^{2}}\right)\\ J_{\mu}&=\hbar\theta\left(i\hbar\frac{\psi\partial_{\mu}\psi^{*}-\psi^{*}\partial_{\mu}\psi}{2}-qA_{\mu}\psi\psi^{*}\right)\end{split}\right\} (17)

with μ{x,y,z}\mu\in\{x,y,z\}. From Eq. (16) we can infer that the term inside the square brackets on the RHS of Eq. (15) does not contribute to the Noether currents. Since the Noether theorem guarantees that

νJν=0,\partial_{\nu}J_{\nu}=0, (18)

we can drop θ\hbar\theta from the four-current density.

Polar form of the conservation equation

We can write this equation in polar form using ψ=ReiS/\psi=Re^{iS/\hbar}. Here we take the modulus R(t,x,y,z):4+R(t,x,y,z):\mathbb{R}^{4}\rightarrow\mathbb{R}^{+} and the phase S(t,x,y,z):4S(t,x,y,z):\mathbb{R}^{4}\rightarrow\mathbb{R}, as this approach is compatible with the quantum operators p^\hat{\textbf{p}} and H^\hat{H} (e.g. Ref. 24). We can now compute explicitly the current density using the polar form of ψ\psi. When R0R\neq 0, we have

ν{t,x,y,z},νψψ=νlnψ,\forall\nu\in\{t,x,y,z\},\,\frac{\partial_{\nu}\psi}{\psi}=\partial_{\nu}\ln\psi,

where lnψ=lnR+iS/\ln\psi=\ln R+iS/\hbar. So, ψνψψνψ=2iR2νS/\psi\partial_{\nu}\psi^{*}-\psi^{*}\partial_{\nu}\psi=-2iR^{2}\partial_{\nu}S/\hbar, even when R=0R=0. Using this result, the four-current density becomes

Jt=[m~tS/c2]R2Jμ=[μSqAμ]R2}\left.\begin{split}J_{t}&=\left[\tilde{m}-\partial_{t}S/c^{2}\right]R^{2}\\ J_{\mu}&=\left[\partial_{\mu}S-qA_{\mu}\right]R^{2}\end{split}\right\} (19)

where μ{x,y,z}\mu\in\{x,y,z\}. Using Eqs. (8) and (19), Eq. (18) yields

tρ+vψρ=ρc2m~c2tSS.\partial_{t}\rho+\textbf{v}_{\psi}\cdot\nabla\rho=-\frac{\rho c^{2}}{\tilde{m}c^{2}-\partial_{t}S}\Box S. (20)

where

vψ=(SqA)c2m~c2tS,\textbf{v}_{\psi}=\frac{(\nabla S-q\textbf{A})c^{2}}{\tilde{m}c^{2}-\partial_{t}S}, (21)

and the density ρ\rho is R2R^{2}. Using Eq. (10), we find

vψ=c22Stm~c2vψ(m~c2tS)m~c2tS.\begin{split}\nabla\cdot\textbf{v}_{\psi}=\frac{c^{2}\nabla^{2}S-\partial_{t}\tilde{m}c^{2}-\textbf{v}_{\psi}\cdot\nabla\left(\tilde{m}c^{2}-\partial_{t}S\right)}{\tilde{m}c^{2}-\partial_{t}S}.\end{split} (22)

Adding ρvψ\rho\nabla\cdot\textbf{v}_{\psi} on both sides of Eq. (20) yields

tρ+(ρvψ)=t(m~c2tS)+vψ(m~c2tS)m~c2tSρ.\begin{split}\partial_{t}\rho+\nabla\cdot(\rho\textbf{v}_{\psi})=\\ -\frac{\partial_{t}\left(\tilde{m}c^{2}-\partial_{t}S\right)+\textbf{v}_{\psi}\cdot\nabla\left(\tilde{m}c^{2}-\partial_{t}S\right)}{\tilde{m}c^{2}-\partial_{t}S}\rho.\end{split} (23)

A necessary and sufficient condition for the conservation of the probability density

At this point, we can easily identify the momentum of ψ\psi as pψ=k\textbf{p}_{\psi}=\hbar\textbf{k}, and its energy as Eψ=ωE_{\psi}=\hbar\omega, where the angular wave vector[25] is given by

k=S/,\textbf{k}=\nabla S/\hbar, (24)

and the angular frequency[25] is

ω=tS/.\omega=-\partial_{t}S/\hbar. (25)

So, we can write Eq. (21) as

vψ=(pψqA)c2m~c2+Eψ.\textbf{v}_{\psi}=\frac{(\textbf{p}_{\psi}-q\textbf{A})c^{2}}{\tilde{m}c^{2}+E_{\psi}}. (26)

As this equation shows, vψ\textbf{v}_{\psi} conforms to the definition of a velocity according to special relativity. Using this insight, we see that the LHS of Eq. (23) is, in fact, a continuity equation for the density ρ\rho inside a velocity field vψ\textbf{v}_{\psi}. The RHS of Eq. (23) is an advection equation (i.e. material derivative) for the energy m~c2+Eψ\tilde{m}c^{2}+E_{\psi}, and vψ\textbf{v}_{\psi} is the speed at which this energy is advected. While both equations are conservation equations, there is a fundamental difference. The continuity equation implies conservation inside a static volume while the advection equation implies conservation inside a volume following along the streamlines of the vector field vψ\textbf{v}_{\psi}.

The necessary and sufficient condition required to conserve the density ρ\rho can be found in Eq. (23), namely

t(m~c2+Eψ)+vψ(m~c2+Eψ)=0.\begin{split}\partial_{t}\left(\tilde{m}c^{2}+E_{\psi}\right)+\textbf{v}_{\psi}\cdot\nabla\left(\tilde{m}c^{2}+E_{\psi}\right)=0.\end{split} (27)

Using Eq. (7) we find that the equation above is equivalent to,

t(Eψqϕ)+vψ(Eψqϕ)=0.\begin{split}\partial_{t}\left(E_{\psi}-q\phi\right)+\textbf{v}_{\psi}\cdot\nabla\left(E_{\psi}-q\phi\right)=0.\end{split} (28)

So, if the effective wave energy EψqϕE_{\psi}-q\phi is conserved then

tρ+(ρvψ)=0.\partial_{t}\rho+\nabla\cdot(\rho\textbf{v}_{\psi})=0. (29)

Hence the density ρ\rho is conserved in the sense of the continuity equation, and it is always positive when the energy in conserved. As a result, ρ\rho can be interpreted as a probability density, and ψ\psi is a probability amplitude.

Dispersion relation

When R=0\Box R=0, Eq. (12) has a simple dispersion relation (see the appendix)

(m~c2tS)2=m02c4+(kqA)2c2.\left(\tilde{m}c^{2}-\partial_{t}S\right)^{2}=m_{0}^{2}c^{4}+(\hbar\textbf{k}-q\textbf{A})^{2}c^{2}. (30)

In this case, Eq. (30) guarantees that m~c2tS0\tilde{m}c^{2}-\partial_{t}S\neq 0 for massive particles. We can use Eqs. (24) and (25) to write the dispersion relation explicitly as a function of the wave vector and the angular frequency,

ω±=m~c2±(m0c2)2+(kqA)2c2.\omega_{\pm}=-\frac{\tilde{m}c^{2}}{\hbar}\pm\sqrt{\left(\frac{m_{0}c^{2}}{\hbar}\right)^{2}+\left(\textbf{k}-\frac{q}{\hbar}\textbf{A}\right)^{2}c^{2}}. (31)

This is the de Broglie dispersion relation for matter-wave [6], where the bias term m~c2/\tilde{m}c^{2}/\hbar comes from using UU rather than EE. Under these conditions, we can now compute the group velocity of the probability amplitude of the particle

ω+k=(kqA)c2m02c4+(kqA)2c2.\frac{\partial\omega_{+}}{\partial\textbf{k}}=\frac{(\hbar\textbf{k}-q\textbf{A})c^{2}}{\sqrt{m_{0}^{2}c^{4}+\left(\hbar\textbf{k}-q\textbf{A}\right)^{2}c^{2}}}. (32)

and

ωk=(k+qA)c2m02c4+(k+qA)2c2.\frac{\partial\omega_{-}}{\partial\textbf{k}}=\frac{(-\hbar\textbf{k}+q\textbf{A})c^{2}}{\sqrt{m_{0}^{2}c^{4}+\left(-\hbar\textbf{k}+q\textbf{A}\right)^{2}c^{2}}}. (33)

The positive group velocity is the velocity of the particle with charge qq and the negative group velocity is the velocity of the dual particle with charge q-q. We see here that the particle and its dual particle travel in opposite directions. Note that we are not using the term “antiparticle” for the dual particle here because we have made no assumption on the actual value of the charge qq. It can be positive or negative. The only requirement is this paper is that qϕ<0q\phi<0. As a result, we see that the particle has an energy m~c2+Eψ>0\tilde{m}c^{2}+E_{\psi}>0 and its dual particle has an energy m~c2+Eψ<0\tilde{m}c^{2}+E_{\psi}<0. Since qq must satisfy qϕ<0q\phi<0 everywhere, then dual particles cannot propagate in this potential since their charge is q-q as it would violate Eq. (5). As a result, this work only applies to the particle only.

Using the dispersion relation above we also find that the velocity vψ\textbf{v}_{\psi} defined in Eq. (21) is the group velocity given by Eq. (32) if m~c2+Eψ>0\tilde{m}c^{2}+E_{\psi}>0 and Eq. (33) if m~c2+Eψ<0\tilde{m}c^{2}+E_{\psi}<0. Note that the amplitude RR has a phase velocity cc since R=0\Box R=0.

So, the particle is formed by a group of waves and its most probable location is where these waves interfere constructively, leading to a match between the particle velocity and the wave group velocity. Since vψ\textbf{v}_{\psi} matches exactly the velocity of a relativistic classical particle in this section, we also conclude that quantum effects are inexistent when R=0\Box R=0.

In reality, we know that particles and antiparticles can propagate in electrical potentials of any sign, the restriction herein comes from the fact that we only consider a single particle in this work. So the restriction of the sign of qϕq\phi are stringent. A many-body version of Eq. (34) would be required to capture the time evolution of the probability amplitude for a particle and its antiparticle inside a potential of arbitrary sign, a topic beyond the scope of this paper.

Discussion about energy exchange

In relativity, mass and energy are exchangeable, as a result, we expect Eq. (5) to be violated when mass is converted to energy or vice versa. However, when the wave gains or loses energy from m~\tilde{m} (i.e. conservative mass acceleration/deceleration or non-conservative electromagnetic effects, see Eq. (27)), or from the potential energy qϕq\phi (i.e. conservative time-independent or non-conservative time-dependent effects, see Eq. (28)), then the energy is conserved because there is there is really no other energies to tap on here, especially because interactions with other particles (e.g. pair creation/annihilation) cannot be captured by Eq. (13).

n interpretation of the space-time coupling

In this section, we highlight the significance of the second order time derivative inside Eq. (13). We will assume here that we can divide by the mass operator and only consider conservative processes. we can rewrite this equation as

tψ=2imψ+qmAψ+1i[qϕ+q22m(ϕ2c2+A2)]ψ.\begin{split}\partial_{t}\psi=&\frac{\hbar}{-2im}\Box\psi+\frac{q}{m}\textbf{A}\cdot\nabla\psi\\ &+\frac{1}{i\hbar}\left[q\phi+\frac{q^{2}}{2m}\left(\frac{\phi^{2}}{c^{2}}+\textbf{A}^{2}\right)\right]\psi.\end{split} (34)

First, it is interesting to note that time and space are treated on equal footings in Eq. (34) when A0A\neq 0. Further, if cc\rightarrow\infty and mm0m\rightarrow m_{0}, this equation reduces to the Schrödinger equation under any conditions. This is a departure of the Klein-Gordon equation with rest mass, which is a relativistic extension of the Schrodinger equation only for a free particle. Since the parallel is quite obvious, the physical meaning of the terms common to both equations will not be discussed here. However, the Klein-Gordon equation with relativistic mass has a second order time derivative, which is not present in the Schrödinger equation. We will now use an analogy with electromagnetism to highlight its physical significance.

For this, we look at an electromagnetic field with a scalar potential ϕE\phi_{E} and vector potential AM\textbf{A}_{M}. We have used subscripts here to make sure there is no confusion with the potentials ϕ\phi and A used in the previous sections. In the presence of a material with conductivity σ\sigma and permittivity ε\varepsilon, we can define a “damped” Lorenz gauge (e.g. see Ref. 26), derived from the standard Lorenz gauge, as

tϕE+𝒞2AM+𝒞2KϕE=0\partial_{t}\phi_{E}+\mathscr{C}^{2}\nabla\cdot\textbf{A}_{M}+\frac{\mathscr{C}^{2}}{K}\phi_{E}=0 (35)

Here 𝒞\mathscr{C} is the speed of light inside the material, given by 𝒞=μ0ε1\mathscr{C}=\sqrt{\mu_{0}\varepsilon}^{-1}, and KK is the magnetic diffusivity given by K=1/(μ0σ).K=1/(\mu_{0}\sigma). When the scalar potential is oscillatory, i.e. ϕE=ϕE(r)eiωt\phi_{E}=\phi_{E}(\textbf{r})e^{-i\omega t}, Eq. (35) turns into AM+μ0σϕE=0\nabla\cdot\textbf{A}_{M}+\mu_{0}\sigma^{*}\phi_{E}=0, allowing to define a complex conductivity σ=σiωε\sigma^{*}=\sigma-i\omega\varepsilon. Complex conductivities are often used inside the Drude model [27, 28], where it is interpreted as a lag between the drive and the response [29]. In vacuum, where σ=0\sigma=0 and ε=ε0\varepsilon=\varepsilon_{0} the magnetic diffusivity, given by

K=c2iωV,K=\frac{c^{2}}{-i\omega_{V}}, (36)

is now purely complex, due to the presence of an oscillating background potential with oscillation frequency ωV\omega_{V}. Under these conditions, the vector potential becomes [26]

tAM=KAM.\partial_{t}\textbf{A}_{M}=K\Box\textbf{A}_{M}. (37)

We can now compare the term in front of the \Box operator in Eqs. (34) and (37). We find that

K=2imK=\frac{\hbar}{-2im}

in Eq. (34), and the frequency ωV\omega_{V} defined in Eq. (36) is given by

ωV=2mc2.\omega_{V}=\frac{2mc^{2}}{\hbar}.

This frequency is exactly the quantum vacuum fluctuation frequency caused by particle and antiparticle pair creation/annihilation both with masses mm. From the parallel between Eq. (37) and Eq. (34), we can infer that the complex conductivity σ=iωVε0\sigma^{*}=i\omega_{V}\varepsilon_{0} creates a delay between the incoming (drive) and outgoing (response) waves in Eq. (34) at any given location.

onclusion

This work derived the Klein-Gordon equation with relativistic mass for a charged particle inside an electromagnetic field. Because the relativistic mass is not Lorentz invariant, this equation cannot describe spinless particles, which require Lorenz invariance. Upon inspection, we found that this equation conserves the probability density and it is always positive, as long as the energy is conserved. In particular, this conservation implies that this equation can be applied to long-lived particles. What is even more interesting, we find that the Schrödinger equation is the quasi-static (i.e. c+c\rightarrow+\infty), non-relativistic limit (mm0m\rightarrow m_{0}) version of this equation under any conditions. So, any model built upon the Schrödinger equation assumes that the speed of light is infinite, and any results obtained from it must have a superluminal behavior. In this context, we can say that the Klein-Gordon equation with relativistic mass is a kind of relativistic Schrödinger equation. Finally, a comparison with electromagnetic theory showed that the second order space-time coupling present in this equation cannot happen without quantum vacuum fluctuations. While not discussed here, the Klein-Gordon equation with relativistic mass can also accommodate spin, using Pauli’s approach [30], where the quantum operator (p^qA)2(\hat{\textbf{p}}-q\textbf{A})^{2} is replaced by [𝝈(p^qA)]2[\boldsymbol{\sigma}\cdot(\hat{\textbf{p}}-q\textbf{A})]^{2}.

cknowledgments

This research was supported by the NSF CAREER award PHY-1943939, and the NSF award PHY-2409038.

ppendix

From Eq. (7) we have m~qϕ+q2ϕ2/2c2=(m02c4m~2c4)/2c2\tilde{m}q\phi+q^{2}\phi^{2}/2c^{2}=(m_{0}^{2}c^{4}-\tilde{m}^{2}c^{4})/2c^{2}. Using this equation, the polar form of the probability amplitude, and the Lorenz gauge, Eq. (13) yields

((SqA)2c2+[m02c4(m~c2tS)2])R2c2R=ic2RS+2i[(m~c2tS)tR+c2(SqA)R],\begin{gathered}\left((\hbar\nabla S-q\textbf{A})^{2}c^{2}+\left[m_{0}^{2}c^{4}-(\tilde{m}c^{2}-\partial_{t}S)^{2}\right]\right)R\\ -\hbar^{2}c^{2}\Box R=i\hbar c^{2}R\Box S\\ +2i\hbar\left[(\tilde{m}c^{2}-\partial_{t}S)\partial_{t}R+c^{2}(\nabla S-q\textbf{A})\cdot\nabla R\right],\end{gathered}

where we dropped eiS/e^{iS/\hbar} from both sides of the equation. Since RR and SS are real, we can split the equation above into

S=2c2R[(m~c2tS)tR+c2(SqA)R]\Box S=\frac{-2}{c^{2}R}\left[(\tilde{m}c^{2}-\partial_{t}S)\partial_{t}R+c^{2}(\nabla S-q\textbf{A})\cdot\nabla R\right]

and

R=((SqA)2c2+m02c4(m~c2tS)2)R2c2.\Box R=\left((\nabla S-q\textbf{A})^{2}c^{2}+m_{0}^{2}c^{4}-(\tilde{m}c^{2}-\partial_{t}S)^{2}\right)\frac{R}{\hbar^{2}c^{2}}.

The first equation times R2R^{2} is simply the conservation equation Eq. (20). While, the second equation should be used to compute the time evolution of RR, it can be seen as the dispersion relation for solutions where R=0\Box R=0.

References

  • Gordon [1926] W. Gordon, Der Comptoneffekt nach der Schrödingerschen Theorie, Zeitschrift für Physik 40, 117 (1926).
  • Klein [1927] O. Klein, Elektrodynamik und Wellenmechanik vom Standpunkt des Korrespondenzprinzips, Zeitschrift für Physik A Hadrons and nuclei 41, 407 (1927).
  • Schrödinger [1926] E. Schrödinger, An undulatory theory of the mechanics of atoms and molecules, Physical review 28, 1049 (1926).
  • Jordan et al. [1964] T. Jordan, A. Macfarlane, and E. Sudarshan, Hamiltonian model of Lorentz invariant particle interactions, Physical Review 133, B487 (1964).
  • Dirac [1928] P. A. M. Dirac, The quantum theory of the electron, Proceedings of the Royal Society of London. Series A 117, 610 (1928).
  • de Broglie [1924] L. de Broglie, Recherches sur la théorie des quanta, Annales de Pysiques 3, 22 (1924).
  • Strange [1998] P. Strange, Relativistic Quantum Mechanics: with applications in condensed matter and atomic physics (Cambridge University Press, 1998) pp. 65–68.
  • Caudrey et al. [1975] P. Caudrey, J. Eilbeck, and J. Gibbon, The sine-gordon equation as a model classical field theory, Il Nuovo Cimento B 25, 497 (1975).
  • Dodd et al. [1982] R. K. Dodd, J. C. Eilbeck, J. D. Gibbon, and H. C. Morris, Solitons and nonlinear wave equations (Academic Press, Inc., London-New York, 1982).
  • Alhaidari et al. [2006] A. Alhaidari, H. Bahlouli, and A. Al-Hasan, Dirac and Klein–Gordon equations with equal scalar and vector potentials, Physics Letters A 349, 87 (2006).
  • Saad et al. [2008] N. Saad, R. Hall, and H. Ciftci, The Klein-Gordon equation with the kratzer potential in d dimensions, Open Physics 6, 717 (2008).
  • Onate et al. [2016] C. Onate, M. Onyeaju, A. Ikot, and J. Ojonubah, Analytical solutions of the Klein–Gordon equation with a combined potential, Chinese Journal of Physics 54, 820 (2016).
  • Dehghan and Shokri [2009] M. Dehghan and A. Shokri, Numerical solution of the nonlinear Klein–Gordon equation using radial basis functions, Journal of computational and Applied Mathematics 230, 400 (2009).
  • Golmankhaneh et al. [2011] A. K. Golmankhaneh, A. K. Golmankhaneh, and D. Baleanu, On nonlinear fractional Klein–Gordon equation, Signal Processing 91, 446 (2011).
  • Partohaghighi et al. [2022] M. Partohaghighi, Z. Mirtalebi, A. Akgül, and M. B. Riaz, Fractal–fractional Klein–Gordon equation: A numerical study, Results in Physics 42, 105970 (2022).
  • Bratsos [2009] A. Bratsos, On the numerical solution of the Klein-Gordon equation, Numerical Methods for Partial Differential Equations 25, 939 (2009).
  • Khader and Kumar [2014] M. Khader and S. Kumar, An accurate numerical method for solving the linear fractional Klein–Gordon equation, Mathematical Methods in the Applied Sciences 37, 2972 (2014).
  • Cline [2021] D. Cline, Variational principles in classical mechanics, 3rd ed. (University of Rochester River Campus Librarie, 2021) p. 212.
  • Greiner [2000] W. Greiner, Relativistic quantum mechanics, 3rd ed. (Springer, 2000) pp. 41–51.
  • Hand and Finch [1998] L. N. Hand and J. D. Finch, Analytical Mechanics, 2nd ed. (Cambridge University Press, Cambridge, 1998) p. 534.
  • Lorenz [1867] L. Lorenz, On the identity of the vibrations of light with electrical currents, The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science 34, 287 (1867).
  • Noether [1918] E. Noether, Invariante Variationsprobleme, Nachrichten von der Gesellschaft der Wissenschaften zu Göttingen, Mathematisch-Physikalische Klasse , 235 (1918).
  • Kosmann-Schwarzbach and Schwarzbach [2011] Y. Kosmann-Schwarzbach and B. E. Schwarzbach, The Noether theorems (Springer, 2011).
  • Madelung [1926] E. Madelung, Eine Anschauliche Deutung der Gleichung von Schrödinger, Naturwissenschaften 14, 1004 (1926).
  • Goldstein et al. [2001] H. Goldstein, C. Poole, and J. Safko, Classical mechanics, 3rd ed. (Pearson, 2001) pp. 434–437.
  • Tyler et al. [2004] R. H. Tyler, F. Vivier, and S. Li, Three-dimensional modelling of ocean electrodynamics using gauged potentials, Geophysical Journal International 158, 874 (2004).
  • Drude [1900a] P. Drude, Zur Elektronentheorie der Metalle, Annalen der physik 306, 566 (1900a).
  • Drude [1900b] P. Drude, Zur Elektronentheorie der Metalle; II. Teil. Galvanomagnetische und Thermomagnetische Effecte, Annalen der physik 308, 369 (1900b).
  • Ashcroft and Mermin [1976] N. W. Ashcroft and N. Mermin, Solid state physics, 1st ed. (Harcourt College Publishers, 1976) p. 16.
  • Pauli [1927] W. Pauli, Zur Quantenmechanik des magnetischen Elektrons, Zeitschrift für Physik 43, 601 (1927).