We propose a novel formulation for a manifestly Lorentz-covariant spinor wave-packet basis. The traditional definition of the spinor wave packet is problematic due to its unavoidable mixing with other wave packets under Lorentz transformations. Our approach resolves this inherent mixing issue. The wave packet we develop constitutes a complete set, enabling the expansion of a free spinor field while maintaining Lorentz covariance. Additionally, we present a Lorentz-invariant expression for zero-point energy.
∗ Department of Mathematics, Tokyo Woman’s Christian University, Tokyo 167-8585, Japan
†Department of Physics, University of Tokyo, Tokyo 113-0033, Japan
1 Introduction
In quantum mechanics, wave packets serve as a crucial conceptual and mathematical foundation. In real observations, we never encounter idealized plane-wave states, characterized by zero uncertainty in momentum and infinite uncertainty in position. They do not belong to the Hilbert space because of its non-normalizability.
In quantum field theory (QFT), the plane-wave -matrix is traditionally used, but this approach results in divergences due to the squared energy-momentum delta function. This makes the plane-wave -matrix computation more of a mnemonic than a rigorous derivation for observables; see e.g. Ref. [1].
Wave packet states have been extensively discussed in various contexts of particle physics phenomenology, including anomalies in vector meson decay [2], corrections to Fermi’s golden rule [3, 4], searches for dark photons [5], applications in quantum computation [6, 7, 8], and studies of neutrino oscillation [9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20].
Despite these applications, theoretical efforts to construct QFT based on wave-packet states have been limited. Previous work has employed Gaussian formalism, utilizing Gaussian wave functions as a complete basis for expanding the free one-particle Hilbert space. This approach has revealed phenomena like time-boundary effects, which are unobservable in plane-wave formalism [21]. However, the Gaussian formalism breaks manifest Lorentz covariance, necessitating a refinement for both aesthetic and pragmatic reasons.
The absence of manifest Lorentz covariance is unsatisfying, as historically, physics has advanced through symmetry-based formalisms, such as the Dirac [22] and Becchi-Rouet-Stora-Tyutin (BRST) [23, 24, 25] formalisms. Therefore, it is desirable to develop a manifestly Lorentz-covariant wave packet formalism.
Practically, the Gaussian formalism complicates calculations due to the lack of Lorentz covariance, evident in the difficulty of deriving explicit analytic formulas for Gaussian wave functions in position space. In contrast, our previous study demonstrated the expansion of scalar fields in QFT using a complete basis of Lorentz-invariant wave packets, facilitating the derivation of analytic formulas for Lorentz-invariant wave functions in scalar fields [26]. Also, the Gaussian wave packet is shown to be a non-relativistic limit of the Lorentz-invariant wave packet for the scalar fields [26].
In this study, we investigate the wave packet basis for spinors within a Lorentz-covariant framework. Traditional methodologies, including Gaussian formalism (refer to Appendix A in Ref. [27] for an overview) and other approaches towards Lorentz-covariant spinor wave packets [28, 29, 30, 31, 32], assume that the spin dependence of spinor wave packets aligns with that of plane waves. However, this assumption introduces a significant complication: it leads to a complex Lorentz transformation law, which unexpectedly intertwines wave-packet states with different central momenta and positions. Such a phenomenon is physically paradoxical. Imagine a scenario where a single particle is depicted by a wave packet with a distinct central momentum and position. The blending of this wave packet with others having varying central momenta and positions effectively results in an unphysical merging of distinct particles under Lorentz transformation, which is a challenging notion for conventional theories.
In response to this issue, this paper introduces a novel definition of the spinor wave packet that circumvents this problematic mixing: , where and are its position and momentum centers, its spin, and its spatial width-squared. The wave packet is defined by
(1)
where is a Lorentz-friendly plane-wave basis state with the momentum and spin ;
further details will be given in subsequent sections, especially in Eq. (34).
Our definition ensures that wave packets remain separate and distinct under Lorentz transformations. We then establish the completeness of this newly defined spinor wave packet within the free one-particle subspace. Expanding on this concept, we demonstrate that the spinor field can be effectively expanded using our spinor wave packet. Additionally, we explore the application of this approach to several well-established operators in the wave packet basis, providing a comprehensive understanding of its implications in quantum field theory.
The paper is organized as follows: Section 2 introduces the new Lorentz-covariant spinor wave-packet basis in the one-particle subspace. Section 3 extends this to the creation and annihilation operators, showing how the free fermion field can be expanded using this basis. Finally, Section 4 presents the expression of several QFT operators in terms of wave packets.
2 Lorentz-covariant spinor wave packet
In this section, we point out that the known representation of a spinor wave packet suffers from mixing with other wave packets under Lorentz transformations, and propose a complete set of Lorentz-covariant spinor wave-packet basis without the difficulty of mixing.
We work in the -dimensional Minkowski space spanned by coordinate system , with spatial dimensions.
We take the almost-plus metric signature ; expressions in the opposite convention can be found in Appendix D.
We only consider a massive field, , and always take -momenta on-shell,
,
throughout this paper unless otherwise stated.
When an on-shell momentum appears in an argument of a function such as , we use both and -dimensional notations interchangeably: .
2.1 Spinor plane waves, revisited
To spell out our notation, we summarize basic known facts on the spinor plane waves.
A free Dirac field can be expanded by plane wave as follows,
(2)
where and are plane-wave solutions of the Dirac equation
(3)
with being the spin in the rest frame of each solution.
Throughout this paper, we suppress the spinor indices for , , , etc. when unnecessary.
These solutions satisfy the following completeness relations,
(4)
and their normalization is
(5)
where is the Dirac adjoint.111
We adopt the spinor notation in Ref. [1]: , where and is the unit matrix in the spinor space. Here, is distinguished from the operator .
The coefficients and in Eq. (2) are the annihilation and creation operators
for particle and anti-particle, respectively, that satisfy the following anticommutation relations:
others
(6)
Free one-particle subspaces of particle and anti-particle are spanned by the following plane-wave bases:
(7)
where and denote the particle and anti-particle of the species , respectively. The anticommutator (6) leads to the inner product:
(8)
where labels the particle and anti-particle.
The normalization (8) leads to the completeness relation (resolution of identity) in the free one-particle subspace of each :
(9)
The Lorentz transformation law of the plane wave reads
(10)
where is the spin- representation of the Winger rotation ; see Appendix B for details.
2.2 Lorentz-covariant spinor wave packet
In this subsection, we first briefly review basic facts on Lorentz-invariant scalar wave packets [33, 34], which is discussed in our previous work [26]. Next, we point out the difficulty in the conventional treatment of the spinor wave packet [28, 29, 30, 31]. Then, we propose a new definition of the spinor wave packet and show that we can avoid this difficulty in our expression.
2.2.1 Brief review of Lorentz-invariant scalar wave packet
For central position and momentum in -dimensions,
a Lorentz-invariant scalar wave packet is defined by [33, 34]:222See e.g. Refs. [30, 26] for reviews.
(11)
where denotes the phase space333
Here, includes the wave-packet central time .
Though is not an independent variable, we also include it for the convenience of writing its Lorentz transformation below.
(12)
and the normalization factor
(13)
provides , in which is the modified Bessel function of the second kind. Here and hereafter, we fix unless otherwise stated.
The wave function and the inner product are obtained as [33, 26]
(14)
(15)
where for any complex vector , we write , namely,
(16)
(17)
with
and .444
The abuse of notation is understood such that a vector-squared is distinguished from the second component of by the context.
We note that there is no branch-cut ambiguity for the square root as long as
[26].
With this state, the momentum expectation value and its (co)variance become [33]
(18)
(19)
where
(20)
In general, a matrix element of becomes
(21)
Let us consider a spacelike hyperplane in the space of central position ; see Appendix A.
One can write the completeness relation in the position-momentum phase space in a manifestly Lorentz-invariant fashion [26] (see also Ref. [33]):
(22)
where denotes the identity operator in the one-particle subspace and
the Lorentz-invariant phase-space volume element is given by
(23)
in which
(24)
is the Lorentz-covariant volume element.
We stress that is not summed nor integrated in the identity (22) and that the identity holds for any fixed .
Let us consider a “time-slice frame” of the central-position space in which becomes an equal-time hyperplane ,
(25)
where the “standard” Lorentz transformation is defined by , with denoting in any frame; note that by definition; see Appendix A for details.
On the constant- hyperplane , the Lorentz-invariant phase-space volume element reduces to the familiar form:
(26)
Note that in the non-relativistic limit .
2.2.2 Difficulty in spin-diagonal representation
In the literature [28, 29, 30, 31] a so to say spin-diagonal one-particle wave-packet state with a spin has been defined as
(27)
where is nothing but the scalar Lorentz-invariant wave packet (11).555
In the literature, the normalization and -dependence [26] have been omitted.
Its normalization becomes
(28)
where is given in Eq. (15).666
An inner product of the spin-diagonal wave-packet state and another state is understood as
.
We will never consider an inner product of the spin-diagonal wave-packet state and a phase-space-diagonal wave-packet state that appears below so that this notation will not cause confusion.
This leads to the following completeness relation in the one-particle subspace
(29)
generalizing the completeness relation of the scalar wave packet (22).
Once the wave-packet state is defined, its Lorentz transformation law is obtained as
(30)
where
(31)
and
(32)
We see that the spin-diagonal choice (27) leads to the complicated transformation law (30) mixing the wave-packet state with the others having various centers of momentum and position.
Below, we will show that we can indeed realize a physically reasonable transformation law, so to say the phase-space-diagonal representation, which evades the mixing with other states (30):
(33)
where is a yet unspecified representation function.
2.2.3 Phase-space-diagonal representation
Instead of the conventional choice (27), we propose to define
(34)
where the key element is
(35)
and is a normalization factor to be fixed below;
in the second step in Eq. (35), we used
, with the charge conjugation matrix () in our notation.
Note that .
Here and hereafter, for notational simplicity, we omit the label that distinguishes the particle and antiparticle unless otherwise stated.
The definition (35) leads to777
One can show it as
(36)
Then it follows that
(37)
The identity (37) results in888
This can be shown as
where we have used Eqs. (10) and (37) and then the unitarity (95) in the second equality.
(38)
As promised, we have realized the phase-space-diagonal representation (33).
Now we show that the normalization is realized by the choice
where we used Eq. (4) in the second line;
we used and in the last line, which is convenient for the antiparticle.
The expectation value is presented in Eq (18), from which we get
(41)
Therefore, using the Dirac equation (3) and then the normalization (5), we see that the choice (39) provides the normalized state.
Finally, the inner product is given by the same procedure:
(42)
where, ; see Eqs. (15) and (21).
Hereafter, we adopt this representation for the Lorentz-covariant spinor wave packet.
2.3 Momentum expectation value
In this subsection, we compute the momentum expectation value of the Lorentz covariant spinor wave packet:
(43)
This will be an important parameter in the following.
Putting Eq. (34), we obtain
(44)
where we used Eq. (4). The expectation value and its covariance , are shown in Eqs (18) and (19). Thus,
(45)
Hence, using the Dirac equation (3) and then the normalization (5), we get
(46)
Therefore, the momentum expectation value is given by
(47)
where
(48)
Note that in the non-relativistic limit .
2.4 Completeness
In this subsection, we will prove the following completeness relation for Lorentz-covariant spinor wave packet,
(49)
where
(50)
in which , , and are given in Eqs. (23), (20), and (48) respectivity.
To prove Eq. (49), we rewrite it as a matrix element for both-hand sides, sandwiched by the plane-wave bases (7):
(51)
where we used Eq. (8) on the right-hand side.
On the left-hand side, we integrate over by exploiting its Lorentz invariance, choosing a coordinate system where it becomes a constant- hyperplane with . Then left-hand side in Eq. (51) becomes
(52)
where we used Eq. (4) in the first line, and Eqs. (46) and (48) in the last line. Thus, Eq. (51), and hence the completeness (49), is proven.
3 Spinor field expanded by wave packets
Now we define the creation and annihilation operators of the Lorentz-covariant wave packet.
We write a free spin- one-particle state of th spinor particle and of its anti-particle .
Similarly to the plane wave case, we define wave-packet creation operators by
(53)
(54)
and annihilation operators by their Hermitian conjugate, with mass dimensions , etc.
Then, the completeness relation (49) on the one-particle subspace reads
(55)
and similarly for the anti-particles.
Then, we can naturally generalize it to an operator relation that is valid on the whole Fock space:
(56)
(57)
Similarly, the completeness of the plane wave (9) leads to the expansion of these creation and annihilation operators:
(58)
(59)
From the above equations, we can derive the anti-commutation relation of the creation and annihilation operators:
(60)
others
(61)
where denotes the identity operator in the whole Fock space, and is the inner product of the Lorentz covariant wave packets, given in Eq. (42).
Finally, the free spinor field can be expanded as
(62)
where the Dirac spinor wave functions are given by
(63)
(64)
in which we have used the scalar wave function (14). Here, and are given in Eq. (16) and below it, respectively.
The normalization conditions of these Dirac spinors are
(65)
where we used Eqs. (5) and (9). The normalization is as same as the case of plane waves (5), except for the integration of .
Next, the completeness relations can be computed by
(66)
where we have used Eqs. (4), (41) and (45). These relations are similar to that of plane waves (4), except for the integration of and factor on the right-hand side.
4 Energy, momentum, and charge
In this section, we rewrite well-known operators in QFT, i.e. the total Hamiltonian, momentum, and charge operators, in the language of the spinor wave packet. Since the wave packet is not the momentum eigenstate, the total Hamiltonian and momentum operators cannot be diagonalized in the wave packet basis. However, the zero-point energy can be described in a fully Lorentz invariant manner using this basis. In Appendix C, we also show the corresponding expressions for the scalar wave packet.
First, let us consider the convergent part of the total Hamiltonian and momentum operators. In the momentum space, these operators are given by
(67)
Putting Eqs. (56) and (57) into the above expression, we get
(68)
where
(69)
We see that the total Hamiltonian and momentum operators are not diagonal on the wave packet basis, unlike the plane-wave eigenbasis.
Let us discuss the divergent part of this operator, coming from the zero-point energy:
(70)
Similarly as above, putting Eq. (57) into this commutator, we obtain
(71)
where we have used the completeness relation (49) in the second line and, in the last line, the expectation value (47) and the Lorentz-invariant phase-space volume element (2.4).
Let the time-like normal vector and spacelike vectors () compose an orthonormal basis: and such that we can decompose into the components parallel and perpendicular to ,
(72)
When we put this into Eq. (71), the perpendicular components vanish in a regularization scheme that makes Lorentz covariance manifest, namely in the dimensional regularization:
(73)
Therefore, the divergent part has only one independent component, which can be interpreted as zero-point energy, defined in a manifestly Lorentz-invariant fashion:
(74)
where is the coefficient in front of in the right-hand side of Eq. (71).
We note that the zero-point energy should be a scalar as we have shown, otherwise, an infinite momentum would appear from a Lorentz transformation.
Physically, we would expect that the zero-point energy is independent of the choice of spacelike hyperplane .
We can show it by exploiting the Lorentz invariance of the expression (74) by choosing (), without loss of generality. Then, the zero-point energy reduces to the well-known form:
(75)
It is remarkable that this zero point energy of the Dirac spinor is exactly times that of a real scalar, shown in Eq. (112) in Appendix, although the expression of momentum expectation values in Eqs. (47) and (18) are completely different between the spinor and scalar.
The factor is the number of degrees of freedom, and the negative sign cancels the bosonic contribution in a supersymmetric theory.
Next, we consider the following charge operator
(76)
Substituting Eq. (56) into the above expression, we obtain
(77)
where we have used
(78)
which follows from Eq. (49). This expression Eq. (77) means that the creation operators and create the wave packet with charge , and respectively. In fact,
(79)
is valid. Here we have used Eq. (78) in the last line.
5 Summary and discussion
In this paper, we have proposed fully Lorentz-covariant wave packets with spin. In the conventional definition of the wave packet, spin dependence of the wave function in the momentum space is just given by Kronecker delta, , and such a wave packet with spin transforms under Lorentz transformation mixing wave-packet states that have different centers of momentum and position. Our proposal overcomes this difficulty.
We have also proven that these wave packets form a complete basis that spans the spinor one-particle subspace in the manifestly Lorentz-invariant fashion.
Generalizing this completeness relation to the whole Fock space, we have shown that the creation and annihilation operators of plane waves can be expanded by that of these wave packets. This relation leads to the expansion of the spinor field in a Lorentz covariant manner. In addition to this, we have expressed the well-known operators in a wave packet basis: the total Hamiltonian, momentum, and charge operators. In particular, we have given the Lorentz covariant expression of zero point energy, in terms of centers of momentum and position of this wave packet.
In the following, we will comment on several future directions. First, as mentioned in the Introduction, the novel Lorentz-covariant basis we propose will be useful in handling wave packet quantum field theory [27, 35, 21, 36], which has previously relied on the saddle-point approximation for computing position-momentum integrals. For future work, we aim to go beyond the leading-order computation and properly account for the effect of wave packet spreading. To achieve these goals, we require an analytical approach without approximations, which is feasible within our formulation.
Second, neutrino oscillation inherently requires a wave-packet formulation, and our newly defined wave packet could significantly impact this field. Previous analyses have primarily used Gaussian wave packets, without considering the spinorial structure—a key focus of this paper. For instance, Giunti’s seminal work [9] addresses neutrino decoherence in quantum mechanics. Akhmedov and Smirnov have also extensively studied neutrino wave packets [13, 19], exploring phenomenological aspects and scattering processes. Many other groups have contributed to this field as well [10, 11, 12, 28, 14, 15, 29, 16, 17, 30, 18, 31, 32, 20]. However, these studies do not fully address the spin aspects of neutrinos. Our proposed spinor wave packet framework, which is Lorentz-covariant, may yield novel predictions for neutrino mixing phenomenology and provide a deeper understanding by incorporating the spinorial structure into the wave-packet formulation.
It may also be interesting to consider Bell’s inequality in the context of this wave packet state. Another possible application might involve vortex collisions; see, e.g., Ref. [37] for a review.
Acknowledgement
We thank Ryusuke Jinno for a useful comment.
This work is supported in part by the JSPS KAKENHI Grant Nos. 19H01899, 21H01107 (K.O.), and 22KJ1050 (J.W.).
Appendix
Appendix A “Slanted” foliation
In this appendix, we briefly introduce “slanted” foliation which is necessary to write down the completeness relation of Lorentz-invariant wave packets in the fully Lorentz-invariant manner.
Let us consider the following spacelike hyperplane:
(80)
where is an arbitrary fixed vector that is timelike-normal and is future-oriented , namely ,
and parametrizes the foliation.
Physically, is the normal vector to the hyperplanes and is the proper time for this foliation.
A schematic figure is given in the left panel in Fig. 1. We can generalize the equal-time foliation of whole Minkowski space to a general foliation by set of these spacelike hyperplanes.
Figure 1: in frame (left) and in frame (right)
In general, we may parametrize a component of in the reference frame as the following linear combination:
(81)
where the “standard vector” is defined to be
(82)
in any frame999
In the language of differential geometry, the basis-independent vector is written as with being the basis vectors in the reference coordinates.
Under the change of basis , where , should remain the same , that is, .
and is the “standard boost to the foliation.”
Concretely, for the vector with ,
(83)
where t denotes a transpose, I is the identity matrix in dimensions,
is given in the matrix representation, and
.
Note that .
Now an equal-time hyperplane in the arbitrary reference frame is written as because on it.
For any given foliation , we may Lorentz-transform from the reference frame to the “time-slice” frame that gives :
(84)
(85)
where we used as usual.
In the time-slice coordinate system , the same plane is written as
(86)
As said above, since on , they are equal-time hyperplanes parametrized by in the coordinate system.
A schematic figure is given in the right panel in Fig. 1.
Appendix B Wigner representation
In this appendix, we briefly review the Wigner representation in the case of massive one-particle state to spell out our notation; see e.g. Ref. [1] for more details.
Here and hereafter, we neglect the label for the particle and anti-particle since it is irrelevant for the current discussion.
The Poincaré transformation on a plane-wave state can be written as
(87)
where
(88)
in which is the generator of the spacetime translation.
Since the translational part is the same as the scalar case, we concentrate on the Lorentz transformation.
Without loss of generality, we can choose to be the of the particle in its rest frame:
(89)
consistently with the definition (8) as we will see below.
Here, is the spin eigenvalue for the rotation in, say, - plane in the rest frame and the standard boost is defined by
(90)
in which is given in Eq. (82).
Concretely, the standard boost to can be written in terms of the “standard boost to a foliation” (83) as101010
In general, these two are different concepts, , since and are different.
(91)
Since has an internal degree of freedom , the Lorentz group representation for this state could be nontrivial. To deal with this, we introduce the well-known procedure, Wigner representation.
First, under the Lorentz transformation, the plane-wave basis transforms as
(92)
where
(93)
Here, is corresponding to the rotation because this transformation does not change the momentum . We call this Wigner rotation in .111111
We sloppily write when it is to be understood as Spin.
Next, we may always write
(94)
where is a finite-dimensional unitary representation of :
(95)
Putting Eq. (94) into Eq. (92), we obtain Eq. (10).
Appendix C Energy, momentum, and number operator in scalar case
We give the energy, momentum, and number operators for a real scalar field in terms of the Lorentz-invariant wave-packet basis. First, we briefly review the scalar wave packet in QFT, and then we will show newly-found expressions of these operators in terms of the Lorentz-invariant scalar wave packets.
The free field is usually expressed in the plane wave basis:
(96)
where and are the creation and annihilation operators of the plane waves, which satisfy , etc.
Now, we define a wave-packet creation operator by [26]
(97)
and an annihilation operator by its Hermitian conjugate.
The completeness of the scalar wave packet (22) leads to the following expansion of the creation and annihilation operators of the plane waves:
(98)
Thus, the free scalar field can be expanded as [26]
Now let us rewrite well-known operators in QFT, i.e. the total Hamiltonian, momentum, and number operators, into the language of the scalar wave packet.
First, in momentum space, the convergent part of the number operator is described by
(100)
Substituting Eq. (98) into the above expression, we obtain
(101)
On the second line, we have used
(102)
which follows from Eq. (22). From Eq. (101), we can read off a Lorentz-covariant number-density operator in the -dimensional phase space:
(103)
We now consider the divergent part of the plane-wave number operator, coming from the zero-point oscillation:
(104)
Putting Eq. (98) into the above expression, we obtain
(105)
Therefore, including the divergent part, the number-density operator can be described by
(106)
From this expression, it can be interpreted that there is one zero-point oscillation per -dimensional phase space volume.
Next, we consider the convergent part of the total Hamiltonian and momentum operators. In the momentum space, these operators are given by
(107)
Putting Eq. (98) into the above expression, we get
(108)
where is given in Eq.(21).
We see that the total Hamiltonian and momentum operators are not diagonal on the wave packet basis, unlike the plane-wave eigenbasis.
Now, let us discuss the divergent part of this operator, coming from the zero-point energy:
(109)
Putting Eq. (98) into the above commutator, we obtain
(110)
where we have used the completeness relation (22) in the second line,
the formula (18) in the third line,
and
the same argument as in Eq. (73) in the last line.
It is noteworthy that the result becomes the same as in the spinor case (71) up to the factor .
We may define the zero-point energy in a manifestly Lorentz-invariant fashion:
(111)
where is the coefficient of in the right-hand side of Eq. (110).
Physically, we expect that the zero-point energy should be independent of the choice of the spacelike hyperplane .
We can show it by exploiting the Lorentz invariance of the expression (111) to choose (), without loss of generality. Then, the zero-point energy reduces to the well-known form:
(112)
Appendix D Conversion of metric convention
We list the corresponding equations in the almost-minus metric signature to those in the main text.
In this paragraph only, we tentatively put the subscript “Main” and “AppD” on those in the main text and Appdendix D, respectively.
The metric sign is reverted: , and ; see footnote 1.
The gamma matrices are related by . The Clifford algebra remains the same: .
[1]
S. Weinberg, The Quantum theory of fields. Vol. 1: Foundations.
Cambridge University Press, 2005.
[2]
K. Ishikawa, O. Jinnouchi, K. Nishiwaki, and K.-y. Oda, Wave-packet
effects: a solution for isospin anomalies in vector-meson decay, Eur.
Phys. J. C83 (2023), no. 10 978,
[arXiv:2308.09933].
[3]
K. Ishikawa, O. Jinnouchi, A. Kubota, T. Sloan, T. H. Tatsuishi, and
R. Ushioda, On experimental confirmation of the corrections to
Fermi’s golden rule, PTEP2019 (2019), no. 3
033B02, [arXiv:1901.03019].
[4]
R. Ushioda, O. Jinnouchi, K. Ishikawa, and T. Sloan, Search for the
correction term to the Fermi’s golden rule in positron annihilation, PTEP2020 (2020), no. 4 043C01,
[arXiv:1907.01264].
[5]
S. Demidov, S. Gninenko, and D. Gorbunov, Light hidden photon production
in high energy collisions, JHEP07 (2019) 162,
[arXiv:1812.02719].
[6]
S. P. Jordan, K. S. M. Lee, and J. Preskill, Quantum Computation of
Scattering in Scalar Quantum Field Theories, Quant. Inf. Comput.14 (2014) 1014–1080, [arXiv:1112.4833].
[7]
S. P. Jordan, K. S. M. Lee, and J. Preskill, Quantum Algorithms for
Quantum Field Theories, Science336 (2012) 1130–1133,
[arXiv:1111.3633].
[8]
Z. Davoudi, C.-C. Hsieh, and S. V. Kadam, Scattering wave packets of
hadrons in gauge theories: Preparation on a quantum computer,
arXiv:2402.00840.
[9]
C. Giunti and C. W. Kim, Coherence of neutrino oscillations in the wave
packet approach, Phys. Rev. D58 (1998) 017301,
[hep-ph/9711363].
[10]
C. Y. Cardall, Coherence of neutrino flavor mixing in quantum field
theory, Phys. Rev. D61 (2000) 073006,
[hep-ph/9909332].
[11]
E. K. Akhmedov, J. Kopp, and M. Lindner, Oscillations of Mossbauer
neutrinos, JHEP05 (2008) 005,
[arXiv:0802.2513].
[12]
J. Kopp, Mossbauer neutrinos in quantum mechanics and quantum field
theory, JHEP06 (2009) 049,
[arXiv:0904.4346].
[13]
E. K. Akhmedov and A. Y. Smirnov, Paradoxes of neutrino oscillations,
Phys. Atom. Nucl.72 (2009) 1363–1381,
[arXiv:0905.1903].
[14]
E. K. Akhmedov and A. Y. Smirnov, Neutrino oscillations: Entanglement,
energy-momentum conservation and QFT, Found. Phys.41 (2011)
1279–1306, [arXiv:1008.2077].
[15]
E. K. Akhmedov and J. Kopp, Neutrino Oscillations: Quantum Mechanics vs.
Quantum Field Theory, JHEP04 (2010) 008,
[arXiv:1001.4815]. [Erratum:
JHEP 10, 52 (2013)].
[16]
J. Wu, J. A. Hutasoit, D. Boyanovsky, and R. Holman, Neutrino
Oscillations, Entanglement and Coherence: A Quantum Field theory Study in
Real Time, Int. J. Mod. Phys. A26 (2011) 5261–5297,
[arXiv:1002.2649].
[17]
E. Akhmedov, D. Hernandez, and A. Smirnov, Neutrino production coherence
and oscillation experiments, JHEP04 (2012) 052,
[arXiv:1201.4128].
[18]
M. Blasone, S. De Siena, and C. Matrella, Wave packet approach to quantum
correlations in neutrino oscillations, Eur. Phys. J. C81
(2021), no. 7 660, [arXiv:2104.03166].
[19]
E. Akhmedov and A. Y. Smirnov, Damping of neutrino oscillations,
decoherence and the lengths of neutrino wave packets, JHEP11
(2022) 082, [arXiv:2208.03736].
[20]
H. Mitani and K.-y. Oda, Decoherence in neutrino oscillation between 3D
Gaussian wave packets, Phys. Lett. B846 (2023) 138218,
[arXiv:2307.12230].
[21]
K. Ishikawa, K. Nishiwaki, and K.-y. Oda, New effect in wave-packet
scattering of quantum fields, Phys. Rev. D108 (2023), no. 9
096013, [arXiv:2102.12032].
[22]
P. A. M. Dirac, The quantum theory of the electron, Proc. Roy.
Soc. Lond. A117 (1928) 610–624.
[23]
C. Becchi, A. Rouet, and R. Stora, The Abelian Higgs-Kibble Model.
Unitarity of the S Operator, Phys. Lett. B52 (1974) 344–346.
[24]
C. Becchi, A. Rouet, and R. Stora, Renormalization of Gauge Theories,
Annals Phys.98 (1976) 287–321.
[25]
I. V. Tyutin, Gauge Invariance in Field Theory and Statistical
Physics in Operator Formalism, arXiv:0812.0580. Lebedev Institute preprint No. 39 (1975).
[26]
K.-y. Oda and J. Wada, A complete set of Lorentz-invariant wave packets
and modified uncertainty relation, Eur. Phys. J. C81 (2021),
no. 8 751, [arXiv:2104.01798].
[27]
K. Ishikawa and K.-y. Oda, Particle decay in Gaussian wave-packet
formalism revisited, PTEP2018 (2018), no. 12 123B01,
[arXiv:1809.04285].
[28]
D. V. Naumov and V. A. Naumov, Relativistic wave packets in a field
theoretical approach to neutrino oscillations, Russ. Phys. J.53 (2010) 549–574.
[29]
D. V. Naumov and V. A. Naumov, A Diagrammatic treatment of neutrino
oscillations, J. Phys.G37 (2010) 105014,
[arXiv:1008.0306].
[30]
D. V. Naumov, On the Theory of Wave Packets, Phys. Part. Nucl.
Lett.10 (2013) 642–650, [arXiv:1309.1717].
[31]
D. Naumov and V. Naumov, Quantum Field Theory of Neutrino Oscillations,
Phys. Part. Nucl.51 (2020), no. 1 1–106.
[32]
V. A. Naumov and D. S. Shkirmanov, Virtual neutrino propagation at short
baselines, Eur. Phys. J. C82 (2022), no. 8 736,
[arXiv:2208.02621].
[33]
G. Kaiser, Phase Space Approach to Relativistic Quantum Mechanics. 1.
Coherent State Representation for Massive Scalar Particles, J. Math.
Phys.18 (1977) 952–959.
[34]
G. R. Kaiser, Phase Space Approach to Relativistic Quantum Mechanics. 2.
Geometrical Aspects, J. Math. Phys.19 (1978) 502–507.
[35]
K. Ishikawa, K. Nishiwaki, and K.-y. Oda, Scalar scattering amplitude in
the Gaussian wave-packet formalism, PTEP2020 (2020), no. 10
103B04, [arXiv:2006.14159].
[36]
A. Edery, Wave packets in QFT: Leading order width corrections to decay
rates and clock behavior under Lorentz boosts, Phys. Rev. D104 (2021), no. 12 125015, [arXiv:2106.13768].
[37]
I. P. Ivanov, Promises and challenges of high-energy vortex states
collisions, Prog. Part. Nucl. Phys.127 (2022) 103987,
[arXiv:2205.00412].