1. Introduction
In nuclear magnetic resonance (NMR) measurements of a nuclear spin, there is a difference between the resonance frequency of that spin when it is in a metal from when it is in vacuum or in an insulator. This is known as the Knight shift [
1]. Although the temperature
T dependence of the Knight shift in a superconductor
has long been considered to be a probe of the spin state of the paired electrons [
2,
3,
4], the only theoretical basis for that assumption was the 1958 Yosida model that assumed the probed nuclear spins could be entirely neglected [
5], and that the only quantity of interest was the
T dependence of the zero-magnetic-field
limit of the electron spin susceptibility of an isotropic and uncorrelated Type-I superconductor [
5]. This led for a BCS singlet-pair-spin superconductor to
, where
,
,
is one-half the BCS energy gap, and we set
, so that
as
, unlike most experimental results [
5].
For isotropic Type-I superconductors in the Meissner state, crushing the sample to a powder of crystallites the cross-sections of which were less than the magnetic penetration depth was usually found to provide a reasonable method for that conventional theory to be applicable [
2,
3,
4]. In the first years following the BCS theory, a few elemental transition metal superconductors were found to behave somewhat differently, as
did not vanish as
[
6], and it was thought that surface impurity spin-orbit scattering could explain the near-cancelation of
in transition metals [
6,
7]. However, surface impurity spin-orbit scattering could not explain the observed non-vanishing
results observed in clean materials. It is now understood that there is also a component to the Knight shift due to the orbital motion of the electrons in a superconductor, and for an anisotropic superconductor, this orbital contribution to the Knight shift depends upon the magnetic induction
direction.
There have since been many examples of unexplained behaviors of the Knight shift in exotic superconductors. Since one possibility of a
T-independent Knight shift result would be a parallel-spin, triplet pair-spin superconducting state, the use of the Knight shift has been considered to be a principal tool for the identification of a triplet pair-spin state. Some examples of triplet-pair-spin or some other types of exotic behavior claimed to exist in unusual materials based upon the unconventional Knight shift
T-dependence are listed in the bibliography [
8,
9,
10,
11,
12,
13,
14]. However, one of those materials was a quasi-one-dimensional organic superconductor [
10,
11], some examples of which often exhibit spin-density waves [
15], and another was the very dirty sodium cobaltate hydrate material [
12,
13,
14]. In the latter example, the upper critical field parallel to that layered compound is Pauli-limited, which normally only occurs when the magnetic field breaks the oppositely-oriented pair spins [
16,
17]. Since dirt drastically suppresses
p-wave superconductivity [
18], the sodium cobaltate hydrate Knight shift results, if correct, are likely to arise from some other mechanism.
Moreover, in highly anisotropic Type-II superconductors, such as the cuprates and heavy fermion materials, other significant breakdowns in the Yosida theory have been found to exist. In the first Knight shift measurements on the cuprate YBa
2Cu
3O
, Bennett et al. found that the Yosida theory appeared to work for the
63Cu spins in the CuO chains for all applied
directions, although the orbital contributions are different for each of those three orthogonal
directions, and it also appeared to work well for the
63Cu spins in the CuO
2 planes when the strong constant
was applied parallel to the CuO
2 layers. However, when
was applied normal to the CuO
2 layers, no
T dependence to the Knight shift was observed in that cuprate [
19]. This result was later described by Slichter as possibly being due to a “fortuitous” cancelation of the effect from an isolated planar
63Cu spin by its interaction with its near-neighbor planar
63Cu spins [
20]. Subsequently, in a number of layered correlated superconductors, the
T dependence of the Knight shift probes of the nuclear spins in the layers with the field applied normal to the layers has been observed to vary strongly with field strength, approaching a constant
in the large normal field strength limit, as first observed by Bennett et al. [
19,
20,
21,
22,
23,
24].
Especially in the case of Sr
2RuO
4, numerous Knight shift measurements of the
17O,
99Ru,
101Ru, and
87Sr have all led to temperature-independent Knight shift measurements [
25,
26,
27], as did polarized neutron scattering experiments [
28]. These experiments were all interpreted as evidence for a parallel-spin pair state in that material. However, several upper critical field measurements with the field parallel to the layers showed strong Pauli limiting effects [
29,
30], which is inconsistent with a parallel-spin pair state [
31,
32]. In addition, the fact that
T-independent Knight shift measurements were obtained for the field both parallel and perpendicular to the RuO
2 layers is incompatible with any of the crystal point-group-compatible
p-wave states. Thus, the only way for a
T-independent Knight shift to legitimately arise from a parallel-pair-spin state in both field directions is for the
-vector (the vector describing the components of the three triplet spin states) to rotate with the magnetic field [
33,
34]. This argument was used to show that while the upper critical field of Sr
2RuO
4 is strongly Pauli limited for the field applied parallel to the layers, it could possibly be consistent with one or more
p-wave helical states, provided that the
-vector is allowed to rotate freely with the magnetic field direction [
32]. This means that spin-orbit coupling with the lattice would have to be negligible. However, there is strong evidence that spin-orbit coupling in Sr
2RuO
4 is very strong at some points on the Fermi surface, ruling out such
-vector rotation possibilities [
35]. More worrisome for the Knight shift measurement results is the fact that carefully performed scanning tunneling measurements of the electronic density of states provided very strong evidence of a nodeless superconducting order parameter orbital symmetry in Sr
2RuO
4 [
17,
36], consistent with a nearly isotropic gap function that is essentially identical on all three of its Fermi surfaces. Since the theories behind the Pauli limiting effects and the BCS gap density of states are very well established, but the Knight shift measurement interpretations rely entirely on the complete neglect of the probed nuclear spins, the development of a microscopic theory of the
T dependence of the Knight shift in anisotropic and correlated Type-II superconductors is sorely needed.
We further note that the time dependence of a spin-1/2 particle in a classic magnetic resonance experiment is now a textbook example of an exactly soluble first quantization quantum mechanics problem giving rise to a Berry phase [
37,
38]. In that case, the Berry, or geometric, phase is a combination of the resonance profile with the frequency of the oscillatory transverse applied magnetic field. In higher spin
I systems, there are
combinations of those two quantities, giving rise to a multiplet of Berry phases, as discussed in the following. Note that the probed nuclear spins of Sr
2RuO
4 are either 5/2 or 9/2. Since nothing was known about the Berry phase in 1958, its possible implications for the interpretation of Knight shift measurements have been generally and perhaps completely ignored in the literature.
In fairness to the pioneering work of Yosida [
5], there have been a few cases in which a complete lack of any
T-dependence to the Knight shift has been confirmed by other experiments consistent with a parallel-pair-spin superconducting state [
39,
40,
41]. These are for the uranium-based compounds UCoGe and UPt
3, for which the
T-independent Knight shift in UCoGe is in agreement with the general assessment of the upper critical field and muon depolarization experiments [
18,
40]. In UPt
3, the seeming incompatibility of the Knight shift and the upper critical field appears to have been resolved by polarized neutron diffraction experiments [
41], favoring a parallel-spin pair state in all three superconducting phases. In the ferromagnetic superconductors UGe
2, UCoGe, and URhGe, the weak Ising-like ferromagnetism appears to allow for a parallel-spin,
p-wave superconducting order parameter in the plane perpendicular to the ferromagnetism, but the Knight shift measurements have not yet been made on URhGe and UGe
2, the latter of which is only superconducting under pressure. In these three ferromagnetic superconductors, there is at least a plausible mechanism for a parallel-spin pair superconducting state, and in URhGe the upper critical field fits the predictions for all three crystal axis directions of a parallel-spin
p-wave polar state fixed to the crystal
a-axis direction normal to the
c-axis Ising ferromagnetic order [
40,
42], and there is a reentrant, high field phase that violates the Pauli limit by a factor of 20 [
40]. In order to obtain further evidence that the classic Yosida interpretation of a
T-independent
can correctly imply a parallel-spin superconducting state, we urge that
73Ge, with a strong nuclear moment, (or possibly
103Rh, with a much weaker nuclear moment)
measurements on URhGe be carefully performed in the low-field superconducting phase.
2. The Model
The first microscopic model of the Knight shift at
in anisotropic and correlated metals was recently presented by Hall and Klemm [
43]. This model assumed that the applied magnetic fields probe the nuclear spins, and the spins of the electrons orbiting the nucleus interact with the nucleus via the hyperfine interaction in the form of a diagonal
tensor with two distinct components
. The assumption
was made to simplify the calculations, as discussed in more detail in the following. After interacting with the nuclear spins, the orbital electrons can be excited into one of multiple bands, each of which was assumed to have an ellipsoidal Fermi surface of arbitrary anisotropy and shape. The orbital motion of the electrons in each of these bands was constrained by the strong, time-independent part
of the magnetic induction
to be in Landau levels, and the electron spins also could interact weakly with
. It was found that the self-energy due to
led to the Knight shift, and that due to
led to the first formulas for the linewidth changes associated with the Knight shift at
. However, since those calculations were made at
, they could not be used to probe the superconducting state. In the following, a method is proposed to do so.
Following Haug and Jauho [
44], we write the Hamiltonian as
, where
is the time-independent part and
is the time-dependent part due to the oscillatory (or pulsed) magnetic field transverse to the constant applied magnetic field
, and the time-independent part consists of the simple (or exactly soluble) part
and the interaction part
that involves the interactions between the particles that must be treated perturbatively. In the case at hand, there are four types of particles: (1) the nuclear spins probed in the NMR experiment, which are assumed to have the general spin
with
substates denoted
; (2) the local orbital electrons surrounding each of the nuclei probed in the NMR experiment; (3) the conduction electrons or holes that propagate from the local nuclei throughout the metal/superconductor; and (4) the superconducting Cooper pairs of electrons or holes. We note that complicated materials such as Sr
2RuO
4 contain multiple Fermi surfaces, which can be a mix of electron and hole Fermi surfaces. In this model, we do not account for competing ferromagnetism or charge-density wave (CDW) or spin-density wave (SDW) formation, at least one of which is normally present in the transition metal dichalcogenides, the organic layered superconductors, the cuprates, the iron pnictides, and the ferromagnetic superconductors. Such competing effects will be the subjects of future studies.
2.1. The Simple Hamiltonian
Since in an NMR experiment, the applied magnetic field can be applied in any direction, we assume the resulting constant magnetic induction
with respect to the crystalline Cartesian
axes. We then quantize the spins along
. We thus write
where
where
is the electronic charge,
creates a nucleus of spin
I in the subspin state
at the atomic position
i,
creates an electron with energy
and spin-1/2 eigenstates indexed by
orbiting that nucleus at site
i, where
is nominally its weak-spin-orbit local electron orbital quantum number set [or its fully relativistic set
],
creates an electron or hole with spin eigenstate
at position
in the
conduction band,
,
,
are respectively the Zeeman energies for the probed nucleus, local orbital electrons, and conduction electrons, respectively,
is the nuclear magneton for the probed nucleus (the value of which can be positive or negative),
is the Bohr magneton,
defines the quantization axis direction for the anisotropic but assumed diagonal
tensor in the
of the
conduction bands with effective mass
in the
spatial direction,
is the
component of the magnetic vector potential at the position
of the conduction electron in the
band,
is the magnetic induction that is independent of
j,
, and the time
t,
, and we set
. Here we use the previous notation [
43], but rearrange the terms in the overall Hamiltonian in order to properly take account of both the
t and
T dependencies essential for probing the superconducting state. We note that for integer or half-integer
I, the nuclei would normally be expected to obey Bose-Einstein or Fermi-Dirac statistics, but since different nuclei correspond to different atoms and do not come in contact with one another, that statistics is not expected to be an important feature of the Knight shift. Equation (
1) is the extension to arbitrary nuclear spin
I of the bare Hamiltonian studied previously, except that
was the time independent part of
[
43]. We note that for a diagonal
tensor,
As a starting point, we assume
is uniform in the probed material, but when the material goes into the superconducting state, and
is in an arbitrary direction with respect to the crystal axes, this is only true at the upper critical field
above which the superconductor becomes a normal metal [
17,
45]. However, in the mixed state for which the time-independent part of the applied magnetic field
satisfies
, if
is along a crystal axis, the direction of
is the same as the direction of
[
17,
46].
2.2. The Time-Independent Interaction Hamiltonian
We write the time-independent interaction part
of the Hamiltonian as
where
where
, and depending upon what is calculated, the superconducting pairing interaction may be written either in real space as
or in momentum space as
Although it appears at first sight to be easier to extend the calculation of the Knight shift into the BCS superconducting state by using
in order to include the Zeeman terms, we have included the momentum-space pairing interaction
for
p-wave superconductors in magnetic fields [
18,
42], for which the simplest single-band parallel-spin pairing interaction
[
18], and a modification of
more naturally treats the pairing of conduction electrons (or holes) in the presence of a strong
.
We note that
present in
corresponds to the correct matrix elements for raising and lowering the
value and also corresponds to our description of the spin-1/2 electron spins [
43]. Of course,
and
are restricted by
. The first three of these terms were presented previously [
43], except for a slightly different normalization factor proportional to
, and respectively represent the hyperfine interaction between the nuclear and surrounding orbital electrons, the effective local electron correlation interaction, and the effective Anderson interaction that allows an orbital electron to leave a local atomic site and jump into a conduction band [
47]. The last term
is responsible for superconducting pairing, and in the form presented allows for pairing between electrons or holes in different bands and with either the same (
) or different (
) spins. In most superconductors, interband pairing is generally considered to be less important than is intraband pairing, but such complications might be important in cases such as Sr
2RuO
4, for which two of the bands are nearly identical. For standard BCS pairing, we would have
at least in the standard approximation. For parallel-spin
p-wave superconductors, one cannot assume the paired electrons are at the same location, but different approximations have been found to give reliable results for the upper critical induction in ferromagnetic superconductors [
18,
32,
42,
48,
49].
2.3. The Time-Dependent Hamiltonian
The crucial part of a magnetic resonance experiment arises from the time-dependent field transverse to the stronger constant magnetic field. In a conventional NMR experiment, the time-dependent induction
oscillates in the plane normal to the strong, constant magnetic induction
. For
, one may then write
, where
and
in the same Cartesian coordinates, and in order not to get confused with the time contours, we may choose
to be along
. This is the classic way to obtain a resonance in the power spectrum associated with flipping an electron or proton spin from up to down, or in a spin
I nucleus, to obtain a regular pattern of resonance frequencies associated with changes in the multiple Zeeman-like nuclear spin levels. Since one generally takes
, this classic case is generally adiabatic [
37,
38], and is the simplest case to treat analytically.
For the above classic NMR case of a single angular frequency
in
, we then have
where
,
, and
, and
is given by Equation (
11) [
43].
3. The Keldysh Contours
Following Haug and Jauho [
44] and with regard to the contours, Rammer and Smith [
50], we may treat the time and temperature dependence of the particles together in the same formulas, as long as we properly order the time integrations around the appropriate contours. When there is only one type of particle, which we take to be a fermion, the fields at the three-dimensional positions
and
evolve in time according to the simple Hamiltonian
,
with its density matrix also involving only
(and the number operator
in the grand canonical ensemble statistics),
and the Green function is given by the two contour integration paths
C and
sketched in
Figure 1,
where the operators in
and
evolve in time via the easily soluble Hamiltonian
. The Greek letter
implies that one needs to consider it as being just above or just below the real axis until the contours merge into one. Roman lettering (
t) indicates the integrals are on the real axis.
For the case of the time-dependent Hamiltonian
making adiabatic changes, as in the case considered here, the two contours
C and
respectively shown in
Figure 1a,b merge into contour
C shown in
Figure 1a. We use the standard short-hand notation
where the particle type, its position, and its energy are still undefined. In order to treat the various time orderings on the contour
C, we define the following in the Heisenberg representation for the full Hamiltonian
,
where the ordinary time-ordering operator
T and inverse-time-ordering operator
describe opposite directions in time, as sketched by lines
and
in
Figure 2. We note that
, so only three of these Green functions are linearly independent.
Here we need to describe three particles, all of which are effectively fermions.
3.1. Bare Nuclear Contour Green Functions
We first consider the nuclei, which are assumed not to interact with one another, as they are fixed in the crystalline locations, which if there is more than one isotope of a particular type with spin
I, may be at a random selection of crystalline sites. In the presence of the constant magnetic induction
, it can be in any one of the
manifold of nuclear Zeeman states, but because each of these local states at the probed nuclear site
i can be either unoccupied or singly occupied, this manifold of local nuclear spin states is precisely that of a fermion with
states. Its occupancy in the local state
on site
i in the grand canonical ensemble is therefore easily seen to be
where
and
is the nuclear chemical potential. We then have for the bare nuclear Green functions with
,
and
are constructed from these according to Equations (25) and (26). There are only three distinct bare neutron Green functions. This is also true when interactions are included [
44]. Although it is somewhat surprising that the nuclear occupation density has the Fermi function form even for integral spin
I, this is due to the nuclear Zeeman magnetic level occupancy being either 0 or 1 for each level on a given probed nuclear site.
3.2. Bare Orbital Electron Contour Green Functions
For the surrounding orbital electrons, we assume that the magnetic induction
is sufficiently weak that it does not change the electronic structure of the orbital electrons or lead to transitions between the orbital electron states and energy levels. Thus, we assume that it only interacts with the orbital electron spins. We note that this is expected to be a good approximation, as the total charge of the nucleus plus its orbital electrons is on the order of one electron charge (for an ion), and the mass of the ion is so large that any Landau levels describing the orbital electrons and their central nucleus is completely negligible in comparison with the Landau levels of the conduction electrons. The only point then to consider for the interaction of
with the orbital electrons is that there can be either 0, 1, or 2 electrons in a given orbital energy
, and two possible magnetic energies for up and down spins. Hence, it is elementary to show that the average orbital electron occupation number in the grand canonical ensemble is
where
is the orbital electron chemical potential. It is then easy to show that the bare orbital electron Green functions are
and the time-ordered and inverse-time-ordered bare orbital electron Green functions are obtained analogously using Equations (25) and (26). Only three of the bare orbital electron Green functions are linearly independent. This is also true when interactions are included [
44].
3.3. Bare Conduction Electron Contour Green Functions
In a normal metal (above the superconducting transition of all superconductors including the cuprates and the record high transition temperature superconductor hydrogen sulfide, which probably transforms to H
3S under the 155 GPa pressure that causes it to become superconducting at 203 K [
51]), the conduction electrons or holes propagate throughout the metal with wave vectors on or nearly on one or more Fermi surfaces. Both the spins and the charges of the conduction electrons interact with
, the spins via the Zeeman interaction and the charges couple to the magnetic vector potential, leading to Landau orbits. Here we assume each of these potentially multiple Fermi surfaces has an ellipsoidal shape, but the shapes and orientations of each of the Fermi surfaces can be different from one another.
We first use the Klemm-Clem transformations to transform each of the ellipsoidal conduction electron band dispersions into spherical forms [
43,
45]. For each ellipsoidal band, the anisotropic scale transformation that preserves the Maxwell equation
transforms the elliptical Fermi surface into a spherical one, but rotates the transformed induction differently in each band. Then, one rotates these bands so that the rotated induction is along the
z direction in each band [
43]. In the
jth band, the conduction electrons behave as free particles with wave vector
along the transformed
direction, but propagate in Landau orbits indexed by the harmonic oscillator quantum number
. Thus, we need to requantize the conduction electron fields as
.
We therefore rewrite the transformed
as
where
are the two-dimensional simple harmonic oscillator quantum numbers of the Landau orbits for band
j,
are the free-particle dispersions along the transformed induction direction,
for a diagonal
tensor describing the spins of the
jth conduction band, and
is the spatially-transformed Landau degeneracy for a single electron in the
jth band. We may then write the conduction electron occupation number as
where
is the chemical potential of the conduction electrons. We note that all of the bands that cross this conduction electron chemical potential make important contributions to the Knight shift.
The bare conduction electron Green functions can then be found to be
and the contour-ordered and inverse-contour-ordered bare conduction electron Green functions are obtained as in Equations (25) and (26), so that there are only three independent bare conduction electron Green functions.
Furthermore, due to the strong
, we also need to spatially transform all of the other terms in the Hamiltonian that contain the conduction electrons. Thus, we have [
43]
4. Transformations in Time of the Operators with the Bare Hamiltonian
In order to proceed with the perturbation expansions, we first need to transform the nuclear, orbital electronic, and conduction electronic operators in real time, using the bare Hamiltonian in Equation (
1). For the nuclear and orbital electronic operators, this is elementary. We have
and integrating the elementary differential equation, we immediately find
in order to use this in Equation (
16). The quantity
is instantly obtained from the Hermitian conjugate of Equation (
52) and letting
, and hence
, so that the time-transformed Equation (
16) becomes
Similarly, for the local orbital electron operators, we have
and
For the spatially-transformed conduction electron operators,
so that the time-transformed Equation (
44). becomes
We note that all three of these transformed Hamiltonians correspond to spin-dependent external field interactions, where the fields are
Then, we time transform the difficult (interaction) parts of the full Hamiltonian. The hyperfine and local electron-electron interactions are elementary to transform. We obtain
Of these, only the transverse (
) part of the hyperfine interaction picks up a time dependence. Before we time transform the remaining two interaction Hamiltonians, we first spatially transform the conduction electron operators in the presence of the magnetic field necessary for the NMR experiment. Then we rewrite
in terms of the spatially-transformed conduction electron fields,
which after time-transformation with respect to
and
becomes
The most important Hamiltonian for the Knight shift in a superconductor is the pairing interaction
, which in position space was written in Equation (
10). Since in a Knight shift measurement, the experimenter first measures the Knight shift in the applied field
and hence the induction
while the superconductor is in its normal (metallic) state, and then cools the material through its superconducting transition at
, it is clear that the correct formulation for the superconducting pairing interaction must be in momentum space, and more precisely, to account for the pairing of the electrons (or holes) while they are in Landau orbits in the normal state. We therefore first rewrite
in a fully spatially-transformed magnetic-induction-quantized form that allows for different pairing interactions, such as those giving rise to various types of spin-singlet and spin-triplet superconductors arising from a multiple-band metal. As a start to understand the orbital motion of the paired superconducting electrons (or holes), we first assume the standard approximation for the evaluation of the upper critical field
that the paired particles of combined charge
move together in Landau levels [
18,
42,
52,
53]. For a BCS superconductor for which
, there is no need to transform the wave vector dependence of the pairing interaction due to the Landau orbits formed by the strong applied field [
18]. Such pairing interactions will be considered elsewhere. Thus, we begin by considering only the simplest case of isotropic intraband pairing of equivalent strength in all of the bands, which after spatial transformation due to the magnetic induction may be written as
and transforming this in time using
, we have
which is independent of
t.
7. Discussion and Conclusions
We have outlined a procedure to obtain a microscopic theory of the Knight shift in an anisotropic Type-II superconductor. This was based upon the Hall-Klemm microscopic model of the effect at
[
43], for which multiple anisotropic conduction bands of ellipsoidal shapes were included. We considered the simplest magnetic resonance case of
with
and
with
oscillating at a single frequency
. For this simple case, the time changes to the system are adiabatic, so that the interaction Keldysh contour
shown in
Figure 1b effectively coincides with contour
C depicted in
Figure 1a, and the integrations can be analytically continued onto the real axis. The procedure can effectively treat any nuclear spin value
I. The conduction electrons (or holes) were quantized in Landau orbits in the applied field in the normal state, and the Hamiltonian for a generalized anisotropic, multiband BCS type-II superconductor was diagonalized, allowing for a full treatment of the superconducting state.
We emphasize that by quantizing the superconducting order parameter in the presence of a strong time-independent magnetic induction
, the energy spacings of the Landau orbits can depend strongly upon the direction of
. At very weak
values, the Landau levels primarily give rise to overall anisotropic constant backgrounds of
and
, with
being predominantly governed by the anisotropic Zeeman interactions and
. However, for sufficiently strong
values in anisotropic materials with layered or quasi-two-dimensional anisotropy, the spacings between the Landau energy levels depends strongly upon the direction of
, so that
could become independent of
T for
, as first observed for
in YBa
2Cu
3O
[
19,
20]. Such behavior could also arise for quasi-one-dimensional materials in all
directions, although to different degrees for
parallel and perpendicular to the most conducting crystal direction.
Since the crucial interaction for the Knight shift is the hyperfine interaction between the probed nuclei and their surrounding orbital electrons, the symmetry of this interaction can be very important. Generally, the hyperfine interaction can arise from the electrons in any of the orbital levels. For
s-orbitals, the Fermi contact term is important, but the induced-dipole induced-dipole interactions can arise from the nucleus of any spin for any spin
and its surrounding electrons in any orbital, and induced-quadrupole induced-quadrupole and higher order interactions can also occur for certain orbitals and nuclear spin values [
60,
61]. In the Hall-Klemm model [
43], the hyperfine interaction crucial for the Knight shift was taken to be diagonal in the spin representations of a lattice with tetragonal symmetry
. In that simple model, the
results indicated that the Knight shift arose from
, and the line width was modified by
. In more realistic examples of correlated and anisotropic materials, the hyperfine interaction would be represented by a symmetric matrix unless time-reversal symmetry-breaking interactions were present. Such matrices can be diagonalized by a set of rotations, but in complicated cases the quantization axes would not necessarily be the same as for the overall crystal structure. Such complications would mix the Knight shift and its linewidth, depending upon the direction of
.
As noted previously, in first quantization, an isolated nuclear spin wave function in an NMR experiment was found to have the form
where
is the nuclear resonance function and the constants
depend upon the initial conditions [
43]. Those authors found this form to hold for
, and in second quantization, up to
, so it is likely to hold for arbitrary
I. In the adiabatic regime, we have
[
37,
38], so that there will be a manifold of geometrical phases that will arise with higher
I values.
We remark that it is possible to generalize this treatment to more complicated
functions, such as a periodic function of square-wave or triangle-wave shape. This can be represented as a Fourier series, but if the primary angular frequency is
, terms of higher multiples
n of
can be present, some of which would violate the adiabatic requirement that they be much smaller than the Zeeman energy spacings. Hence, this experiment would make some amount of non-adiabatic changes that could drive the system out of thermal equilibrium, and the two contours
C and
shown in
Figure 1 and discussed above would not coincide, greatly complicating the analysis.