Abstract
The derivation of the total induced current density vector field, in the presence of static and uniform magnetic and electric fields, is illustrated in a more clear and formally correct language together with a discussion on the charge-current conservation law not presented before for the spin–orbit coupling contribution. The theory here exposed turns out to be in fully agreement with the theory of Special Relativity and it is applicable to open-shell molecules in the presence of a nonvanishing spin orbit coupling. The discussion here exposed turns out to be accurately valid for a strictly central field due to the chosen approximation of the spin–orbit coupling Hamiltonian, but it is appropriate to deal correctly with molecular systems. The ab initio calculation of spin current densities has been implemented at both unrestricted Hartree–Fock and unrestricted DFT levels of theory. Some maps of spin currents on molecules of interest, i.e., the CH3 radical and the superoctazethrene molecule are also illustrated.
1. Introduction
The interaction of matter with electric and magnetic fields or both has always fascinated the scientific community. Particularly in the case of open shell systems, NMR and EPR spectroscopies1−3 are useful means to identify their structure being sometimes the existence of these species very short. However, the aim of the present paper is not to discuss about these spectroscopies but rather to analyze in detail the expressions obtained for the current density vector field induced by static and uniform magnetic and electric fields when the interaction between electrons spins and the applied fields is properly taken into account using the Foldy–Wouthuysen diagonalization of the Dirac Hamiltonian.4 A magnetic field induces current distributions in the electron cloud of a molecule. Some of these current distributions can be derived starting from the continuity equation, using the Schrödinger equation with an hydrodynamical approach,5 but contributions that arises from electron spin cannot be obtained in this way (being the spin operator self-adjoint) and different procedures have been proposed during the years to overcome this problem.2,3,6−9 Of these procedures we recall in particular the Gordon decomposition of the Dirac four-current6,10−12 that introduces a magnetization current, see eq 38 for definition, starting from a correct, relativistic spin 1/2 theory. However, this decomposition does not allow to obtain the spin–orbit coupling current derived first by Hodge and co-workers for the hydrogen atom9 adopting the Landau approach7 and then extended to deal with many electron systems.13 The aim of the present paper is to discuss about the continuity equation related to the total induced current density in the presence of a non vanishing one electron spin–orbit coupling interaction in open-shell systems.
2. Outline of Notation and Theoretical Methods
Within the Born–Oppenheimer (BO) approximation,14 for a molecule with n electrons
and N clamped nuclei, charge, mass, position, canonical
and angular momentum of the k-th electron are indicated,
in the configuration space, by – e, me, rk, ,
, k = 1, 2, ..., n, using boldface letters for electronic vector operators.
Analogous quantities for nuclei are Zae, Ma, Ra, etc., for a = 1, 2, ..., N. Throughout this work, SI units are used and standard tensor formalism
is employed, e.g., the Einstein convention of implicit summation over
two repeated Greek indices is applied. The third-rank Levi–Civita
skew-tensor is indicated by ϵαβγ. The imaginary unit is represented by a Roman i. Let us introduce
the general definition of n-electron probability
density matrix functions for a stationary state wave function Ψ(X)
![]() |
1 |
of electronic space-spin coordinates xk = rk ⊗ ηk, k = 1, 2, ..., n, where
![]() |
2 |
By integrating over the spin variable η1, a spatial probability density matrix is obtained
![]() |
3 |
Similarly, the spin-density matrix is defined as
![]() |
4 |
with the α component of the spin operator
equal to
![]() |
5 |
and σα the Pauli matrices
![]() |
6 |
The diagonal elements of the density matrix, eq 1, γ(r) = γ(r; r), give the electronic charge distribution of the state ρ(r) = −eγ(r), and the diagonal elements of eq 4 give the spin density, described by the axial vector Qα(r) = Qα(r; r). For the reference state Ψa the probability density becomes
![]() |
7 |
Our starting point is the generalized Foldy–Wouthuysen Hamiltonian,4 within the Breit–Pauli picture, in the Born–Oppenheimer approximation,14 for molecular interacting systems with static and uniform magnetic and electric fields, i.e.,4,8,13,15
![]() |
8 |
In the previous equation primes mean
that, performing the double
summation, k ≠ j and a ≠ a′, μB is the Bohr magneton, g is the
electron spin g-factor, is the electric field operator defined
as
![]() |
9 |
and is the electron mechanical momentum operator
defined to be
![]() |
10 |
It is worth noticing that in Hamiltonian
(eq 8) we have removed
the electron rest mass energy,
so that the zero of our energy axis agrees with the conventional nonrelativistic
one. Other high order terms have been not taken into account. The
reason to avoid such terms, like the mass-velocity correction, as
defined in the Cowan–Griffin Hamiltonian,16 is that it makes the Hamiltonian unbounded from below,17,18 indeed it converges only for , whereas the spectrum of operator
is (0, + ∞).19 The leading relativistic corrections are futile
because it is well-known that the expectation values
, n ≥ 5, diverge
for S states of the hydrogen atom and, in general, for the ground
state of any atomic or molecular system.20−22 In order to
take into account these or other effects it is preferable to use a
fully relativistic 4-components approach being in principle simpler
to apply. An important aspect of the discussion that will follow is
that spin currents as derived in the present manuscript can be implemented
both in a traditional (non perturbed) open-shell Hartree–Fock
and open-shell DFT calculation because contributions as the spin–orbit
coupling and the electron spin Zeeman Hamiltonians can be neglected
in a linear response approach calculation at first insight. The implementation,
here reported, seems to be extendable in a current density functional
theory picture too.23,24
3. Semi-Relativistic Quantum Mechanical Current Density
To show how the QM expression for the total many body current density J can be obtained, the Landau–Lifshitz approach7 is used which is based on relation
![]() |
11 |
To use this idea for a QM system, Landau and Lifshitz argued that the classical Hamiltonian Hc is to be identified with the expectation value of the QM Hamiltonian according to
![]() |
12 |
Due to eq 11, only terms containing the vector potential A (or the magnetic field B) must be taken into account. For a complete derivation of the Landau–Lifshitz approach, see ref (9). Looking at the interaction Hamiltonian 8, one can see that the vector potential appears only in three terms. Then, defining
![]() |
13 |
and using eq 1 we can focus only on the term8
![]() |
14 |
The use of the reduced Hamiltonian 13 and the reason to neglect
contributions (such as terms
like, for example, Fermi contact, spin dipolar interaction, Darwin,
and all the others not written in eq 8, see refs (8, 25) for a complete description), except the two-electron spin–orbit
coupling term, is that they make no direct contribution to the induced
electronic current density vector field in a finite field approach.
That is not true in case of a perturbative approach.2,26 Furthermore, being in the Born–Oppenheimer approximation,
it is clear that the nuclear Zeeman interaction and the term representing
the nuclear magnetic moment acting on the moving electrons25 cannot be taken into account to derive an expression
for the approximate total induced current density being nuclei clamped
in space. The reason why the two-electron spin–orbit coupling
contribution has been neglected in Hamiltonian 8 is that the author wants to keep as simple
as possible the derivations and the discussion carried out in the
present manuscript, being in line with the original paper by Foldy
and Wouthuysen.4 The use of this approximate
form of the spin–orbit operator turns out to be accurately
valid for a strictly central field.25 Despite
this, that is not exactly true in the case of molecular systems (as
it is the case with atomic systems), the conclusions reported at the
end of the present manuscript seem to fully reflect those obtained
using a fully relativistic 4-components approach in molecules where
Special Relativity plays a fundamental role at least at qualitative
level.27 For this reason the model here
proposed seems to be appropriate to deal correctly with molecular
systems. When spin-effects are not taken into account in , it is possible to complete the integration
over spin variable and write the following equation
![]() |
15 |
recovering the spinless density (eq 3) introduced before. This is the case of the first term in eq 13. To obtain the current density vector J from eq 11, we need to calculate the variation δHc with respect to an infinitesimal change of the vector potential9,13
![]() |
16 |
Compared to ref (13), a more clear and formally
correct derivation of the many-body induced
current density will be given here. First, let us consider the term . By expanding
we have
![]() |
17 |
Substituting this expression and the definition 3 in eq 15, we can write
![]() |
18 |
because
![]() |
19 |
and a second order variation δA·δA is not considered being vanishing small at first-order approximation. From the previous equations, it follows that
![]() |
20 |
Let us consider now the term · (δAΨ). If we apply the vector identity
![]() |
21 |
with f = Ψ and V = δA, we obtain
![]() |
22 |
Multiplying the last equation on the left by Ψ*, one can write
![]() |
23 |
The term Ψ*Ψ·(δA), on the r.h.s of the previous equation, can be identified with f
·V, in
which f = Ψ*Ψ and V = δA, so using again
the identity 21, one
obtains
![]() |
24 |
Considering that
![]() |
25 |
eq 23 can be rearranged in the form
![]() |
26 |
Applying the divergence theorem for the first term on the r.h.s of the last identity and considering that the wave function goes to zero at infinity, we have
![]() |
27 |
from which relation 20 can be rewritten as
![]() |
28 |
Comparing this last equation with eq 11, it follows that the contribution given from the first term in eq 13 to the total induced current density vector is
![]() |
29 |
conventionally rewritten as
![]() |
30 |
with
![]() |
31 |
![]() |
32 |
A similar procedure can be applied to the spin Zeeman and to the one-electron spin orbit coupling Hamiltonians. Considering the spin Zeeman Hamiltonian first and substituting A(r) with eq 16, as before, we obtain that
![]() |
33 |
from which it is possible to rewrite
![]() |
34 |
Now taking into account the vector identity
![]() |
35 |
and applying the divergence theorem for the first term on the r.h.s of the previous equation, considering that the wave function goes to zero at infinity, we have
![]() |
36 |
from which relation 34 can be rewritten as
![]() |
37 |
Using eqs 11 and 4, we obtain from the Spin Zeeman Hamiltonian as contribution to the total induced current density vector the expression
![]() |
38 |
called also magnetization current in analogy with the one obtained in classical electrodynamics.9,28 Now, let us use relation 14 for the last term of eq 13, i.e. the one-electron spin–orbit coupling Hamiltonian, substituting A(r) with eq 16, we get
![]() |
39 |
from which it follows
![]() |
40 |
Using the vector identity
![]() |
41 |
we obtain
![]() |
42 |
that using eq 11 enable use to achieve as contribution to the total current density
![]() |
43 |
As approximation to the total induced many-body current density,13 for a generic open-shell system in the Born–Oppenheimer approximation and in the case of a strictly central field, (the two electron spin–orbit coupling interaction has been not taken into account) by collecting all the previous terms, it is possible to write
![]() |
44 |
The presence of other terms usually
included in the Hamiltonian,
like the Darwin correction for example or the Coulomb interaction
between charged particle, (i.e., terms without the vector potential A), do not change the results that have been
illustrated in eq 44 as discussed before. The current density defined in eq 44 is by definition gauge-invariant
only for an exact calculation or in the limit of a complete basis
set. This gauge dependence comes from the nonrelativistic part of
the current density, see eq 30.29 In the SI system, units of J are . As stated before, the results here proposed
are to be taken into account only in the presence of static and uniform
magnetic and electric fields, so dynamic currents are not considered.30−35 The Hamiltonian 13 describes the interaction with the intramolecular perturbation,
that is, the intrinsic magnetic dipoles μa = γaIa, expressed via the
magnetogyric ratio
and spin
, with na an integer, of nucleus a via the vector potential
, and with an external, spatially uniform
and time-independent magnetic field
![]() |
45 |
![]() |
46 |
![]() |
47 |
Due to the form of the vector potential here chosen, a classical Larmor contribution coming from the nuclear magnetic dipole is also expected, indeed26eq 32 can be rewritten as
![]() |
48 |
Here, according to definition 4, it is clear that
![]() |
49 |
In accordance with the Wigner–Eckart theorem, the spin densities are all the same except for a proportionality constant,3,8,25 so it is therefore expedient to introduce a reduced spin density scalar function common to all components of the multiplet
![]() |
50 |
with
![]() |
51 |
where γα(r) and γβ(r) represent the densities of spin-up and spin-down electrons,
respectively. Recently a topological analysis of the spin density
has been reported in literature.36 For
an open shell molecule, as we can see from the previous equations,
we have many spin–orbit coupling current densities as the number
of atoms contained, but due to the presence of , these currents are negligible except for
very heavy atoms where spin–orbit coupling plays a fundamental
role or in systems with high spin multiplicity in the nuclear proximity.13 The eq 44 for a closed shell system reduces to
![]() |
52 |
The continuity equation associated with the total current defined in eq 44 that is verified, in tensorial notation, is
![]() |
53 |
because the magnetization current has no divergence due to the presence of the curl in its definition, see eq 38, and the non relativistic current (eq 30) can also be obtained with an approach that adopts the Coulomb gauge. That ensures that its divergence is zero for an exact calculation. Now, let us consider the term defined in eq 53. If we use the vectorial relation
![]() |
54 |
given for two generic vectors X and Y, we have that
![]() |
55 |
Now the second term can be evaluated using the following Maxwell’s equation37
![]() |
56 |
in the case of a static magnetic field to be
![]() |
57 |
so we have
![]() |
58 |
Due to the scalar product between an axial vector and a polar one, i.e., the electric field generated by point charges and the magnetization current density vector field, we have38,39
![]() |
59 |
from which it follows that
![]() |
60 |
The condition ∇αJα(r) = 0 is fully satisfied only if the state functions are exact eigenfunctions of a model Hamiltonian and therefore satisfy the off-diagonal hypervirial theorem for the position operator, i.e., in HF, DFT, or Full-CI approaches40,41 as illustrated in Figure 1 for the divergence of first-order magnetically induced current density in the benzene molecule tending toward the complete basis set limit. This problem anyway, comes from the nonrelativistic current density vector, eq 30, because, as discussed before for spin contributions, see eqs 38 and 43, the charge-current conservation condition is satisfied for symmetry reasons related to the involved quantities, regardless electronic structure calculation method and the adopted basis set. Looking at the Larmor current induced by nuclear magnetic dipole moments, i.e. the last term of eq 48, it seems that a divergence of this current is always expected (also in the case of an exact calculation or in a complete basis set limit). That is not true because it can be shown that using Rayleigh–Schrödinger perturbation theory, at first order in the magnetic field given by nuclear magnetic dipoles, also a paramagnetic term can be obtained2
![]() |
61 |
with
![]() |
62 |
![]() |
63 |
![]() |
64 |
Figure 1.
Diverging color map of ∇αJα induced by a static magnetic field Bzϵ3 for the benzene molecule calculated on three different planes at BHandHLYP level of theory. The CPK (Corey-Pauling-Koltun) color scheme colors “atom” objects by the atom (element) type. From top to bottom, we have the four different basis sets adopted, i.e., aug-pcseg-0, aug-pcseg-1, aug-pcseg-2, and aug-pcseg-3 respectively. From left to right, we have the molecular plane, −0.5 and −1 au, respectively.
The sum of these paramagnetic and diamagnetic terms is gauge invariant and the integral conservation condition is satisfied in the case of a complete basis set and exact eigenfunctions of a model Hamiltonian.2
![]() |
65 |
The same seems to be expected in a finite field calculation at all order in the applied magnetic field. It is not in the scope of this paper to show a plot of the paramagnetic contribution given by nuclear magnetic dipoles. The condition ∇αJα = 0 is compatible with the true induced relativistic current density. To show this let us consider that the Dirac’s equation of an electron in an electromagnetic field reads42
![]() |
66 |
with the 4-spinor Ψ defined as
![]() |
67 |
and I4 that is the 4 × 4 identity matrix. With reference to the Landau–Lifshitz approach used before, now it is possible to define
![]() |
68 |
and then to obtain using again the relation (14) and substituting A(r) with eq 16
![]() |
69 |
From previous equation it follows that
![]() |
70 |
Using eq 11, the relativistic current density vector can be written as
![]() |
71 |
and the satisfied charge-current conservation condition is in the case of static and uniform magnetic and electric fields42
![]() |
72 |
4. Implementation of Induced Spin Current Densities in a Linear Response Approach
The theoretical formulation of the total induced current density vector field can be straightly implemented within the UHF and UKS frameworks.3,13,29 The first term on the rhs of eq 44 describes the system response to an external applied magnetic field. In this context an explicit expression for the first term on the rhs of (44) can be provided according to the well-known equations for the first order-induced current density using a CTOCD approach29,43−52 or using London orbitals12,53 to avoid the origin dependence of the calculated current density vector field, being the same a subobservable according to a definition proposed by Hirschfelder.54 For the Zeeman current density, with the understanding that the reference frame is always chosen so that the quantization axis coincides with the direction of applied magnetic field, we can define3
![]() |
73 |
Similarly, for the spin–orbit coupling current, we have
![]() |
74 |
with
![]() |
75 |
and
![]() |
76 |
where the density matrices are
![]() |
77 |
![]() |
78 |
defined using the Szabo and Ostlund notation.55 Orbital coefficients can be obtained from a Gaussian56 calculation or generated by the SYSMOIC suite of programs.29 The full procedure for the calculation of total induced current density vector field has been implemented in the freely available SYSMOIC program package.29 The implementation here reported seems to be easily extendable also to nonperturbative approaches24,57 when a magnetic field is applied only in the z direction in which the electron spin Zeeman Hamiltonian can be expressed as a scalar quantity, but this is not true for the one-electron spin–orbit coupling Hamiltonian.25 For the Larmor contribution given by the nuclear magnetic dipole μa
![]() |
79 |
the implementation at first order is straightforward, being
![]() |
80 |
and
![]() |
81 |
Note that μa is a vector with components equal to the maximum z-component expectation value of the magnetic dipole moment, for a given nucleus, in units of nuclear magnetons.
5. Calculation Details
Being in the case
of spin currents the continuity equation satisfied
for symmetry reasons, we have computed only the divergence of non
relativistic current density, eq 30, calculated at first order in the applied magnetic
field for the benzene molecule, prototypical of aromatic behavior
for a magnetic field applied in the z direction.
Thanks to its small size, very accurate computations have been carried
out, using the BHandHLYP functional,58 recently
shown to perform quite well,41 adopting
basis sets of contracted functions which include terms of high angular
momentum, taken from BSE. In particular, the aug-pcseg-X basis set
series (with X = 0, 1, 2, 3) has been adopted.59 BHandHLYP perturbed coefficients for r × and
operators have been computed using the
Gaussian 16 program package.56 Geometry
was first optimized using the same functional with aug-cc-pVTZ basis
set. In order to show some maps of induced spin currents, the CH3 radical in a doublet state and an open-shell graphene like
system, i.e., the superoctazethrene60 have
been taken into account. The geometries have been optimized with an
unrestricted B3LYP calculation adopting the 6-31G(d) basis set. The
keyword guess = mix has been used in the case of superoctazethrene
to account for the singlet open-shell character. Then single point
calculations made with the same functional and a 6-311+G(2d,p) basis
set adopting the 6d 10f keywords have been performed to obtain the
unperturbed coefficients needed for the evaluation of spin density
and spin currents using the SYSMOIC program facilities.29 All the electronic structure calculations have
been performed using the Gaussian 16 program package.56
6. Results and Discussion
Diverging color
maps of ∇αJα of first-order induced current density are illustrated
in Figure 1 for a magnetic
field applied along the z direction in the benzene
molecule using a CTOCD-DZ2 approach.47 As
we can observe, an increase in the basis set size reduces the divergence
of the current density vector field in all directions. Four maps of
induced spin currents for a magnetic field applied along the z-axis (that is considered as the quantization axis) on
the molecular plane are illustrated in Figures 2–5 for the CH3 radical and superoctazethrene
molecules in the D3h and C2h symmetry point groups respectively for the electron
Zeeman and the spin orbit coupling currents. From these maps it follows
that the topological structure of the two kind of spin currents is
not so different. A difference can be seen near the nuclear positions
where a divergence due to the point charge character of the electric
field is expected for the spin–orbit coupling current, as well
as a current cusp. Another difference is the fast decay of the spin
orbit coupling current, compared to the spin magnetization current,
observed moving away from nuclear positions due to the electric field
dependence as . A plot of the total spin current is not
useful due to the different magnitude of these currents not taken
into account in the maps here shown being the spin–orbit coupling
current smaller than the spin magnetization current roughly by a factor
. Looking at Figures 4 and 5 different tropicities
can be observed in both maps together with a clear evidence of the
regions in space where unpaired electrons have a higher probability
to be found. It is clear by inspection of equations and maps that
spin–orbit coupling (SOC) currents enhance the curvature and
give rise to a previously unobserved current cusps according to the
fully 4-components relativistic calculation as seen in ref (27). A plot of the Larmor
contributions given by the nuclear magnetic dipoles is also presented
in Figure 6 together
with its divergence in Figure 7.
Figure 2.
Magnetization current density vector field, eq 38, induced on the molecular plane in the CH3 radical by a static magnetic field Bzϵ3 considered to be coincident with the quantization axis represented with a streamlines map.
Figure 5.
Spin orbit coupling current density vector field, eq 43, on the molecular plane
induced
in the superoctazethrene molecule by a static magnetic field Bzϵ3 considered to be coincident with the quantization axis. Note that
to produce this map, in the implementation done in atomic units, the
factor has been not taken into account.
Figure 4.
Magnetization current density vector field, eq 38, on the molecular plane induced in the superoctazethrene molecule by a static magnetic field Bzϵ3 considered to be coincident with the quantization axis.
Figure 6.
Larmor current density vector field, given by nuclear
magnetic
dipoles, at −1 a.u. on the molecular plane induced in the superoctazethrene
molecule by a static magnetic field Bzϵ3. Note that to produce
this map, in the implementation done in atomic units, the ratio has been not taken into account.
Figure 7.
Divergence of Larmor current density vector field, given
by nuclear
magnetic dipoles, at −1 a.u. on the molecular plane induced
in the superoctazethrene molecule by a static magnetic field Bzϵ3. Note that to produce this map, in the implementation done in atomic
units, the ratio has been not taken into account.
Figure 3.
Spin orbit coupling current density vector field, eq 43, on the molecular plane
induced
in the CH3 radical by a static magnetic field Bzϵ3 considered
to be coincident with the quantization axis represented with a streamlines
map. Note that to produce this map, in the implementation done in
atomic units, the factor has been not taken into account.
7. Conclusions
In the present manuscript it has been illustrated that the introduction of a one-electron spin–orbit coupling contribution in the approximate total induced current density vector field, as defined in eq 44, does not change the charge-current conservation condition expected to be fulfilled (∇·J = 0) in the case of applied static and uniform magnetic and electric fields for an exact calculation or in the limit of a complete basis set. The divergence comes to be zero for the spin magnetization current eq 38 due to the presence of the curl. For the spin–orbit coupling current, eq 43, instead, the divergence is zero thanks to the scalar product between an axial vector and a polar one as discussed before.38,39 To illustrate that the divergence of the total induced current density vector field is not zero in approximate calculations some maps of this scalar quantity have been shown in a linear response approach using a CTOCD-DZ2 method to avoid the origin-dependence of first order induced current density. The condition ∇αJα = 0 is fully satisfied only if the state functions are exact eigenfunctions of a model Hamiltonian and therefore satisfy the off-diagonal hypervirial theorem for the position operator, i.e., in HF, DFT or Full-CI approaches.40,41 Some maps of spin currents have been illustrated for the CH3 radical and the superoctazethrene60 molecules. A contribution coming from the two-electron spin–orbit coupling interaction is also expected, as discussed in the main text, in the definition of the total many body induced current density vector field, due to the presence of the vector potential, but it is expected to be smaller than the others like the spin magnetization current, the nonrelativistic and the one-electron spin orbit coupling ones, being in heavy elements, the one-electron term of the spin–orbit coupling Hamiltonian the dominating part.61,62 Furthermore, neglecting the two-electron spin–orbit coupling contribution it is possible to implement currents, as defined in eq 44, in standard nonrelativistic UHF and UDFT calculations in a linear approach fashion or in a nonrelativistic current density functional theory picture.23,24 For the reasons here exposed and that the features of a full 4-components relativistic calculation are observed, it seems that the equations here reported can be applied to molecular systems where the strictly central field condition is lost. A plot of the Larmor contribution given by nuclear magnetic dipoles is also shown together with a map of its divergence different from zero also in a complete basis set limit. This divergence is erased by the divergence of the paramagnetic contribution as shown in terms of the integral conservation condition eq 65.
Acknowledgments
The author would like to thank Professor Riccardo Zanasi, Professor Guglielmo Monaco, and Professor Paolo Lazzeretti for helpful discussions. Financial support from FARB 2019 and FARB 2020 is gratefully acknowledged.
The author declares no competing financial interest.
Dedication
This paper is dedicated to Professor Paolo Lazzeretti on the occasion of his 80th birthday.
References
- Lazzeretti P. Ring currents. Prog. Nucl. Magn. Reson. Spectrosc. 2000, 36, 1–88. 10.1016/S0079-6565(99)00021-7. [DOI] [Google Scholar]
- Lazzeretti P. Gauge invariance of the nuclear spin/electron orbit interaction and NMR spectral parameters. J. Chem. Phys. 2012, 137, 074108. 10.1063/1.4745077. [DOI] [PubMed] [Google Scholar]
- Soncini A. Charge and Spin Currents in Open-Shell Molecules: A Unified Description of NMR and EPR Observables. J. Chem. Theory Comput. 2007, 3, 2243–2257. 10.1021/ct700169h. [DOI] [PubMed] [Google Scholar]
- Foldy L. L.; Wouthuysen S. A. On the Dirac Theory of Spin 1/2 Particles and Its Non-Relativistic Limit. Phys. Rev. 1950, 78, 29–36. 10.1103/PhysRev.78.29. [DOI] [Google Scholar]
- Madelung E. Quantentheorie in hydrodynamischer Form. Zeitschrift für Physik 1927, 40, 322–326. 10.1007/BF01400372. [DOI] [Google Scholar]
- Gordon W. Der Strom der Diracschen Elektronentheorie. Z. Physik 1928, 50, 630–632. 10.1007/BF01327881. [DOI] [Google Scholar]
- Landau L. D.; Lifshitz E. M.; Landau L. D.. Quantum mechanics: Non-relativistic theory, 3rd ed.; Course of Theoretical Physics Vol. 3; Landau L. D., Lifshitz E. M., Eds.; Elsevier: Singapore, 2007. [Google Scholar]
- McWeeny R.Methods of Molecular Quantum Mechanics, 2nd ed.; Theoretical Chemistry; Academic Press: London, 1992. [Google Scholar]
- Hodge W. B.; Migirditch S. V.; Kerr W. C. Electron spin and probability current density in quantum mechanics. American Journal of Physics 2014, 82, 681–690. 10.1119/1.4868094. [DOI] [Google Scholar]
- Bast R.; Jusélius J.; Saue T. 4-Component relativistic calculation of the magnetically induced current density in the group 15 heteroaromatic compounds. Chem. Phys. 2009, 356, 187–194. 10.1016/j.chemphys.2008.10.040. [DOI] [Google Scholar]
- Engel E.; Dreizler R. M.. Density Functional Theory: An Advanced Course; Theoretical and Mathematical Physics; Springer: Heidelberg, 2011. [Google Scholar]
- Sundholm D.; Dimitrova M.; Berger R. J. F. Current density and molecular magnetic properties. Chem. Commun. 2021, 57, 12362–12378. 10.1039/D1CC03350F. [DOI] [PubMed] [Google Scholar]
- Summa F. F.; Citro R. Many Body Current Density from Foldy–Wouthuysen Transformation of the Dirac–Coulomb Hamiltonian. Physchem 2022, 2, 96–107. 10.3390/physchem2020007. [DOI] [Google Scholar]
- Born M.; Oppenheimer R. Zur Quantentheorie der Molekeln. Ann. Phys. (Berlin, Ger.) 1927, 389, 457–484. 10.1002/andp.19273892002. [DOI] [Google Scholar]
- Helgaker T.; Coriani S.; Jørgensen P.; Kristensen K.; Olsen J.; Ruud K. Recent Advances in Wave Function-Based Methods of Molecular-Property Calculations. Chem. Rev. 2012, 112, 543–631. 10.1021/cr2002239. [DOI] [PubMed] [Google Scholar]
- Cowan R. D.; Griffin D. C. Approximate relativistic corrections to atomic radial wave functions. J. Opt. Soc. Am. 1976, 66, 1010. 10.1364/JOSA.66.001010. [DOI] [Google Scholar]
- Farazdel A.; Smith V. H. Invalidity of the ubiquitous mass-velocity operator in quasirelativistic theories. Int. J. Quantum Chem. 1986, 29, 311–314. 10.1002/qua.560290304. [DOI] [Google Scholar]
- Thaller B. Normal forms of an abstract Dirac operator and applications to scattering theory. Journal of Mathematical Physics 1988, 29, 249–257. 10.1063/1.528182. [DOI] [Google Scholar]
- Almlöf J.; Fægri K.; Grelland H. A variational approach to relativistic effects in LCAO calculations. Chem. Phys. Lett. 1985, 114, 53–57. 10.1016/0009-2614(85)85054-5. [DOI] [Google Scholar]
- Bethe H. A.; Salpeter E. E.. Quantum Mechanics of One- and Two-Electron Atoms; Springer: Berlin, 1957. [Google Scholar]
- Thakkar A. J. Relativistic kinetic energies and mass–velocity corrections in diatomic molecules. J. Chem. Phys. 1986, 85, 4509–4514. 10.1063/1.451771. [DOI] [Google Scholar]
- Ketley I.; Moss R. On the expectation values of relativistic corrections to the Hamiltonian. Mol. Phys. 1983, 49, 1289–1295. 10.1080/00268978300101941. [DOI] [Google Scholar]
- Vignale G. In Time-Dependent Density Functional Theory; Series Title: Lecture Notes in Physics; Beig R., Beiglböck W., Domcke W., Englert B.-G., Frisch U., Hänggi P., Hasinger G., Hepp K., Hillebrandt W., Imboden D., Jaffe R. L., Lipowsky R., Löhneysen H. v., Ojima I., Sornette D., Theisen S., Weise W., Wess J., Zittartz J., Marques M. A., Ullrich C. A., Nogueira F., Rubio A., Burke K., Gross E. K. U., Eds.; Springer: Berlin, 2006; Vol. 706; pp 75–91. [Google Scholar]
- Irons T. J. P.; Garner A.; Teale A. M. Topological Analysis of Magnetically Induced Current Densities in Strong Magnetic Fields Using Stagnation Graphs. Chemistry 2021, 3, 916–934. 10.3390/chemistry3030067. [DOI] [Google Scholar]
- McWeeny R.Spins in Chemistry: Current Chemical Concepts; Academic Press: New York, 1970. [Google Scholar]
- Lazzeretti P.Science and Technology of Atomic, Molecular, Condensed Matter & Biological Systems; Elsevier, 2013; Vol. 3; pp 209–243. [Google Scholar]
- Berger R. J. F.; Repisky M.; Komorovsky S. How does relativity affect magnetically induced currents?. Chem. Commun. 2015, 51, 13961–13963. 10.1039/C5CC05732A. [DOI] [PubMed] [Google Scholar]
- Panofsky W. K. H.; Phillips M.. Classical Electricity and Magnetism; Addison-Wesley Publishing Company, 1962. [Google Scholar]
- Monaco G.; Summa F. F.; Zanasi R. Program Package for the Calculation of Origin-Independent Electron Current Density and Derived Magnetic Properties in Molecular Systems. J. Chem. Inf. Model. 2021, 61, 270–283. 10.1021/acs.jcim.0c01136. [DOI] [PubMed] [Google Scholar]
- Lazzeretti P. Gauge invariance and origin independence of electronic charge density and current density induced by optical fields. J. Chem. Phys. 2018, 149, 154106. 10.1063/1.5052352. [DOI] [PubMed] [Google Scholar]
- Lazzeretti P. Continuity equations for electron charge densities and current densities induced in molecules by electric and magnetic fields. J. Chem. Phys. 2019, 151, 114108. 10.1063/1.5124250. [DOI] [PubMed] [Google Scholar]
- Summa F. F.; Monaco G.; Zanasi R.; Pelloni S.; Lazzeretti P. Electronic Currents Induced by Optical Fields and Rotatory Power Density in Chiral Molecules. Molecules 2021, 26, 4195. 10.3390/molecules26144195. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Summa F. F.; Lazzeretti P. Electronic Currents and Anapolar Response Induced in Molecules by Monochromatic Light. Chemistry 2021, 3, 1022–1036. 10.3390/chemistry3030073. [DOI] [Google Scholar]
- Summa F. F.; Monaco G.; Zanasi R.; Lazzeretti P. Dynamic toroidizability as ubiquitous property of atoms and molecules in optical electric fields. J. Chem. Phys. 2022, 156, 054106. 10.1063/5.0082731. [DOI] [PubMed] [Google Scholar]
- Summa F. F.; Monaco G.; Zanasi R.; Lazzeretti P. J. Chem. Phys. 2022, 156, 154105. 10.1063/5.0089605. [DOI] [PubMed] [Google Scholar]
- Bruno G.; Macetti G.; Lo Presti L.; Gatti C. Spin Density Topology. Molecules 2020, 25, 3537. 10.3390/molecules25153537. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jackson J. D.Classical Electrodynamics, 3rd ed.; Wiley: New York, 1999. [Google Scholar]
- Stewart A. M. Vector potential of the Coulomb gauge. Eur. J. Phys. 2003, 24, 519–524. 10.1088/0143-0807/24/5/308. [DOI] [Google Scholar]
- Stewart A. Longitudinal and transverse components of a vector field. Sri Lankan J. Phys. 2012, 12, 33. 10.4038/sljp.v12i0.3504. [DOI] [Google Scholar]
- Epstein S. T.The Variation Method in Quantum Chemistry; Physical Chemistry, A Series of Monographs Vol. 33; Academic Press: New York, 1974. [Google Scholar]
- Summa F. F.; Monaco G.; Lazzeretti P.; Zanasi R. Assessment of the Performance of DFT Functionals in the Fulfillment of Off-Diagonal Hypervirial Relationships. Phys. Chem. Chem. Phys. 2021, 23, 15268. 10.1039/D1CP01298C. [DOI] [PubMed] [Google Scholar]
- Greiner W.; Reinhardt J.. Field Quantization; Springer: Berlin, 1996. [Google Scholar]
- Keith T. A.; Bader R. F. Calculation of magnetic response properties using a continuous set of gauge transformations. Chem. Phys. Lett. 1993, 210, 223–231. 10.1016/0009-2614(93)89127-4. [DOI] [Google Scholar]
- Lazzeretti P.; Malagoli M.; Zanasi R. Computational approach to molecular magnetic properties by continuous transformation of the origin of the current density. Chem. Phys. Lett. 1994, 220, 299–304. 10.1016/0009-2614(94)00158-8. [DOI] [Google Scholar]
- Zanasi R.; Lazzeretti P.; Malagoli M.; Piccinini F. Molecular magnetic properties within continuous transformations of origin of the current density. J. Chem. Phys. 1995, 102, 7150. 10.1063/1.469109. [DOI] [Google Scholar]
- Lazzeretti P.; Zanasi R. Molecular magnetic properties via formal annihilation of paramagnetic contribution to electronic current density. Int. J. Quantum Chem. 1996, 60, 249–259. . [DOI] [Google Scholar]
- Fowler P.; Zanasi R.; Cadioli B.; Steiner E. Ring currents and magnetic properties of pyracylene. Chem. Phys. Lett. 1996, 251, 132–140. 10.1016/0009-2614(96)00120-0. [DOI] [Google Scholar]
- Zanasi R. Coupled Hartree-Fock calculations of molecular magnetic properties annihilating the transverse paramagnetic current density. J. Chem. Phys. 1996, 105, 1460. 10.1063/1.472008. [DOI] [Google Scholar]
- Steiner E.; Fowler P. W. Four- and two-electron rules for diatropic and paratropic ring currents in monocyclic π-systems. Chem. Commun. 2001, 2220–2221. 10.1039/b104847n. [DOI] [PubMed] [Google Scholar]
- Steiner E.; Fowler P. W. Patterns of Ring Currents in Conjugated Molecules: A Few-Electron Model Based on Orbital Contributions. J. Phys. Chem. A 2001, 105, 9553–9562. 10.1021/jp011955m. [DOI] [Google Scholar]
- Soncini A.; Lazzeretti P.; Zanasi R. Invariance of magnetic-field induced current density to a continuous transformation of the origin of the coordinate system. Chem. Phys. Lett. 2006, 421, 21–26. 10.1016/j.cplett.2006.01.036. [DOI] [Google Scholar]
- Monaco G.; Summa F. F.; Zanasi R. Atomic size adjusted calculation of the magnetically induced current density. Chem. Phys. Lett. 2020, 745, 137281. 10.1016/j.cplett.2020.137281. [DOI] [Google Scholar]
- Sundholm D.; Fliegl H.; Berger R. J. Calculations of magnetically induced current densities: theory and applications. WIREs Comput. Mol. Sci. 2016, 6, 639–678. 10.1002/wcms.1270. [DOI] [Google Scholar]
- Hirschfelder J. O. Quantum mechanical equations of change. I. J. Chem. Phys. 1978, 68, 5151–5162. 10.1063/1.435635. [DOI] [Google Scholar]
- Szabo A.; Ostlund N. S.. Modern Quantum Chemistry: Introduction to Advanced Electronic Structure Theory; Dover Publications: Mineola, N.Y, 1996. [Google Scholar]
- Frisch M. J.; Trucks G. W.; Schlegel H. B.; Scuseria G. E.; Robb M. A.; Cheeseman J. R.; Scalmani G.; Barone V.; Petersson G. A.; Nakatsuji H.; Li X.; Caricato M.; Marenich A. V.; Bloino J.; Janesko B. G.; Gomperts R.; Mennucci B.; Hratchian H. P.; Ortiz J. V.; Izmaylov A. F.; Sonnenberg J. L.; Williams-Young D.; Ding F.; Lipparini F.; Egidi F.; Goings J.; Peng B.; Petrone A.; Henderson T.; Ranasinghe D.; Zakrzewski V. G.; Gao J.; Rega N.; Zheng G.; Liang W.; Hada M.; Ehara M.; Toyota K.; Fukuda R.; Hasegawa J.; Ishida M.; Nakajima T.; Honda Y.; Kitao O.; Nakai H.; Vreven T.; Throssell K.; Montgomery J. A. Jr.; Peralta J. E.; Ogliaro F.; Bearpark M. J.; Heyd J. J.; Brothers E. N.; Kudin K. N.; Staroverov V. N.; Keith T. A.; Kobayashi R.; Normand J.; Raghavachari K.; Rendell A. P.; Burant J. C.; Iyengar S. S.; Tomasi J.; Cossi M.; Millam J. M.; Klene M.; Adamo C.; Cammi R.; Ochterski J. W.; Martin R. L.; Morokuma K.; Farkas O.; Foresman J. B.; Fox D. J.. Gaussian16 Revision C.01. 2016; Gaussian Inc. Wallingford CT.
- Tellgren E. I.; Teale A. M.; Furness J. W.; Lange K. K.; Ekström U.; Helgaker T. Non-perturbative calculation of molecular magnetic properties within current-density functional theory. J. Chem. Phys. 2014, 140, 034101. 10.1063/1.4861427. [DOI] [PubMed] [Google Scholar]
- Becke A. D. A new mixing of Hartree-Fock and local density-functional theories. J. Chem. Phys. 1993, 98, 1372–1377. 10.1063/1.464304. [DOI] [Google Scholar]
- Jensen F. Segmented Contracted Basis Sets Optimized for Nuclear Magnetic Shielding. J. Chem. Theory Comput. 2015, 11, 132–138. 10.1021/ct5009526. [DOI] [PubMed] [Google Scholar]
- Zeng W.; Gopalakrishna T. Y.; Phan H.; Tanaka T.; Herng T. S.; Ding J.; Osuka A.; Wu J. Superoctazethrene: An Open-Shell Graphene-like Molecule Possessing Large Diradical Character but Still with Reasonable Stability. J. Am. Chem. Soc. 2018, 140, 14054–14058. 10.1021/jacs.8b09075. [DOI] [PubMed] [Google Scholar]
- Kiyonaga H.; Morihashi K.; Kikuchi O. Calculation of contributions of one- and two-electron spin-orbit coupling terms to the parity-violating energy shifts for amino acids and helical alkanes. J. Chem. Phys. 1998, 108, 2041–2043. 10.1063/1.475583. [DOI] [Google Scholar]
- Marian C. M. Spin-orbit coupling and intersystem crossing in molecules: Spin-orbit coupling. WIREs Comput. Mol. Sci. 2012, 2, 187–203. 10.1002/wcms.83. [DOI] [Google Scholar]