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

Generalized Continuity Equations for Schrödinger and Dirac Equations

A. Katsaris Department of Physics, University of Athens, GR-15784 Athens, Greece    P. A. Kalozoumis Department of Engineering and Informatics, Hellenic American University, 436 Amherst Street, Nashua, NH 03063, USA Materials Science Department, School of Natural Sciences, University of Patras, Patras 265 04, Greece    F. K. Diakonos fdiakono@phys.uoa.gr Department of Physics, University of Athens, GR-15784 Athens, Greece
Abstract

The concept of the generalized continuity equation (GCE) was recently introduced in [J. Phys. A: Math. and Theor. 52, 1552034 (2019)], and was derived in the context of NN independent Schrödinger systems. The GCE is induced by a symmetry transformation which mixes the states of these systems, even though the NN-system Lagrangian does not. As the NN-system Schrödinger Lagrangian is not invariant under such a transformation, the GCE will involve source terms which, under certain conditions vanish and lead to conserved currents. These conditions may hold globally or locally in a finite domain, leading to globally or locally conserved currents, respectively. In this work, we extend this idea to the case of arbitrary SU(N)SU(N)-transformations and we show that a similar GCE emerges for NN systems in the Dirac dynamics framework. The emerging GCEs and the conditions which lead to the attendant conservation laws provide a rich phenomenology and potential use for the preparation and control of fermionic states.

I Introduction

Symmetries and the emerging conservation laws are of fundamental importance in Physics. For continuous transformations, the link between conservation laws and symmetries is provided by Noether’s theorem Noether1918 . Based on the Lagrangian formulation of the underlying dynamics, Noether’s theorem allows for applications both in classical field theory and in non-relativistic quantum mechanics, leading to symmetry induced continuity equations directly linked to conservation laws. In the context of Schrödinger dynamics, these continuity equations usually refer to a single quantum state of the considered system and they are connected to internal symmetries of the Hilbert space.

Recently, it was shown that symmetry-induced bilocal continuity equations emerge in quantum Hermitian and non-Hermitian systems Spourdalakis2016 from symmetry transformations leaving the corresponding Lagrangian density invariant. They generalize the usual continuity equations, involving two distinct wave fields which obey different dynamics, when the non-Hermitian part in the potential term is non-zero. The corresponding two-field Lagrangian is invariant under dilatation and phase transformations of these fields. The theoretical framework developed in Spourdalakis2016 addresses the impact of global symmetry transformations on quantum Schrödinger states and indicates that the currents associated with the bilocal continuity equations act as correlators between the two fields. However, the existence of global symmetries often can be an idealized scenario. In realistic systems symmetries can be broken due to defects, impurities and boundary conditions. Moreover, the presence of a potential term depending only on space leads to stationary states, which violate the equal treatment between space and time, indicating the breaking of Lorentz invariance.

A case of particular interest is when a symmetry transformation, even though globally broken, appears to hold in restricted spatial domains. The existence of such local symmetries, being present to a broad range of photonic Kalozoumis2013b , acoustic Kalozoumis2015a and, quantum Kalozoumis2013a systems induce interesting spectral properties and exhibit a new class of non-local currents (NLC) which have been shown to characterize rigorously the symmetry breaking Kalozoumis2014a . The question is, if these currents, which are connected to local symmetries, can be extracted variationally from a suitable Lagrangian density. Clearly, the GCE derived in Spourdalakis2016 is not directly transferable to a Lagrangian density possessing local symmetries, and a more general mathematical framework is needed for the description of local conservation laws assigned to systems with local symmetries. Such an approach has been recently proposed in Diakonos2019 . The variational scheme developed there, utilizing the concept of a super-Lagrangian, allowed to derive a GCE obeyed by the local symmetry induced NLC. This mathematical framework led to GCEs linking pairwise wave fields which belong to different Schrödinger problems and involve the dynamics of single particle evolution in different potentials Vi(𝐱)V_{i}(\mathbf{x}) with i=1,,Ni=1,...,N. Within this scenario, the extended symmetry transformation applied to the super-Lagrangian is chosen to belong to the SU(N)SU(N) Lie group and its algebraic form determines the form of the occurring source terms. As it has been shown Diakonos2019 , under certain conditions the source terms vanish and GCEs lead to generalized conservation laws. This consistent variational framework, allowed the description of a special class of currents, which are locally conserved in restricted spatial domains whenever symmetries of the potential terms are valid within these domains. From a group theoretical point of view, this analysis was restricted to transformations, which involved only generators of the SU(N)SU(N) Cartan sub-algebra and the super-Lagrangian was designed to describe only Schrödinger dynamics.

Although, the Schrödinger equation can handle a multitude of quantum mechanical problems, it is not suitable to describe microscopic systems at high energies, such as relativistic fermions. The need to incorporate special theory of relativity in a quantum mechanical context led to the Dirac equation, formulated by Paul Dirac in 1928. In this work, we extend the concept of the super-Lagrangian Diakonos2019 for non-relativistic Schrödinger problems, to a relativistic framework. We introduce GCEs which contain Dirac fermionic states and we consider SU(N)SU(N) symmetry transformations with generators beyond the Cartan sub-algebra, covering this way all the possible SU(N)SU(N) transformations. Thus, the set of possible source terms in the GCEs will be significantly increased, giving rise to a novel class of symmetry-induced invariant currents. Our results provide a unified treatment of generalized conservation laws in the presence of potentials with globally broken, domain-wise sustained symmetries for both Dirac and Schrödinger-type problems.

The paper is organized as follows: In Section II we introduce the super-Lagrangian for Dirac fermions, we derive the corresponding GCE, and extract the global and local conservation laws that emerge. In Section III we apply the concept of the super-Lagrangian to the case of two, single-Fermion Dirac systems and we derive the corresponding global and local conservation laws. In Section IV, non-relativistic systems are considered again and we generalize the results found in Diakonos2019 for SU(N)SU(N) symmetry transformations. Our results are summarized in Section V.

II Super-Lagrangian for Dirac Fermions and the Generalized Continuity Equation

The Dirac Lagrangian density i\mathcal{L}_{i} is,

i=ψ¯i(i∂̸Vi(x))ψi\mathcal{L}_{i}=\bar{\psi}_{i}(i\not{\partial}-V_{i}(\vec{x}))\psi_{i} (1)

where ψi\psi_{i} is a Dirac spinor. i\mathcal{L}_{i} describes the dynamics of fermionic states in the presence of the potential term Vi(x)V_{i}(\vec{x}). Applying the usual variational approach to i\mathcal{L}_{i} leads to the Dirac equation for the field ψi\psi_{i}. The super-Lagrangian is defined as the sum of the individual Lagrangian densities which describe the dynamics of independent quantum systems in different potential landscapes Vi(x)V_{i}(\vec{x}), i=1,,Ni=1,\dots,N. For these NN independent Dirac systems the super-Lagrangian can be written as,

=i=1Ni=Ψ¯(H0V)Ψ=0+~\mathcal{L}=\sum_{i=1}^{N}\mathcal{L}_{i}=\bar{\Psi}(H_{0}-V)\Psi=\mathcal{L}_{0}+\tilde{\mathcal{L}} (2)

where Ψ\Psi, Ψ¯\bar{\Psi} are defined as,

Ψ=(ψ1ψ2ψN)\Psi=\begin{pmatrix}\psi_{1}\\ \psi_{2}\\ \cdots\\ \psi_{N}\end{pmatrix} (3)

and

Ψ¯=Ψγ0=(ψ¯1ψ¯2ψ¯N)\bar{\Psi}=\Psi^{\dagger}\otimes\gamma^{0}=\left(\begin{array}[]{cccc}\bar{\psi}_{1}&\bar{\psi}_{2}&\cdots&\bar{\psi}_{N}\end{array}\right) (4)

respectively and ψi\psi_{i} are the corresponding Dirac spinors.

The quantities 0\mathcal{L}_{0} and ~\tilde{\mathcal{L}} are defined as,

0=Ψ¯H0Ψ;~=Ψ¯VΨ\mathcal{L}_{0}=\bar{\Psi}H_{0}\Psi~{}~{}~{};~{}~{}~{}\tilde{\mathcal{L}}=-\bar{\Psi}V\Psi (5)

with

H0=i∂̸𝟙NH_{0}=i\not{\partial}\otimes\mathds{1}_{N} (6)

and

V=diag(V1,V2,,VN),V=\text{diag}(V_{1},V_{2},\cdots,V_{N}), (7)

being the kinetic and potential terms, respectively.

To derive the GCEs from the super-Lagrangian in Eq. (2), we make an arbitrary transformation of the spinors Ψ\Psi. Note, that the source terms in the resulting GCEs are subject to this transformation. We are interested in the conditions under which these source terms vanish. In this case, we consider an SU(N)SU(N) transformation of Ψ\Psi, which for an infinitesimal variation δθa\delta\theta_{a} can be expressed as

UΨ=eiδθaTaΨ(1iδθaTa)ΨU\Psi=e^{-i\delta\theta_{a}T^{a}}\Psi\simeq(1-i\delta\theta_{a}T^{a})\Psi (8)

where TaT_{a} denotes the generators of the SU(N)SU(N) group and δθa\delta\theta_{a} are arbitrary small parameters. The variation of the Lagrangian density \mathcal{L} yields that δ0=0\delta\mathcal{L}_{0}=0, δ~0\delta\mathcal{\tilde{L}}\neq 0 and consequently δ=δ~0\delta\mathcal{L}=\delta\mathcal{\tilde{L}}\neq 0, with δ~=~~\delta\mathcal{\tilde{L}}=\mathcal{\tilde{L}}^{\prime}-\mathcal{\tilde{L}}. These considerations lead to a current conservation law with a vanishing 44-divergence.

If \mathcal{L} was invariant under the transformation UU then, according to Noether’s theorem, a typical continuity equation would be obtained, associated with this invariance. However, here \mathcal{L} is not invariant, since δ~0\delta\mathcal{\tilde{L}}\neq 0, and Noether’s approach leads to a GCE with non-vanishing source terms, dictated by the performed symmetry transform in Eq. (8). In order to calculate δ~\delta\mathcal{\tilde{L}} we employ the generators TaT_{a} of the SU(N)SU(N) group and we write the potential matrix VV as,

V(x)=V0(x)𝟙N+k=1N21Ck(x)TkV(x)=V_{0}(x)\mathds{1}_{N}+\sum_{k=1}^{N^{2}-1}C_{k}(x)T_{k} (9)

where

V0(x)=Tr(V𝟙N)N;Ck(x)=Tr(VTk).\ V_{0}(x)=\frac{Tr(V\cdot\mathds{1}_{N})}{N}~{}~{}~{};~{}~{}~{}C_{k}(x)=Tr(V\cdot T_{k}).

The explicit form of Ck(x)C_{k}(x) and the details for their derivation are presented in Appendix A. The fact that δ=~~\delta\mathcal{L}=\mathcal{\tilde{L}}^{\prime}-\mathcal{\tilde{L}} only, leads to

δ=Ψ¯(UVUV)Ψ.\delta\mathcal{L}=-\bar{\Psi}(U^{\dagger}VU-V)\Psi. (10)

Using Eq. (9) it is straightforward to find that

UVUV=δθaCkfakcTcU^{\dagger}VU-V=-\delta\theta_{a}C_{k}f_{ak}^{c}T_{c} (11)

and consequently

δ=δθafabcΨ¯CbTcΨ.\delta\mathcal{L}=\delta\theta^{a}f_{abc}\bar{\Psi}C_{b}T_{c}\Psi. (12)

On the other hand the variation δ\delta\mathcal{L} of the Lagrangian can be also written as,

δ=i=1N[ψiδψi+(μψi)δ(μψi)\displaystyle\delta\mathcal{L}=\sum_{i=1}^{N}[\frac{\partial\mathcal{L}}{\partial\psi_{i}}\delta\psi_{i}+\frac{\partial\mathcal{L}}{\partial(\partial_{\mu}\psi_{i})}\delta(\partial_{\mu}\psi_{i})
+ψi¯δψi¯+(μψi¯)δ(μψi¯)]\displaystyle+\frac{\partial\mathcal{L}}{\partial\bar{\psi_{i}}}\delta\bar{\psi_{i}}+\frac{\partial\mathcal{L}}{\partial(\partial_{\mu}{\bar{\psi_{i}})}}\delta(\partial_{\mu}\bar{\psi_{i}})] (13)

where ii enumerates the Dirac spinors. Employing the Euler-Lagrange equations for each fermion ii,

ψiμ((μψi))=0\displaystyle\frac{\partial\mathcal{L}}{\partial\psi_{i}}-\partial_{\mu}\big{(}{\frac{\partial\mathcal{L}}{\partial(\partial_{\mu}\psi_{i})}}\big{)}=0
ψ¯iμ((μψ¯i))=0\displaystyle\displaystyle{\frac{\partial\mathcal{L}}{\partial\bar{\psi}_{i}}}-\partial_{\mu}\big{(}\displaystyle{\frac{\partial\mathcal{L}}{\partial(\partial_{\mu}\bar{\psi}_{i})}}\big{)}=0 (14)

and the relations,

(μψ¯i)=0,δ(μψi)=μ(δψi).\frac{\partial\mathcal{L}}{\partial(\partial_{\mu}\bar{\psi}_{i})}=0~{}~{},~{}~{}\delta(\partial_{\mu}\psi_{i})=\partial_{\mu}(\delta\psi_{i}). (15)

we find that the Lagrangian variation is given by

δ=i=1N[μ(μψiδψi)].\delta\mathcal{L}=\sum_{i=1}^{N}[\partial_{\mu}\big{(}{\frac{\partial\mathcal{L}}{\partial\partial_{\mu}\psi_{i}}}\delta\psi_{i}\big{)}]. (16)

Using Eq. (9) and after some algebra, we find

δ=μ(Ψ¯(γμ𝟙N)TaΨ)δθa\delta\mathcal{L}=\partial_{\mu}(\bar{\Psi}(\gamma^{\mu}\otimes\mathds{1}_{N})T_{a}\Psi)\delta\theta^{a} (17)

Combining Eq. (12) and Eq. (17) and keeping only first order terms in δθa\delta\theta^{a} we obtain a GCE for the Dirac equation,

μ(Ψ¯(γμTa)Ψ)=fabcΨ¯(CbTc)Ψ,\partial_{\mu}(\bar{\Psi}\left(\gamma^{\mu}\otimes T_{a}\right)\Psi)=f_{abc}\bar{\Psi}\left(C_{b}\otimes T_{c}\right)\Psi, (18)

which is associated to SU(N)SU(N) symmetry transformations. Equation (18) is one of the core results of this work. The term “generalized” in the derived continuity equation has a two-fold meaning. Firstly, Eq. (18) involves non-zero source terms and therefore it does not, directly, lead to a 4-current conservation. Secondly, the 4-current Jaμ=Ψ¯γμTaΨJ^{\mu}_{a}=\bar{\Psi}\gamma^{\mu}T_{a}\Psi is not the usual current, but an abstract mathematical quantity that connects the solutions of different Dirac problems [see the definition of Ψ\Psi in Eq. (3)]. Thus, Eq. (18), in fact, corresponds to N21N^{2}-1 different equations and the number of the emerging 4-currents is equal to the number of the SU(N)SU(N) group’s generators. In the case of a diagonal potential matrix VV, as the one discussed here, and due to the property faac=0f_{aac}=0 of the structure constants fabcf_{abc}, the remaining not trivial currents will be those related to the non-diagonal elements of an SU(N)SU(N) matrix in the fundamental representation. These currents are not all independent of each another. They appear in Hermitian conjugate pairs and, in this sense, for every emerging current, there is a corresponding Hermitian conjugate, which is described by Eq. (18), as well.

Of particular interest is the case where the source terms of the GCE in Eq. (18) vanish, leading to a new class of conservation laws. Here we will examine the possible scenarios which lead to vanishing source terms in Eq. (18). These can be classified in the following two categories: (i) Global conservation laws. The source terms become zero due to a global property of the potential and holds for any spinor Ψ\Psi. In this case the emerging conservation law will be valid globally. (ii) Local conservation laws. The source terms become zero due to a local property of the potential and holds for any Ψ\Psi. As local we consider here a property of the potentials which is valid within a finite spatial domain. The emerging conservation law will also be valid only within the spatial region where this property of the potentials holds.

In principle, there is a third category related to the accidental case where the source terms become zero from a suitable combination of Ψ\Psi and VV properties. Apparently, this would require a fine tuning in Ψ\Psi and VV and the emerging conservation law would not necessarily correspond to a symmetry of the potentials. Nevertheless, this case will not be considered here due to the lack of a specific example leading to a non-trivial generalized current.

Refer to caption
Figure 1: (a) Schematic of two different potential landscapes V1V_{1} and V2V_{2} of two distinct setups which have a common profile indicated by the region 𝒟\mathcal{D}. (b) Schematic of two different potential landscapes V1V_{1} and V2V_{2} of two distinct setups. The domain 𝒟\mathcal{D} of V1V_{1} is linked to the domain 𝒟¯\bar{\mathcal{D}} of V2V_{2} via an inversion symmetry transform.

The most straightforward way to set the source terms to zero is by setting the coefficients Cb=0C_{b}=0. Observing the explicit form of the coefficients as shown in the Appendix A and for a diagonal matrix VV, as in our case, this condition can be realized if in a spatial domain 𝒟\mathcal{D} holds,

Vi(x)=Vj(x);ijV_{i}(\vec{x})=V_{j}(\vec{x})~{}~{}~{};~{}~{}~{}i\neq j (19)

where i,j=1,2,,Ni,j=1,2,\cdots,N, enumerate the different components of Ψ\Psi. Obviously, this means that in every region 𝒟\mathcal{D} where Eq. (19) holds, a locally conserved 4-current emerges. Such a case, where two different potential landscapes of two distinct problems, have a common profile within a region 𝒟\mathcal{D}, is schematically shown in Fig. 1 (a). Note, that since the potential term ViV_{i}, which appears in the ii-th Dirac equation, is a matrix Vi=Vi𝟙4V_{i}=V_{i}\cdot\mathds{1}_{4}, the coefficients CbC_{b} will also be matrices Cb=Cb𝟙4C_{b}=C_{b}\cdot\mathds{1}_{4}. This, in turn, leads to the commutation property

Ψ¯CbfabcTcΨ=CbfabcΨ¯TcΨ,\bar{\Psi}C_{b}f_{abc}T_{c}\Psi=C_{b}f_{abc}\bar{\Psi}T_{c}\Psi, (20)

which can be generalized for the case where the potential terms and the coefficients CbC_{b} are not proportional to the unit matrix but to γ0\gamma^{0}, i.e. Vi=Viγ0V_{i}=V_{i}\cdot\gamma^{0} and Cb=Cbγ0C_{b}=C_{b}\cdot\gamma^{0}. Then, Eq. (20) becomes

Ψ¯CbfabcTcΨ=CbfabcΨTcΨ.\bar{\Psi}C_{b}f_{abc}T_{c}\Psi=C_{b}f_{abc}\Psi^{\dagger}T_{c}\Psi. (21)

A similar procedure can be developed in the presence of gauge fields. Using an SU(N)SU(N) gauge transformation of Ψ\Psi, the Lagrangian density 0\mathcal{L}_{0} becomes 0=i𝟙N\mathcal{L}_{0}=i\not{D}\otimes\mathds{1}_{N} with =γμDμ\not{D}=\gamma^{\mu}D_{\mu} and Dμ=μ+iAμaTaD_{\mu}=\partial_{\mu}+iA^{a}_{\mu}T_{a} the covariant derivative containing the gauge field AμaA^{a}_{\mu}. Following the process described above, we find that the corresponding GCE is,

μ(Ψ¯γμTaΨRμνdfabdAνb)=fabcCbΨ¯TcΨ.\partial_{\mu}(\bar{\Psi}\gamma^{\mu}T_{a}\Psi-R^{\mu\nu d}f_{ab}^{d}A_{\nu}^{b})=f_{abc}{C_{b}}\bar{\Psi}T_{c}\Psi. (22)

The detailed derivation of Eq. (22) can be found in the Appendix B.

III Generalized continuity equation for two Dirac Systems

To illustrate the properties of the Dirac GCE in a more transparent way, we will focus on the case of two distinct systems each consisting of one Dirac Fermion. This composite system is described by the super-Lagrangian =1+2\mathcal{L}=\mathcal{L}_{1}+\mathcal{L}_{2},

=[ψ¯1ψ¯2][i∂̸V100i∂̸V2][ψ1ψ2].\mathcal{L}=\left[\begin{array}[]{cc}\bar{\psi}_{1}&\bar{\psi}_{2}\end{array}\right]\left[\begin{array}[]{cc}i\not{\partial}-V_{1}&0\\ 0&i\not{\partial}-V_{2}\end{array}\right]\left[\begin{array}[]{c}\psi_{1}\\ \psi_{2}\end{array}\right]. (23)

Recalling that the representation of the SU(2)SU(2) generators are Ta=σa/2T_{a}=\sigma^{a}/2, where σa\sigma^{a} are the Pauli matrices, and that the 3D Levi-Civita symbol ϵabc\epsilon_{abc} corresponds to the structure constants of the group (fabc=ϵabcf_{abc}=\epsilon_{abc}), we find that, for a global transformation, the emerging GCEs, which hold for any pair of arbitrary states ψ1,ψ2\psi_{1},\psi_{2}, according to Eq. (18) are,

μ(ψ¯1γμψ2)\displaystyle\partial_{\mu}(\bar{\psi}_{1}\gamma^{\mu}\psi_{2}) =i(V1V2)ψ¯1ψ2\displaystyle=i(V_{1}-V_{2})\bar{\psi}_{1}\psi_{2} (24)
μ(ψ¯2γμψ1)\displaystyle\partial_{\mu}(\bar{\psi}_{2}\gamma^{\mu}\psi_{1}) =i(V2V1)ψ¯2ψ1\displaystyle=i(V_{2}-V_{1})\bar{\psi}_{2}\psi_{1} (25)
μ(ψ¯1γμψ1)\displaystyle\partial_{\mu}(\bar{\psi}_{1}\gamma^{\mu}\psi_{1}) =μ(ψ¯2γμψ2),\displaystyle=\partial_{\mu}(\bar{\psi}_{2}\gamma^{\mu}\psi_{2}), (26)

where the CbC_{b} coefficients are given by Eq. (9). Obviously Eq. (26) is trivial and Eqs. (24) and (25) are Hermitian conjugates. Hence, for the rest of our analysis, and without loss of generality, we consider only Eq. (24), aiming to determine the conditions which eliminate the right-hand-side (r.h.s.) of Eq. (24).

III.1 Global conservation laws

We first consider the global conservation case i.e V1(x)=V2(x)xV_{1}(\vec{x})=V_{2}(\vec{x})~{}~{}\forall\vec{x}, where the source term in Eq. (24) vanishes in the entire space. An interesting scenario occurs when Eq. (24) involves eigenstates of the same Dirac equation which -in the absence of degeneracy- differ in energy. Note, that this holds always for 1-D systems. In higher dimensions, different eigenstates which share the same energy eigenvalue constitute also a possible realization of the global conservation scenario.

Assuming 1-D solutions of the form

ψi(x,t)=eiEitϕi(x);i=1,2\psi_{i}(\vec{x},t)=e^{-iE_{i}t}\phi_{i}(\vec{x})~{}~{}~{};~{}~{}~{}i=1,2 (27)

and inserting Eq. (27) into Eq. (24), we find

t[ϕ¯1(x)eiE1tγ0ϕ2(x)eiE2t]+\displaystyle\partial_{t}[\bar{\phi}_{1}(x)e^{iE_{1}t}\gamma^{0}\phi_{2}(x)e^{-iE_{2}t}]+
ei(E1E2)tddx[ϕ¯1(x)γ1ϕ2(x)]=0,\displaystyle e^{i(E_{1}-E_{2})t}\frac{d}{dx}[\bar{\phi}_{1}(x)\gamma^{1}\phi_{2}(x)]=0, (28)

The quantity J12(x)=ϕ¯1(x)γ1ϕ2(x)J_{12}(x)=\bar{\phi}_{1}(x)\gamma^{1}\phi_{2}(x) is identified as the spatial part of the conserved current. The spatial integral, over xx, of the time derivative in the l.h.s. of Eq. (III.1) leads to the condition,

x1x2ϕ¯1(x)γ0ϕ2(x)𝑑x=iJ12(x2)J12(x1)(E1E2).\int_{x_{1}}^{x_{2}}\bar{\phi}_{1}(x)\gamma^{0}\phi_{2}(x)dx=i\small\frac{J_{12}(x_{2})-J_{12}(x_{1})}{(E_{1}-E_{2})}. (29)

If x1x_{1}\rightarrow-\infty and x2=+x_{2}\rightarrow=+\infty, we recover the global generalized charge Q=+ϕ¯1(x)γ0ϕ2(x)𝑑xQ=\int_{-\infty}^{+\infty}\bar{\phi}_{1}(x)\gamma^{0}\phi_{2}(x)dx. This, in turn, can be expressed via currents J12J_{12} at ±\pm\infty, according to Eq. (29).

III.2 Local conservation laws

Equation (24) also supports the scenario of local conservation laws. We assume that the equality V1(x)=V2(x)V_{1}(\vec{x})=V_{2}(\vec{x}) holds only in a finite spatial domain 𝒟\mathcal{D}, as shown in Fig. 1 (a) and that there is, at least, one common energy state E1=E2=EE_{1}=E_{2}=E for the two distinct systems involved in the Lagrangian (23). Even though, the two eigenstates share the same eigenvalue, this is not a degeneracy case. These eigenstates correspond to different Dirac equations, e.g. in Fig. 1 (a) there are x\vec{x}-regions in which V1(x)V2(x)V_{1}(\vec{x})\neq V_{2}(\vec{x}). Otherwise, this would be a case of global conservation law. Note that restricting our analysis to systems which a common energy value simplifies significantly the GCE and only spatial variations remain. This condition can be always satisfied for two different scattering systems. For bound systems, on the other hand, it can be achieved only with appropriate fine tuning.

Due to the condition of Eq. (19) which holds for x𝒟\vec{x}\in\mathcal{D} [see Fig. 1 (a)], we get

(ψ¯1(x)γψ2(x))=0,x𝒟.\vec{\nabla}(\bar{\psi}_{1}(\vec{x})\vec{\gamma}\psi_{2}(\vec{x}))=0~{}~{}~{},~{}~{}~{}\vec{x}~{}\in~{}\mathcal{D}. (30)

This, in turn, leads to the divergence-free current,

J12(x)=ψ¯1(x)γψ2(x),x𝒟.\vec{J}_{12}(\vec{x})=\bar{\psi}_{1}(\vec{x})\vec{\gamma}\psi_{2}(\vec{x})~{}~{}~{},~{}~{}~{}\vec{x}~{}\in~{}\mathcal{D}. (31)

The subscripts 1,21,~{}2 denote the solutions of the Dirac equations which involve the potentials V1V_{1} and V2V_{2}, respectively. The vanishing divergence holds also for the Hermitian conjugate J12J_{12}^{\dagger} of the current J12J_{12}. It should be stressed here that this current is a completely different quantity from the usual probability current. In fact, its existence is not attributed to the presence of a particular source in region 𝒟\mathcal{D} [see Fig. 1 (a)], but it is based on the fact that outside region 𝒟\mathcal{D} the potential landscapes in the two problems differ from each other. Restricting our analysis to 1-D, Eq. (30) becomes,

x(ψ¯1(x)γxψ2(x))=0,\partial_{x}(\bar{\psi}_{1}(x)\gamma^{x}\psi_{2}(x))=0, (32)

yielding a spatially constant current x𝒟\forall x\in\mathcal{D}

J12=ψ¯1(x)σxψ2(x)=const.,x𝒟,J_{12}=\bar{\psi}_{1}(x)\sigma_{x}\psi_{2}(x)=\text{const.}~{}~{}~{},~{}~{}~{}x~{}\in~{}\mathcal{D}, (33)

where we have employed the usual convention for the γ\gamma-matrices in 1-D,

γ0=σz;γx=σx\gamma^{0}=\sigma_{z}~{}~{}~{};~{}~{}~{}\gamma^{x}=\sigma_{x} (34)

The constant current in 𝒟\mathcal{D} acts as a correlator between the solutions of two different 1-D Dirac systems, based on the equality of the potentials V1(x)V_{1}(x) and V2(x)V_{2}(x) in this region. Obviously, if the condition V1(x)=V2(x)V_{1}(x)=V_{2}(x) holds for more than one regions 𝒟\mathcal{D}, i.e in the regions 𝒟j\mathcal{D}_{j}, with j=1,2,,Mj=1,2,\cdots,M, then the respective currents Jj,12=cjJ_{j,12}=c_{j} will be constant in every region 𝒟j\mathcal{D}_{j} while, in general, cicjc_{i}\neq c_{j} for iji\neq j.

Finally, we discuss an interesting sub-case which also belongs to the class of local conservation laws. It is based on a relation of the form,

V1(x)=V2[F(x)],V_{1}(x)=V_{2}[F(x)], (35)

holding within a finite domain 𝒟\mathcal{D}. In Eq. (35) F(x)F(x) is a linear transformation under which the Laplacian operator is invariant. Notice that since the Dirac field is not a scalar, one has to take into account the spinor transformation under the specific geometric transform (leaving the Laplacian invariant) in the super-Lagrangian. To explain this in more detail let us restrict to the one spatial dimension. In 1-D, F(x)F(x) describes either an inversion Π\Pi with respect to a point aa or a translation TT by LL, as shown in Fig. 1 (b). Treating such transforms requires attention. For instance, the transformed spinor, in the inversion case, is given by ψ(x)=Πψ[F(x)]=σzψ(x)\psi^{\prime}(x^{\prime})=\Pi\psi[F(x)]=\sigma_{z}\psi(-x). The symmetry transform F(x)F(x) can be formally written as,

F:xx¯=F(x)=σx+ρ,F:~{}x\mapsto\overline{x}=F(x)=\sigma x+\rho, (36)

where the parameters σ\sigma and ρ\rho are as follows:

σ=1,\displaystyle\sigma=-1,~{} ρ=2aF=Π\displaystyle\rho=2a~{}\Rightarrow F=\Pi (Parity)\displaystyle~{}(\textrm{Parity}) (37)
σ=+1,\displaystyle\sigma=+1,~{} ρ=LF=T\displaystyle\rho=L~{}\Rightarrow F=T (Translation)\displaystyle~{}(\textrm{Translation}) (38)

The symmetry condition (35) leads, in turn, to a current which is spatially conserved in 𝒟\mathcal{D} and is expressed as,

JF,21=ψ¯1(x)σxψ2(σx+ρ)=const.J_{F,21}=\bar{\psi}_{1}(x)\sigma_{x}\psi_{2}(\sigma x+\rho)=\text{const.} (39)

or

JF,21=ψ¯1(x)σxψ2(x)=const.J_{F,21}=\bar{\psi}_{1}(x)\sigma_{x}\psi^{\prime}_{2}(x^{\prime})=\text{const.} (40)

Note, that the conservation of the generalized current JF,21J_{F,21} in the spatial domain 𝒟\mathcal{D} and its symmetry-transformed counterpart 𝒟¯\bar{\mathcal{D}} involves the original ψ1(x)\psi_{1}(x) solution which corresponds to the potential landscape V1V_{1} and the transformed solution ψ2(x)\psi^{\prime}_{2}(x^{\prime}) which corresponds to the potential landscape V2V_{2}. We will demonstrate the construction of such a current using the setup illustrated in Figure 2.

Refer to caption
Figure 2: Potential landscapes which are linked via an inversion symmetry transformation with the exception of x=0x=0 where a δ\delta-barrier is located. The color coding indicates the areas of the wavefunctions which take part in the construction of the respective generalized currents.

The landscapes (a)(a) and (b)(b) are symmetric under an inversion symmetry transformation, with the exception of the point x=0x=0 where the δ\delta-barrier λδ(x)\lambda\delta(x) (λ>0\lambda>0) is located in (a)(a). The δ\delta-barrier separates the xx-space into two distinct domains, namely D=(,0)D_{-}=(-\infty,0) and D+=(0,+)D_{+}=(0,+\infty). The generalized current acquires a different value in each of these two domains. The coloring indicates the wavefunction domains participating in the generalized current in each case. Based on Eq. (40) we have :

J12=ψ¯1(x)σxσzψ2(x)={c,xDc+,xD+J_{12}=\bar{\psi}_{1}(x)\sigma_{x}\sigma_{z}\psi_{2}(-x)=\begin{cases}c_{-}&,~{}~{}x\in D_{-}\\ c_{+}&,~{}~{}x\in D_{+}\end{cases} (41)

The constants c±c_{\pm} can be expressed through the left ψi(0)\psi_{i}(0^{-}) and right ψi(0+)\psi_{i}(0^{+}) (i=1,2i=1,~{}2) limiting values of the Dirac field at x=0x=0, as,

J12={ψ¯1(0)σxσzψ2(0+),xDψ¯1(0+)σxσzψ2(0),xD+,J_{12}=\begin{cases}\bar{\psi}_{1}(0^{-})\sigma_{x}\sigma_{z}\psi_{2}(0^{+})&,~{}~{}x\in D_{-}\\ \bar{\psi}_{1}(0^{+})\sigma_{x}\sigma_{z}\psi_{2}(0^{-})&,~{}~{}x\in D_{+},\end{cases} (42)

Combining the properties of the Pauli matrices, the definition of ψ¯(x)\bar{\psi}(x) in 1-D, and the following expressions as given in 1ddirac :

ψ1(0+)\displaystyle\psi_{1}(0^{+}) =exp(iλσy)ψ1(0)\displaystyle=\exp\left({-i\lambda\sigma_{y}}\right)\psi_{1}(0^{-}) (43)
ψ2(0+)\displaystyle\psi_{2}(0^{+}) =ψ2(0),\displaystyle=\psi_{2}(0^{-}), (44)

we find that the constant generalized current can be expressed at x=0x=0 as,

J12={ψ1(0)σxψ2(0),xDψ1(0)exp(iλσy)σxψ2(0),xD+J_{12}=\begin{cases}-\psi^{\dagger}_{1}(0^{-})\sigma_{x}\psi_{2}(0^{-})&,~{}~{}x\in D_{-}\\ -\psi^{\dagger}_{1}(0^{-})\exp\left({i\lambda\sigma_{y}}\right)\sigma_{x}\psi_{2}(0^{-})&,~{}~{}x\in D_{+}\end{cases} (45)

Since exp(iλσy)\exp\left(i\lambda\sigma_{y}\right) represents a rotation matrix:

exp(iλσy)=(cosλsinλsinλcosλ)\exp\left(i\lambda\sigma_{y}\right)=\left(\begin{array}[]{cc}\cos\lambda&\sin\lambda\\ -\sin\lambda&\cos\lambda\end{array}\right) (46)

we observe that the current J12J_{12} in D+D_{+} differs from J12J_{12} in DD_{-} in containing the spinor σxψ2(0)\sigma_{x}\psi_{2}(0^{-}) with a phase shift induced by the δ\delta-barrier. The phase shift angle is given by λ\lambda. Obviously, when λ=0\lambda=0 there is a single value for the constant J12J_{12} in the entire xx-space.

IV SU(N)SU(N) complete set of GCEs for Schrödinger systems

In this section we return to the Schrödinger framework and we generalize the results obtained in Diakonos2019 for non-relativistic quantum systems. To this end, we use a complete set of SU(N)SU(N) transformations instead of those belonging to the Cartan sub-algebra, as in Diakonos2019 . In addition, we include a gauge field in our analysis. The super-Lagrangian in the Schrödinger framework is written as,

=i=1Ni=Ψ(H0V)Ψ=0+~+h.c,\mathcal{L}=\sum_{i=1}^{N}\mathcal{L}_{i}=\Psi^{\dagger}(H_{0}-V)\Psi=\mathcal{L}_{0}+\mathcal{\tilde{L}}+h.c, (47)

with

i=Ψi(x,t)[itp^22mVi(x)]Ψi(x,t)+h.c.\mathcal{L}_{i}=\Psi_{i}^{\star}(\vec{x},t)\left[i\partial_{t}-\frac{\hat{\vec{p}}^{2}}{2m}-V_{i}(\vec{x})\right]\Psi_{i}(\vec{x},t)~{}+~{}h.c. (48)

and

Ψ=(Ψ1(x,t)Ψ2(x,t)ΨN(x,t)).\Psi=\begin{pmatrix}\Psi_{1}(\vec{x},t)\\ \Psi_{2}(\vec{x},t)\\ \vdots\\ \Psi_{N}(\vec{x},t)\end{pmatrix}. (49)

In Eqs. (48), (49), Ψi(x,t)\Psi_{i}(\vec{x},t) is the wavefunction of the ii-th system involved in the Schrödinger super-Lagrangian. Obviously,

0=ΨH0Ψ;~=ΨVΨ\mathcal{L}_{0}=\Psi^{\dagger}H_{0}\Psi~{}~{}~{};~{}~{}~{}\mathcal{\tilde{L}}=-\Psi^{\dagger}V\Psi (50)

with

H0=(itp^22m)𝟙NH_{0}=(i\partial_{t}-\frac{\hat{\vec{p}}^{2}}{2m})\otimes\mathds{1}_{N} (51)

and

V=diag(V1,V2,,VN)V=\text{diag}(V_{1},V_{2},\ldots,V_{N}) (52)

The symmetry transformation of Eq. (8) leads directly to the variation relations δ0=0,δ~0\delta\mathcal{L}_{0}=0,~{}\delta\mathcal{\tilde{L}}\neq 0 and as expected, once again δ=δ~0\delta\mathcal{L}=\delta\mathcal{\tilde{L}}\neq 0. Since δ~=~~\delta\mathcal{\tilde{L}}=\mathcal{\tilde{L}}^{\prime}-\mathcal{\tilde{L}} we find that the Lagrangian density variation is given by

δ=2Ψ(UVUV)Ψ.\delta\mathcal{L}=-2\Psi^{\dagger}(U^{\dagger}VU-V)\Psi. (53)

Substituting Eq. (9) we find that,

UVUV=δθaCkfakcTcU^{\dagger}VU-V=-\delta\theta_{a}C_{k}f_{ak}^{c}T_{c} (54)

and consequently

δ=2δθaCbfabcΨTcΨ.\delta\mathcal{L}=2\delta\theta^{a}C_{b}f_{abc}\Psi^{\dagger}T_{c}\Psi. (55)

The variation of δ\delta\mathcal{L} can also be written as

δ=i=1N[ΨiδΨi+(μΨi)δ(μΨi)+\displaystyle\delta\mathcal{L}=\sum_{i=1}^{N}[\frac{\partial\mathcal{L}}{\partial\Psi_{i}}\delta\Psi_{i}+\frac{\partial\mathcal{L}}{\partial(\partial_{\mu}\Psi_{i})}\delta(\partial_{\mu}\Psi_{i})+
ΨiδΨi+(μΨi)δ(μΨi)],\displaystyle\frac{\partial\mathcal{L}}{\partial{\Psi^{\star}_{i}}}\delta{\Psi^{\star}_{i}}+\frac{\partial\mathcal{L}}{\partial(\partial_{\mu}{{\Psi_{i}^{\star}})}}\delta(\partial_{\mu}{\Psi^{\star}_{i}})], (56)

which after some algebraic calculations leads to,

δ=[t(ΨTaΨ)+im[(Ψ)TaΨ]+h.c]δθa.\delta\mathcal{L}=[\partial_{t}(\Psi^{\dagger}T_{a}\Psi)+\frac{i}{m}\vec{\nabla}[\vec{\nabla}(\Psi^{\dagger})T_{a}\Psi]+h.c]\delta\theta^{a}. (57)

Combining Eq. (55) and Eq. (57) we finally get

t(ΨTaΨ)+i2m(ΨTaΨΨTaΨ)=\displaystyle\partial_{t}(\Psi^{\dagger}T_{a}\Psi)+\frac{i}{2m}\vec{\nabla}(\vec{\nabla}\Psi^{\dagger}T_{a}\Psi-\Psi^{\dagger}T_{a}\vec{\nabla}\Psi)=
fabcCbΨTcΨ\displaystyle f_{abc}C_{b}\Psi^{\dagger}T_{c}\Psi (58)

which generalizes the GCE found in Diakonos2019 .

Note, that the source terms both in the Schrödinger and the Dirac GCEs have the same form [see Eq. (20) and Eq. (IV)]. The conservation of the corresponding generalized currents implies the same conditions both for the relativistic and the non-relativistic case. On the contrary, the form of the conserved current differs between these two cases, i.e. in the non-relativistic GCE the current contains derivatives of the wave field while in the relativistic version there are no derivatives involved. This is due to the difference in the form of the kinetic term in either case.

V Summary & Conclusions

In this work we develop a systematic approach to derive generalized 4-currents built from fermionic fields which describe solutions of Dirac problems involving different external potential terms. This is achieved through the construction of a super-Lagrangian which allows the synthesis of a composite system acting as a generator for all involved sub-problems along the lines presented in Diakonos2019 in the context of Schrödinger (non-relativistic) quantum mechanics. The derived currents are related to symmetry transforms of the super-Lagrangian and they obey generalized continuity equations which involve source terms dictated by the specific transform. These source terms occur because the super-Lagrangian is not invariant under this transformation. Following this procedure, we have explored possible scenarios based on conditions on the form of the external potentials leading to the vanishing of the source terms globally or in finite spatial domains. In the second case local conservation laws manifest, relating solutions of Dirac equations which differ in the external potential terms. We have illustrated in a simple example how such a current is constructed in practice. In contrast to the Schrödinger case presented in Diakonos2019 the generalized Dirac currents do not involve derivative terms and therefore they can be considered as a direct mapping between solutions of different Dirac equations. It would be interesting to apply this procedure to specific problems and explore the insight gained by such a mapping. The corresponding local conservation laws may be useful for the control as well as the preparation of fermionic states through suitable modulation of the external potential(s). Here we have considered exclusively SU(N)SU(N) transformations of the super-field formed from the solutions of the different Dirac problems. Future perspectives of this work would be the investigation of other symmetry transformations and how new conservation laws could emerge.

Appendix A Analysis of V-matrix to SU(N) Representations of Generators

A general N×NN\times N VV-matrix is analyzed to the basis of the generators of the SU(N)SU(N) group using the formula,

V=(V11V1NVN1VNN)=V0𝟙N+k=1N21CkTkV=\left(\begin{array}[]{ccc}V_{11}&\cdots&V_{1N}\\ \vdots&\ddots&\vdots\\ V_{N1}&\cdots&V_{NN}\end{array}\right)=V_{0}\mathds{1}_{N}+\sum_{k=1}^{N^{2}-1}C_{k}T_{k}

where

V0=Tr(V𝟙N)N;Ck=Tr(VTk).V_{0}=\frac{Tr(V\cdot\mathds{1}_{N})}{N}~{}~{}~{};~{}~{}~{}C_{k}=Tr(V\cdot T_{k}).

The explicit form of the coefficient V0V_{0} is,

V0=V11+V22++VNNNV_{0}=\frac{V_{11}+V_{22}+\cdots+V_{NN}}{N} (59)

For k=n21k=n^{2}-1 (n=2,3,,Nn=2,3,\dots,N), the coefficients CkC_{k} which are related to the diagonal representations of the generators are,

C3\displaystyle C_{3} =V11(x)V22(x)\displaystyle=V_{11}(x)-V_{22}(x)
C8\displaystyle C_{8} =V11(x)+V22(x)2V33(x)3\displaystyle=\frac{V_{11}(x)+V_{22}(x)-2V_{33}(x)}{\sqrt{3}}
\displaystyle\vdots~{}~{} \displaystyle\qquad\qquad~{}~{}~{}\vdots
CN21\displaystyle C_{N^{2}-1} =V11(x)+(N1)VNN(x)N(N1)2.\displaystyle=\frac{V_{11}(x)+\cdots-(N-1)V_{NN}(x)}{\sqrt{\frac{N(N-1)}{2}}}.

Accordingly, the coefficients which are related to the non-diagonal representations of the generators are,

Ck\displaystyle C_{k} =Vk,k+1(x)+Vk+1,k(x)\displaystyle=V_{k,k+1}(x)+V_{k+1,k}(x)
Ck+1\displaystyle C_{k+1} =[Vk,k+1(x)Vk+1,k(x)]i\displaystyle=[V_{k,k+1}(x)-V_{k+1,k}(x)]i

where,

k\displaystyle k =n2+2m\displaystyle=n^{2}+2m
n\displaystyle n =1,2,,N\displaystyle=1,2,\dots,N
m\displaystyle m =0,1,,n1.\displaystyle=0,1,\dots,n-1.

Employing our analysis for the full-element VV-matrix, we derive the GCEs for an SU(N)SU(N) global infinitesimal transformation for NN Dirac problems:

μ(ψ¯iγμψj)\displaystyle\partial_{\mu}(\bar{\psi}_{i}\gamma^{\mu}\psi_{j}) =i(ViiVjj)ψ¯iψj\displaystyle=i(V_{ii}-V_{jj})\bar{\psi}_{i}\psi_{j}
+iVji(ψ¯jψjψ¯iψi)\displaystyle+iV_{ji}(\bar{\psi}_{j}\psi_{j}-\bar{\psi}_{i}\psi_{i})
+iVkiψ¯kψjiVjkψ¯iψk\displaystyle+iV_{ki}\bar{\psi}_{k}\psi_{j}-iV_{jk}\bar{\psi}_{i}\psi_{k}
μ(ψ¯iγμψiψ¯jγμψj)\displaystyle\partial_{\mu}(\bar{\psi}_{i}\gamma^{\mu}\psi_{i}-\bar{\psi}_{j}\gamma^{\mu}\psi_{j}) =2iVjiψ¯jψi2iVijψ¯iψj\displaystyle=2iV_{ji}\bar{\psi}_{j}\psi_{i}-2iV_{ij}\bar{\psi}_{i}\psi_{j}
+iVjkψ¯jψkiVkjψ¯kψj\displaystyle+iV_{jk}\bar{\psi}_{j}\psi_{k}-iV_{kj}\bar{\psi}_{k}\psi_{j}
+iVkiψ¯kψiiVikψ¯iψk\displaystyle+iV_{ki}\bar{\psi}_{k}\psi_{i}-iV_{ik}\bar{\psi}_{i}\psi_{k}

where

i,j,k=1,2,,N\displaystyle i,j,k=1,2,\dots,N
ijk\displaystyle i\neq j\neq k

and kk sum over all the remaining indices. The non-diagonal terms express a direct connection between the ii-th Dirac problem with the N1N-1 remaining problems.

Appendix B Gauge Transformation and GCE

We present the basic steps towards a GCE which is induced by a gauge transformation, for NN Dirac problems. A gauge symmetric Lagrangian under a U=eiθa(x)TaU=e^{-i\theta^{a}(x)T_{a}} transformation is,

=Ψ¯(im)Ψ14RμνaRμνa{\mathcal{L}=\bar{\Psi}(i\not{D}-m)\Psi-\frac{1}{4}R_{\mu\nu}^{a}R^{\mu\nu a}}

where

Dμ\displaystyle D_{\mu} =μ+iAμaTa\displaystyle=\partial_{\mu}+iA^{a}_{\mu}T_{a}
Rμνa\displaystyle R_{\mu\nu}^{a} =μAνaνAμafbcaAμbAνc\displaystyle=\partial_{\mu}A_{\nu}^{a}-\partial_{\nu}A_{\mu}^{a}-f_{bc}^{a}A_{\mu}^{b}A_{\nu}^{c}

and AμaA_{\mu}^{a} are scalar fields. Demanding DμΨ=UDμΨD_{\mu}^{\prime}\Psi^{\prime}=UD_{\mu}\Psi, this Lagrangian becomes invariant under the corresponding gauge transformation. The Lagrangian of these different problems is,

=Ψ¯(i∂̸aTaV)Ψ14RμνaRμνa,{\mathcal{L}=\bar{\Psi}(i\not{\partial}-\not{A^{a}}T_{a}-V)\Psi-\frac{1}{4}R_{\mu\nu}^{a}R^{\mu\nu a}},

where VV is a diagonal matrix. In that case the Lagrangian becomes,

=Ψ¯(iV)Ψ14RμνaRμνa=0+~\displaystyle\mathcal{L}=\bar{\Psi}(i\not{D}-V)\Psi-\frac{1}{4}R_{\mu\nu}^{a}R^{\mu\nu a}=\mathcal{L}_{0}+\tilde{\mathcal{L}}

with

0\displaystyle\mathcal{L}_{0} =Ψ¯iΨ14RμνaRμνa\displaystyle=\bar{\Psi}i\not{D}\Psi-\frac{1}{4}R_{\mu\nu}^{a}R^{\mu\nu a}
~\displaystyle\tilde{\mathcal{L}} =Ψ¯VΨ\displaystyle=-\bar{\Psi}V\Psi

Following the same procedure we find that

δ~=Ψ¯θa(x)CbfabcTcΨ,\displaystyle{\delta\tilde{\mathcal{L}}=\bar{\Psi}\theta^{a}(x){C_{b}}f_{ab}^{c}T_{c}\Psi},

and

δ=μ(Ψ¯γμTaΨ+RμνcfbacAνb)θa(x).\delta\mathcal{L}=\partial_{\mu}(\bar{\Psi}\gamma^{\mu}T_{a}\Psi+R^{\mu\nu c}f_{ba}^{c}A_{\nu}^{b})\theta^{a}(x).

Finally, we find that the GCE generated by a gauge SU(N)SU(N) transformation is,

μ(Ψ¯γμTaΨRμνdfabdAνb)=Ψ¯CbfabcTcΨ.\displaystyle\partial_{\mu}(\bar{\Psi}\gamma^{\mu}T_{a}\Psi-R^{\mu\nu d}f_{ab}^{d}A_{\nu}^{b})=\bar{\Psi}\ C_{b}f_{ab}^{c}T_{c}\Psi.

This GCE is similar to the GCEs for global transformations, having only an extra term in the 4-gradient, as expected.

References

  • (1) E. Noether, Invariante Variationsprobleme, Nachr. D. König. Gesellsch. D. Wiss. Zu Göttingen, Math-phys. Klasse 1918: 235–25 (1918).
  • (2) A. G. B. Spourdalakis, G. Pappas, C. V. Morfonios, P. A. Kalozoumis, F. K. Diakonos and P. Schmelcher, "Generalized continuity equations from two-field Schrodinger Lagrangians", Phys. Rev. A 94, 052122 (2016).
  • (3) P. A. Kalozoumis, C. Morfonios, N. Palaiodimopoulos, F. K. Diakonos, and P. Schmelcher, Phys. Rev. A 88, 033857 (2013).
  • (4) P.A. Kalozoumis, O. Richoux, F.K. Diakonos, G. Theocharis and P. Schmelcher, Phys. Rev. B 92, 014303 (2015).
  • (5) P. A. Kalozoumis, C. Morfonios, F. K. Diakonos, and P. Schmelcher, Phys. Rev. A 87, 032113 (2013).
  • (6) P. A. Kalozoumis, C. Morfonios, F. K. Diakonos and P. Schmelcher, Phys. Rev. Lett. 113, 050403, (2014).
  • (7) F. K. Diakonos and P. Schmelcher, "Super-Lagrangian and variational principle for generalized continuity equations", J. Phys. A: Math. and Theor. 52, 1552034 (2019).
  • (8) Kennedy, P. and Dombey, N. "Low momentum scattering in the Dirac equation." J. Phys. A: Math. Theor. 35, 6645 (2002).
Summary of GCEs for Schrödinger and Dirac Problems
Schrödinger Dirac
SU(N) 3-D t(ΨTaΨ)+i2m(ΨTaΨΨTaΨ)=CbfabcΨTcΨ\partial_{t}(\Psi^{\dagger}T_{a}\Psi)+\frac{i}{2m}\vec{\nabla}(\vec{\nabla}\Psi^{\dagger}T_{a}\Psi-\Psi^{\dagger}T_{a}\vec{\nabla}\Psi)=C_{b}f_{abc}\Psi^{\dagger}T_{c}\Psi μ(Ψ¯γμTaΨ)=fabcΨ¯CbTcΨ\partial_{\mu}(\bar{\Psi}\gamma^{\mu}\otimes T_{a}\Psi)=f_{abc}\bar{\Psi}C_{b}\otimes T_{c}\Psi
Common Energy Condition i2m(ΨTaΨΨTaΨ)=CbfabcΨTcΨ\frac{i}{2m}\vec{\nabla}(\vec{\nabla}\Psi^{\dagger}T_{a}\Psi-\Psi^{\dagger}T_{a}\vec{\nabla}\Psi)=C_{b}f_{abc}\Psi^{\dagger}T_{c}\Psi (Ψ¯γTaΨ)=fabcΨ¯CbTcΨ\vec{\nabla}(\bar{\Psi}\vec{\gamma}\otimes T_{a}\Psi)=f_{abc}\bar{\Psi}C_{b}\otimes T_{c}\Psi
Vanishing Source Terms Condition i2m(ΨTaΨΨTaΨ)=0\frac{i}{2m}\vec{\nabla}(\vec{\nabla}\Psi^{\dagger}T_{a}\Psi-\Psi^{\dagger}T_{a}\vec{\nabla}\Psi)=0 (Ψ¯γTaΨ)=0\vec{\nabla}(\bar{\Psi}\vec{\gamma}\otimes T_{a}\Psi)=0
Divergence Free Currents : J12a(x)=i2m[Ψ(x)TaΨ(x)Ψ(x)TaΨ(x)]\vec{J}_{12}^{a}(\vec{x})=\frac{i}{2m}\left[\vec{\nabla}\Psi^{\dagger}(\vec{x})T_{a}\Psi(\vec{x})-\Psi(\vec{x})^{\dagger}T_{a}\vec{\nabla}\Psi(\vec{x})\right] J12a(x)=Ψ¯(x)γTaΨ(x)\vec{J}_{12}^{a}(\vec{x})=\bar{\Psi}(\vec{x})\vec{\gamma}\otimes T_{a}\Psi(\vec{x})
SU(N) 1-D t(ΨTaΨ)+i2mx(xΨTaΨΨTaxΨ)=CbfabcΨTcΨ\partial_{t}(\Psi^{\dagger}T_{a}\Psi)+\frac{i}{2m}\partial_{x}(\partial_{x}\Psi^{\dagger}T_{a}\Psi-\Psi^{\dagger}T_{a}\partial_{x}\Psi)=C_{b}f_{abc}\Psi^{\dagger}T_{c}\Psi t(ΨTaΨ)+x(Ψ¯γxTaΨ)=CbfabcΨ¯TcΨ\partial_{t}(\Psi^{\dagger}T_{a}\Psi)+\partial_{x}(\bar{\Psi}\gamma^{x}T_{a}\Psi)=C_{b}f_{abc}\bar{\Psi}T_{c}\Psi
Common Energy Condition i2mx(xΨTaΨΨTaxΨ)=CbfabcΨTcΨ\frac{i}{2m}\partial_{x}(\partial_{x}\Psi^{\dagger}T_{a}\Psi-\Psi^{\dagger}T_{a}\partial_{x}\Psi)=C_{b}f_{abc}\Psi^{\dagger}T_{c}\Psi x(Ψ¯γxTaΨ)=CbfabcΨ¯TcΨ\partial_{x}(\bar{\Psi}\gamma^{x}T_{a}\Psi)=C_{b}f_{abc}\bar{\Psi}T_{c}\Psi
Vanishing Source Terms Condition i2mx(xΨ(x)TaΨ(x)Ψ(x)TaxΨ(x))=0\frac{i}{2m}\partial_{x}(\partial_{x}\Psi^{\dagger}(x)T_{a}\Psi(x)-\Psi^{\dagger}(x)T_{a}\partial_{x}\Psi(x))=0 x(Ψ¯(x)γxTaΨ(x))=0\partial_{x}(\bar{\Psi}(x)\gamma^{x}\otimes T_{a}\Psi(x))=0
Conserved Currents J12a=i2m[xΨ(x)TaΨ(x)Ψ(x)TaxΨ]=const.J^{a}_{12}=\frac{i}{2m}\left[\partial_{x}\Psi^{\dagger}(x)T_{a}\Psi(x)-\Psi(x)^{\dagger}T_{a}\partial_{x}\Psi\right]=\text{const.} J12a=Ψ¯(x)γxTaΨ(x)=const.J^{a}_{12}=\bar{\Psi}(x)\gamma^{x}\otimes T_{a}\Psi(x)=\text{const.}
SU(2) 1-D t(Ψ1Ψ2)+i2mx(xΨ1Ψ2Ψ1xΨ2)=i(V1V2)Ψ1Ψ2\displaystyle\partial_{t}(\Psi_{1}^{\star}\Psi_{2})+\frac{i}{2m}\partial_{x}(\partial_{x}\Psi_{1}^{\star}\Psi_{2}-\Psi_{1}^{\star}\partial_{x}\Psi_{2})=i(V_{1}-V_{2})\Psi_{1}^{\star}\Psi_{2} t(Ψ2Ψ1)+i2mx(xΨ2Ψ1Ψ2xΨ1)=i(V2V1)Ψ2Ψ1\displaystyle\partial_{t}(\Psi_{2}^{\star}\Psi_{1})+\frac{i}{2m}\partial_{x}(\partial_{x}\Psi_{2}^{\star}\Psi_{1}-\Psi_{2}^{\star}\partial_{x}\Psi_{1})=i(V_{2}-V_{1})\Psi_{2}^{\star}\Psi_{1} t(Ψ1Ψ1)+i2mx(xΨ1Ψ1Ψ1xΨ1)=\displaystyle\partial_{t}(\Psi_{1}^{\star}\Psi_{1})+\frac{i}{2m}\partial_{x}(\partial_{x}\Psi_{1}^{\star}\Psi_{1}-\Psi_{1}^{\star}\partial_{x}\Psi_{1})= t(Ψ2Ψ2)+i2mx(xΨ2Ψ2Ψ2xΨ2)\displaystyle\partial_{t}(\Psi_{2}^{\star}\Psi_{2})+\frac{i}{2m}\partial_{x}(\partial_{x}\Psi_{2}^{\star}\Psi_{2}-\Psi_{2}^{\star}\partial_{x}\Psi_{2}) μ(ψ¯1γμψ2)\displaystyle\partial_{\mu}(\bar{\psi}_{1}\gamma^{\mu}\psi_{2}) =i(V1V2)ψ¯1ψ2\displaystyle=i(V_{1}-V_{2})\bar{\psi}_{1}\psi_{2} μ(ψ¯2γμψ1)\displaystyle\partial_{\mu}(\bar{\psi}_{2}\gamma^{\mu}\psi_{1}) =i(V2V1)ψ¯2ψ1\displaystyle=i(V_{2}-V_{1})\bar{\psi}_{2}\psi_{1} μ(ψ¯1γμψ1)\displaystyle\partial_{\mu}(\bar{\psi}_{1}\gamma^{\mu}\psi_{1}) =μ(ψ¯2γμψ2)\displaystyle=\partial_{\mu}(\bar{\psi}_{2}\gamma^{\mu}\psi_{2})
Common Energy Condition i2mx(xΨ1Ψ2Ψ1xΨ2)=i(V1V2)Ψ1Ψ2\displaystyle\frac{i}{2m}\partial_{x}(\partial_{x}\Psi_{1}^{\star}\Psi_{2}-\Psi_{1}^{\star}\partial_{x}\Psi_{2})=i(V_{1}-V_{2})\Psi_{1}^{\star}\Psi_{2} x(ψ¯1γxψ2)\displaystyle\partial_{x}(\bar{\psi}_{1}\gamma^{x}\psi_{2}) =i(V1V2)ψ¯1ψ2\displaystyle=i(V_{1}-V_{2})\bar{\psi}_{1}\psi_{2}
Condition for V1=V2V_{1}=V_{2} i2mx(xΨ1Ψ2Ψ1xΨ2)=0\displaystyle\frac{i}{2m}\partial_{x}(\partial_{x}\Psi_{1}^{\star}\Psi_{2}-\Psi_{1}^{\star}\partial_{x}\Psi_{2})=0 x(ψ¯1γxψ2)\displaystyle\partial_{x}(\bar{\psi}_{1}\gamma^{x}\psi_{2}) =0\displaystyle=0
Conserved Current J12=i2m[xΨ1Ψ2Ψ1xΨ2]=const.\displaystyle J_{12}=\frac{i}{2m}\left[\partial_{x}\Psi_{1}^{\star}\Psi_{2}-\Psi_{1}^{\star}\partial_{x}\Psi_{2}\right]=\text{const.} J12=ψ¯1γxψ2\displaystyle J_{12}=\bar{\psi}_{1}\gamma^{x}\psi_{2} =const.\displaystyle=\text{const.}