1. Introduction
The phenomenon of entanglement is a distinguishing feature of quantum theory and an increasingly important topic in contemporary physics, from the continuing debate of philosophical questions deriving from the papers of Einstein, Podolsky, and Rosen [
1] to new technological applications in computation, communications, and radar [
2,
3,
4]. The representation of entanglement as a superposition of states
belonging to the direct product space
necessarily involves a precise and rigorous characterization of each Hilbert space
. Moreover, a covariant theory of relativistic entanglement requires detailed attention to the basis eigenstates of the commuting operators formed from the generators of a unitary representation of the Poincaré group.
In particular, the treatment of temporal entanglement must be performed in a fully covariant relativistic framework that places space and time on an equal footing in a consistent manner. This can be seen in the comparative analysis of experiments showing spatial and temporal interference: just as a quantum wavefunction split in space by a pair of slits produces spatial interference fringes, a wavefunction split in time can produce temporal interference fringes.
The demonstration by Davisson and Germer [
5] of the wavelike nature of particles in the double-slit experiment was a foundational contribution to quantum theory.
As shown schematically in
Figure 1, an electron passes through a slit at either
or
leading to a spatial superposition
producing interference fringes in space. In the temporal double-slit experiment shown schematically in
Figure 2, Lindner [
6] et. al. showed that an aperture opened and closed at temporal separation can produce interference in time. An ultra-short laser pulse ionizes an atom when
causing an electron to be emitted at either
or
and accelerated toward the detector. The resulting wavefunctions were shown to interfere coherently, producing temporal fringes.
In analogy to (
2), one might be tempted to describe this emission by a temporal superposition
showing an energy state producing interference fringes in time. Although wavefunction interference is superficially similar for space and time separations, the nonrelativistic states defined at different times in (
3) belong to different Hilbert spaces and cannot be said to form a coherent superposition. To see this, we simply recall that the time evolution of an unperturbed nonrelativistic state is determined by a Hamiltonian-generated unitary transformation
But expressing the proposed temporal superposition in the form
there is no obvious way to define the unitary evolution of this state or formulate a nonrelativistic many-body theory. Another temporal interference experiment has been proposed by Palacios et. al. [
7] involving a sequential double ionization of helium to produce a pair of entangled electrons in a singlet state. Horwitz and Arshansky [
8,
9] have discussed the conceptual difficulties associated with these experiments, and presented a rigorous relativistic analysis.
The challenges associated with formulating a consistent relativistic quantum mechanics for a single particle and its application to the description of many-body states are well-known. These difficulties are often reviewed in motivating the transition to quantum field theory (QFT). At minimum, constructing a consistent theory of superposition of states for the experiments of Lindner and Palacios requires a relativistic Hilbert space that treats space and time on the same footing. By definition, this space consists of the coherent eigenstates of a complete set of operators in a given representation defined with respect to a shared continuous parameterization. These operators are formed from the generators of the Poincaré group, with states characterized with respect to an induced representation of spin through the subgroup of the covering group for . As seen above, the time coordinate cannot be used to define the coherent states or parameterize their unitary evolution.
In this paper, we discuss the construction of relativistic Hilbert spaces in a framework developed by Stueckelberg, Horwitz, and Piron (SHP) to overcome the most serious difficulties associated with covariant quantum mechanics. This approach is described at length in a series of books [
10,
11,
12,
13] and only more recent references will be given explicitly in the following. Here we provide an updated review of the SHP framework, while introducing a novel representation of spin as a property of extra dimensions, organized as follows.
Section 2 presents an short review of the SHP formalism for spinless particles, and briefly introduces a new approach to representation of spin as a characteristic of an extended phase space. The extension of the classical phase space to include extra dimensions is formalized in
Section 3 and the electromagnetic interaction in the extended space is explored.
Section 4 reviews the construction of unitary representations of the Poincaré group for quantum states, and presents the induced representation of spin in the extra dimensions. Although these developments follow the standard presentation given by Wigner, they are implemented in the extended space, along side usual 4D spacetime, and so lead to new interpretations. Quantum mechanics in the extended phase space is then developed in
Section 5. This section generalizes the presentation by Horwitz and Arshansky, leading to an electromagnetic Hamiltonian in the new dimensions, along with solutions. In
Section 6 we construct entangled bipartite states and specify the singlet state to a discussion of the Palacios experiment.
2. The Stueckelberg, Horwitz, Piron (SHP) Formalism
The Stueckelberg, Horwitz, Piron (SHP) formalism [
10,
11,
12,
13] is a framework for special and general relativity in the classical and quantum realms developed to overcome the difficulties associated with the problem of time. In describing an antiparticle as a particle moving backward in time, Stueckelberg introduced an external parameter
, similar to the Newtonian time in nonrelativistic physics, so that the spacetime event
can change direction in coordinate time
while
advances monotonically. In order to permit the event velocity
to evolve continuously through the spacelike region between the forward and reverse lightlike regions, the 8D phase space
must be unconstrained, so that the usual constraint
is reduced to the status of a conservation law that applies under appropriate interactions. Unlike the convention in SHP theory, we take the flat metric to be
for consistency with familiar descriptions of spin. Horwitz and Piron extended the Stueckelberg framework into a fully covariant canonical theory with many-body interactions, whose applications include solutions to relativistic bound state and scattering problems under central forces, a microdynamical theory of electrodynamics, and more recently, motion in curved spacetime and evolution of the gravitational field [
14,
15,
16].
A free classical event in SHP theory is described by the Lagrangian and equivalent Hamiltonian
where
, with Euler-Lagrange and Hamilton equations
to which interactions may be added in much the same manner as in nonrelativistic theory. The Stueckelberg-Schrodinger equation
describes a single quantum mechanical event, and may also describe the reduced motion of a pair of interacting events after separation of the free center of mass motion. Solutions to this equation undergo unitary evolution in
as
By generalizing the central force problem for particles of spacelike separation
as
Horwitz and Arshansky found relativistic generalizations of the classical bound state solutions. States with the correct spectrum and multiplicity are obtained by choosing an arbitrary spacelike vector
and solving (
8) in a coordinate parameterization that guarantees
and
, where
. These solutions lower the spectrum by breaking the
symmetry of the original system to
. As a result, wavefunctions that transform under the full Lorentz group must be constructed as an induced representation with respect to
, leading to generators of
in a form that includes
and
. The Lie algebra for these transformations was found and new wavefunctions were then obtained as eigenstates of the commuting operators formed from these generators, providing an induced representation of
. Although the value of
was chosen arbitrarily, Land and Horwitz showed that it shifts under radiative transitions. By treating
as a dynamical quantity and extending the classical phase space to include
, the gauge invariant Hamiltonian acquires an interaction term in which the electromagnetic field is coupled in a natural way to the generators of the induced
. Thus, the Zeeman and Stark effects were found by Land and Horwitz in manifestly covariant form as a first order perturbation, reproducing the expected lifting of energy degeneracy.
Equations (
6) to (
8) describe a spinless particle. In standard relativistic quantum theory [
17] massive particles may be characterized in their rest frame, where
. Spin is introduced, following Wigner, by studying the unitary representations of the Poincaré group, involving an induced representation of
with respect to the momentum
. But in the SHP formalism
is unconstrained and may vanish or be negative for real particles, precluding the identification of a rest frame. Therefore, Horwitz, Piron, and Reuse modified Wigner’s method by specifying a position state and inducing a representation of
with respect to an arbitrarily chosen timelike unit vector designated
. The resulting spin operators act in the spacelike hypersurface normal to
, characterizing the chosen
subgroup of
. Because a singlet state belongs to an irreducible representation of the direct product of identical spin-1/2 states in the same representation, the entanglement requires that the individual wavefunctions be defined at the same value of
and with the same value of
.
In this paper, we will treat as a dynamical quantity, as was done for the induced representation of in the analysis of the Zeeman and Stark effects for relativistic bound states. More specifically, we introduce an auxiliary momentum conjugate to a position and study the classical and quantum dynamics on the extended phase space . Applying Wigner’s method, we construct a unitary representation of the Poincaré group for quantum states, leading to a representation of spin induced on rather than . Since a superposition can only be formed from coherent states in the same Hilbert space, characterized by a representation of spin, a singlet state formed from pair of free particles must have identical values of , even as this extended momentum evolves dynamically. In the following, we explore the implications of this requirement.
3. Classical Extended Phase Space
We write the classical phase space
and introduce the fields
and
to write the classical Lagrangian
for which the action is invariant under the gauge transformations
Variation with respect to
and
leads to the Lorentz force
where
are the field strengths.
We may define the canonical momenta
and perform a Legendre transformation on the classical Lagrangian (
12) to obtain the classical gauge invariant scalar Hamiltonian
In
Section 5 we will see how the quantum Hamiltonian acquires a spin interaction term. As discussed in
Section 2, a pair of quantum particles must have the same value of
in order to enter a singlet state. To get a sense of how this quantity behaves in an external field, we consider two classical particles with initial conditions
and write the Hamilton equations
In the case that
and
we have
and note that by (
16),
in this case. Since the forces and initial conditions are the same for each particle, they will evolve without divergence in their values of
. We note that transforming the interaction terms in the Lagrangian (
12) as
and introducing kinetic terms for the fields, we may write a field Lagrangian as
Variation of this Lagrangian with respect to the potentials and leads to Maxwell-like field equations for which the sources are and , suggesting that the field solutions will be of the type and
Nevertheless, if either potential depends on both
x and
then we have
so that the forces on
and
may differ at spacetime separated locations
and
, leading to a relative change in their values. In the quantum context, divergent evolution in
may indicate a mechanism of decoherence.
4. Unitary Representation of the Poincaré Group
Basis quantities for quantum states are found as eigenstates of the commuting operators formed from the generators of a unitary representation of the Poincaré group
for
and translation
a. The spacetime Lorentz transformations
and translations
have generators
from which one may write unitary transformations
where
and
satisfy the well-known Lie algebra
leading to spinor representations.
We treat the extra dimensions as an inequivalent representation of the Poincaré group, and so in the extended phase space, the transformations of
and
are generated by
defined in analogy to (
28) but independent of
and
. With these, we may consider extended unitary transformations of the quantum states of the type
whose generators satisfy the Poincaré algebra. Although the spacetime 4-momentum is unconstrained and
is not necessarily non-negative definite, we now stipulate that the momentum
must satisfy the mass-shell constraint
.
4.1. Representations of the Lorentz Group
For later reference, we briefly review the spinor and vector representations of the Lorentz group, obtained from the combined generators
. The operators are partitioned into boost and rotation generators as
with commutation relations
decomposing the Lorentz transformation as
. Under space reflection
P,
and
which leads to the left and right handed operators
with commutation relations
The
and
generate two inequivalent representations of the
Lie algebra with Casimir operators
so that representations
are characterized as
. Since
, the spin of a given representation is given by
. For the two-component spinor
and
are represented by Pauli matrices, and (
36) leads to
which transform as
where
belong to the group
, and we have
. In particular, pure rotations and boosts are
For
the anti-Hermitian operator
acts as
and taking the complex conjugate of a left handed Lorentz transformation
we see that
. Thus,
transforms as
, implementing the parity transformation
. Introducing
, the matrices
provide two inequivalent bases for
. Deploying the indices of
C as
the spinor with upper index is defined as
with
C playing the role of a metric. Under Lorentz transformation
where here
C plays the role of a similarity transformation between the equivalent representations
and
.
The vector representation of
by
is found by writing
which conserves
because
.
4.2. The Little Group and Wigner Operator
By imposing the mass-shell constraint on the extended momenta
we may follow Wigner’s method for representation of spin. The little group
consists of all Lorentz transformations
for which
preserves the momentum
. One may construct the little group for
by choosing a standard vector
with known
and defining the Wigner operator
generated by
such that
In the
representation this becomes
where we denote by
the
matrix associated with the Wigner operator
. For any
associated with
we may find the corresponding element of
and by inversion we may find the little group
from a known little group
as
It follows that every Lorentz transformation
can be expressed as a combination of a little group element
and the Wigner operator
in the form
establishing the isomorphism
.
For the discussion of spin, it is convenient to choose the standard timelike
vector as pure time
and so we write
making explicit the connection with the arbitrary timelike unit vector in the work of Horwitz et. al. This leads to the requirement
so that
is the group of spatial rotations represented by
and a general
is the
rotation group in the spacelike hypersurface normal to
. An explicit form of the standard transformation
is the pure boost
so that
where
is the velocity parameter.
4.3. Basis Quantities for States with Spin
Horwitz et. al. defined quantum states in the position representation as eigenstates of
, but here we work in the extended momentum representation and define eigenstates of the operators
and
. With the notation
we introduce the Pauli–Lubanski pseudovector in the extended sector
where
indicate the orthogonality of
to
. The scalar
commutes with the other operators and in the special frame
takes the form
which represents the total spin as the Casimir invariant of the rotation subgroup in the hypersurface normal to
.
Writing
for the diagonal spin component, while suppressing the total spin
J and any additional eigenvalues associated with internal symmetries, we expect a Lorentz transformation to act on the momentum eigenvalue as
where
is a unitary representation of
. This is demonstrated by expressing the transformed operator
as
so that
The wavefunction for a spinless momentum state is
so that frame covariance requires
The Lorentz transformed momentum state is
and so
confirming that
The matrix element of
for the spinless state is thus
where the delta functions represent the action of
on the left, taking
to
and
to
, while the last equality follows from
.
For states with spin the identity operator is
and so the transformation acts as
with the matrix element
where
is a discrete matrix representation of the action of the Lorentz transformation on the spin state. Thus
expresses the action of the unitary Lorentz transformation. Unitarity requires
and using the matrix element this becomes
showing that
is a unitary matrix.
The Wigner operator
is a pure boost constructed from the operators
and so does not act on
p or
. The unitary representation
is defined such that
Inserting this definition into (
85) and multiplying both sides by
leads to
so that writing
, combining unitary matrices on the LHS, and using (
56) this becomes
The matrix element for
is thus
showing that
belongs to the little group
. This can be written as
where
A is the
representation of
and
is the Wigner matrix representation of the rotation
Since this combination is an element of
and a rotation, it belongs to the subgroup
and is therefore unitary, and we have simply
Together, the matrix element
in (
84) is a combination of a pure boost with a rotation of the spin indices in the hypersurface normal to
.
In wavefunction notation with
as defined in (
92) the transformation can be written
and multiplying both sides by
this becomes
showing that
undergoes Lorentz transform as
. As we saw in
Section 4.1 there is an inequivalent representation that transforms under the complex conjugate representation as
, so together
are the fundamental spinorial bases for the
representation of the Lorentz group.
5. Quantum Mechanics in the Extended Phase Space
5.1. Bispinors
Using the basis states defined in (
97) we construct a bispinor in the Weyl basis as
leading to the Dirac spinor
in which the upper and lower sectors have even and odd parity, respectively. This field must transform as
where
mixes components of the bispinor and acts on the standard gamma matrices as
in analogy to the
transformation. Comparing the forms of the transformation
leads to the condition
which is satisfied by
It is convenient here to adopt the Hestenes approach to Clifford algebra [
18] and treat
as the 4D basis vectors. Thus, we write vectors and bivectors as
with the products
and so
The explicit matrix form of
is
so that only
is diagonal. The four bivectors
are defined as
and take the explicit form
Writing the transverse projection of the basis vectors as
Thus, and each have 3 independent components, satisfying the Lie algebra and generating boosts and rotations in the spacelike hypersurface transverse to .
5.2. Quantum Hamiltonian
Using
, the projections of spacetime momentum can be expressed as
and the parity transformation is given by
Generalizing the work of Horwitz and Arshansky to the extended phase space, we write the longitudinal and transverse parts of the momenta in the form
Using
we define
so that the four operators
may be treated as Hermitian with respect to the standard
matrices. Using
we have
so that
leading to the expression
as the free particle quantum Hamiltonian. With the minimal gauge substitution
we expand the longitudinal components of the Hamiltonian as
and
where the wedge products all vanish because the longitudinal components are colinear and so these terms are proportional to
. The transverse components are
where the interaction terms act on a state as
Using
and the commutation relations
we are left with
and similarly
where again
Putting together the longitudinal and transverse components we find the electromagnetic Hamiltonian as
which is equivalent to the classical spinless Hamiltonian (
19) with the addition of the spin interaction terms. As in
Section 3, terms containing
do not directly appear in the Hamiltonian or couple to the spin operator
.
5.3. Plane wave Solutions
Taking
in the Hamiltonian (
138), the free particle Stueckelberg-Schrodinger equation
admits the plane wave solution
where
is a constant amplitude. In the special frame
we can write the four independent amplitudes as
where
is some normalization and
The general plane wave solution (
143) is found by boosting
to
and transforming the state as
where the phase of the plane wave is a Lorentz invariant and
We can now write the four independent solutions as
where
so that using
we find the transformed amplitudes as
Using (
49) the conjugate bispinor is
where
so that
leading to
5.4. The Spin Operator
For the bispinor representation, the Pauli–Lubanski pseudovector (
67) takes the form of the four matrices
where the second equality follows from (
112). The scalar product
is independent of
and commutes with all the other generators. By direct calculation we find the explicit forms
with the 3D inner product
and cross product
, from which we obtain
describing the bispinor as a spin-1/2 state. In the special frame
we have
so that acting on the plane wave solutions,
has eigenvalue
on
and
, and eigenvalue
on
and
.
Since
is orthogonal to
it has three independent components in the spacelike hypersurface, for which we may find an orthonormal basis
for
. Expanding
where
is the
component of
in the
basis. From (
164) we see that the components
satisfy
as expected for the 3-vector spin operator
.
For simplicity we may consider a boost along one of the three space axes
a so that
leading to
Taking the boost along the 3-axis this becomes
so that
and
are diagonal while
and
are off-diagonal.
Using the Gram-Schmidt method we can define the orthonormal basis
starting with
and leading to
so that although
has a 0-component, it is spacelike in the 4D spacetime and normal to
n. Using
we find
so that
and
which is diagonal and independent of
n. Using the
n-dependent bispinors (
152) and (
153) we see
as expected.
6. Entanglement
6.1. A Singlet State
We write the plane wave state (
149) as
where
describes a spinless plane wave on the extended spacetime. A two-body plane wave state is the direct product
and so the singlet state is
where the spacetime part
is symmetric under
while the spin part
requires
and is antisymmetric under
. Since the one-particle states must transform under the same representation of
, we may form singlets from the pairs
and
.
6.2. The Palacios Experiment
The experiment proposed by Palacios et. al. involves the sequential double ionization of helium, producing a pair of entangled electrons with a small time separation. The electrons are emitted with energies and separated in time by (femtoseconds) and each with an emission pulse width of . Since the SHP formalism treats time/space and energy/momentum on an equal footing, the uncertainly relation holds rigorously, and in this case leads to which is seen to apply for this process. The electrons may thus be considered identical particles and may form a singlet state.
Describing the singlet using (
183) — (
185) the spin part describes the anti-symmetric combination of spin up and spin down states, where the states must be chosen with the same parity. In the special frame
the spin factor is either
or
and the spacetime part in the extended phase space is the symmetric combination
where taking
on-shell for the free particles, (
181) reduces to
Now (
189) is just
and defining
this simplifies to
which is compatible with the expression obtained by Palacios et. al. and indicates interference fringes over time
.
7. Summary
The phenomenon of quantum entanglement is an increasingly important topic in contemporary physics, and new developments in this area will depend on determining appropriate frameworks for its representation. The description of entanglement as a superposition of distinguishable states necessarily involves a precise and rigorous characterization of the Hilbert space of each constituent particle. Although nonrelativistic methods are adequate in many applications, relativistic considerations become relevant at high energy, short distance, and as we have seen here, for describing temporal interference. Therefore, a proper treatment of relativistic states requires a fully covariant framework that places space and time on an equal footing in a natural and consistent manner. Moreover, the covariant Hilbert space must include the coherent eigenstates of the commuting operators found from the Poincaré algebra, and be defined with respect to a shared continuous parameterization left invariant by these operators.
As we have seen, the Stueckelberg-Horwitz-Piron (SHP) formalism satisfies these requirements by introducing a Poincaré invariant evolution parameter and relaxing the mass-shell constraint on spacetime momenta. But since is unconstrained and has no definitive rest-frame value, Wigner’s construction of spin states in the spacelike hypersurface normal to cannot be carried out. Horwitz et. al. circumvented this problem by introducing an arbitrary timelike unit vector , defining spin states in the spacelike hypersurface normal to , and inducing the representation on the orbit of these arbitrarily defined states. Because the possible spin states are defined with respect to a representation of determined by the chosen value of , coherent states and their superpositions must share this choice.
In this paper, we have attempted to provide a dynamical interpretation to by associating it with the momentum in an extended phase space. By constructing a classical Lagrangian and Hamiltonian gauge theory on the extended phase space and considering the classical equations of motion in the extended electromagnetic field, we identified the conditions under which the values of can diverge for a particle pair, indicating a possible mode of disruption of a quantum singlet. Defining the generators of Poincaré transformation on the extended phase space, we developed the spinor representations of the Lorentz group. To introduce spin states, we constructed the little group and Wigner operator acting on the momenta in the extended sector, leading to a Pauli-Lubanski operator and pure spin operator in this sector. The matrix elements for a Lorentz transformation on these states consists of a pure boost in the spacetime and extended sectors and a rotation of the spin indices in the spacelike hypersurface of the extended sector orthogonal to . Projecting onto the longitudinal and transverse components of the momentum we found the electromagnetic quantum Hamiltonian, and wrote plane wave solutions on the extended phase space. In the absence of an external field, these free particle states evolve with constant and permit the construction of singlet states defined in the same representation of . Such states contain a spacetime part symmetric under exchange of particles, and an antisymmetric spin part. We finally presented an analysis of the proposed Palacios experiment, showing the appearance of interference fringe from a temporal superposition.
As discussed in
Section 2, the Zeeman and Stark effects were found by treating an arbitrary vector quantity required for the induced representation of
as a dynamical quantity. In a subsequent paper, we will discuss the transitions of singlet states discussed in Section (
Section 6) under perturbations induced by an external electromagnetic field. We anticipate that under certain conditions, such transitions may shift the respective values of
differentially, and thus place the constituent particles into different representations of spin. Such a process would imply perturbative decoherence and breaking of the singlet state.
Conflicts of Interest
The authors declare no conflicts of interest.
References
- Einstein, A.; Podolsky, B.; Rosen, N. Can Quantum-Mechanical Description of Physical Reality Be Considered Complete? Phys. Rev. 1935, 47, 777–780. [Google Scholar] [CrossRef]
- Caleffi, M.; Amoretti, M.; Ferrari, D.; Illiano, J.; Manzalini, A.; Cacciapuoti, A.S. Distributed quantum computing: A survey. Computer Networks 2024, 254, 110672. [Google Scholar] [CrossRef]
- Lederman, M.; Pereg, U. Secure Communication with Unreliable Entanglement Assistance. In Proceedings of the 2024 IEEE International Symposium on Information Theory (ISIT); 2024; pp. 1017–1022. [Google Scholar] [CrossRef]
- Zhuang, Q. Quantum advantage on the radar. Nature Physics 2023, 19, 1384–1385. [Google Scholar] [CrossRef]
- Davisson, C.; Germer, L.H. Diffraction of Electrons by a Crystal of Nickel. Phys. Rev. 1927, 30, 705–740. [Google Scholar] [CrossRef]
- Lindner, F.; Schätzel, M.G.; Walther, H.; Baltuška, A.; Goulielmakis, E.; Krausz, F.; Milošević, D.B.; Bauer, D.; Becker, W.; Paulus, G.G. Attosecond Double-Slit Experiment. Phys. Rev. Lett. 2005, 95, 040401. [Google Scholar] [CrossRef] [PubMed]
- Palacios, A.; Rescigno, T.N.; McCurdy, C.W. Two-Electron Time-Delay Interference in Atomic Double Ionization by Attosecond Pulses. Phys. Rev. Lett. 2009, 103, 253001. [Google Scholar] [CrossRef] [PubMed]
- Horwitz, L.; Arshansky, R.I. Relativistic entanglement. Physics Letters A 2018, arXiv:quant-ph/1707.03294]382, 1701–1708. [Google Scholar] [CrossRef]
- Horwitz, L.P. Spin and entanglement in general relativity. The European Physical Journal Plus 2021, 136. [Google Scholar] [CrossRef]
- Horwitz, L.P. Relativistic Quantum Mechanics; Springer: Dordrecht, Netherlands, 2015. [Google Scholar] [CrossRef]
- Horwitz, L.P.; Arshansky, R.I. Relativistic Many-Body Theory and Statistical Mechanics; 2053-2571, Morgan & Claypool Publishers, 2018. [CrossRef]
- Land, M.; Horwitz, L.P. Relativistic classical mechanics and electrodynamics; Morgan and Claypool Publishers, 2020.
- Horwitz, L. Concepts in Relativistic Dynamics; World Scientific Publishing Co., 2023. Publisher Copyright: © 2023 by World Scientific Publishing Co. Pte. Ltd. All rights reserved. [CrossRef]
- Horwitz, L.P. An Elementary Canonical Classical and Quantum Dynamics for General Relativity. Journal of Physics: Conference Series 2019, 1239, 012014. [Google Scholar] [CrossRef]
- Horwitz, L.P. An elementary canonical classical and quantum dynamics for general relativity. The European Physical Journal Plus 2019, 134, 313. [Google Scholar] [CrossRef]
- Land, M. A vielbein formalism for SHP general relativity. Journal of Physics: Conference Series 2023, 2482, 012006. [Google Scholar] [CrossRef]
- Novozhilov, Y.V. Introduction to Elemenatary Particle Theory; Pergamon Press: Oxford, 1975. [Google Scholar]
- Hestenes, D. Space-time algebra; Documents on modern physics, Gordon and Breach, 1966.
|
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content. |
© 2024 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).