1. Introduction
Bethe’s wave-function was initially devised to diagonalise the Heisenberg[
1,
2] Hamiltonian for an infinite, one-dimensional (
) lattice and was later extended[
3,
4] to the
Hubbard Hamiltonian. However, since those analyses resort heavily to technical peculiarities, associated with the concerned Hamiltonians, Bethe’s wave-function could not be applied to any
model[
5], nor to other
Hamiltonians. Therefore this article is aimed at showing that Bethe’s wave-function gives access to the eigenspectrum of
every realistic
Hamiltonian. Then the groundstate energy will be compared with data[
6], obtained previously for the Hubbard Hamiltonian[
3,
4]. An additional comparison will be carried out with the data, resulting from the
correlated Fermi gas model, introduced recently[
7] to account for the properties of interacting electrons in normal metals. This method will be further applied to electrons interacting on neighbouring sites[
8], which was believed so far to lie out of the scope of Bethe’s wave-function[
1,
2,
3,
4].
Here is the outline : the proof of Bethe’s wave-function being a many-electron eigenstate is laid out in section I and the general expression of the corresponding eigenvalue is worked out; the two-body scattering is studied in section II for two different Hamiltonians, whereas section III will be concerned with the prominent role of a boundary condition; the groundstate energy, associated with each of the two mentioned models, is reckoned and compared with data, obtained by other methods, in sections IV and V; the various many-electron states, discussed here, are analysed comparatively in section VI; at last, the main results are summarised in the conclusion.
2. 1-Diagonalisation
Let us consider
of electrons moving on a
lattice, comprising an
even number
of atomic sites, labelled by the index
. The lattice parameter is taken equal to unity and each site can accommodate at most two electrons of opposite spin
, which implies
. The Hamiltonian
H, governing the electron motion, then reads
wherein
describe the one-electron hopping between first neighbours and the two-electron interaction, respectively. Hence
reads[
9]
wherein the sum is carried out over
,
with
and
are one-electron creation and annihilation operators on the Wannier[
9] state
with
referring to the no electron state. Then
t designates the hopping integral
with
standing for the electron-nucleus Coulomb potential. The resulting one-electron energy dispersion
reads[
9]
will be referred to as
for two electrons sitting on the same site, which characterises the Hubbard model, or on first neighbours. They read[
8]
wherein
and
J designate[
9] Coulomb and exchange integrals, respectively, and
refers to the electron-electron Coulomb potential.
The
n-electron states make up a Hilbert space of dimension
, subtended by the basis
with each
reading
. The
electron, having spin
, sits on the site
. The sites are ordered such that
, except in case of double occupancy, characterised by
and
. Let us introduce now a sequence of
n real numbers
and
defined as
with
being the position vector of site
i.
The group of permutations
P of
n objects is assumed to act on
as follows
wherein there is a one-to-one correspondence
between
and
. At last, the Hamiltonian
H is projected onto the subspace, subtended by
of
’s, so that every eigenvector of
H is found to read as a linear combination of the
’s
wherein
of
’s are complex numbers to be determined below. This last term on the right-hand side of Eq.(
3) is recognised to be identical to the expression of Bethe’s wave-function in Eq.(9) of [
4] and extends thereby its validity to
any Hamiltonian, as illustrated below for
. However noteworthy is that
may become bigger than
for
n close to
N, in which case the
’s are no longer linearly independent from one another.
The expression of the eigenvalue
, associated with
in Eq.(
3), will now be worked out for
in the case of the Hubbard Hamiltonian
. To that end, the particular
n-electron state
, sketched in Fig.1, is defined as
while assuming the boundary conditions
Figure 1.
Sketch of
, defined by Eqs.(
4,
8) for
, respectively, and sites
; dots and
signs refer to an empty site or one accommodating either one electron of spin 1 or
or two electrons of opposite spin
, respectively.
Figure 1.
Sketch of
, defined by Eqs.(
4,
8) for
, respectively, and sites
; dots and
signs refer to an empty site or one accommodating either one electron of spin 1 or
or two electrons of opposite spin
, respectively.
Thus
is seen to have the properties
Eq.(
6) then entails
with
defined in Eq.(
1). Extending this equation to every
and applying it further to the Schrödinger equation
, with
defined by Eq.(
3), yields finally the energy per site
as
Proceeding for
similarly as done for
,
, sketched in Fig.1, will be defined for
even as
Assuming again the conditions in Eq.(
5),
is inferred to have the properties
which entails for every
P
Taking advantage of
, as done in Eq.(
7), yields for
the same expression as already given in Eq.(
7). This latter expression of
is recognised to be identical to Eq.(11) of [
4]. Likewise, since it consists in a sum over one-particle energies, it is typical of a many-electron
scattering state, so that the many
bound electron states of
, addressed elsewhere[
8], are left out of the purview of this work. Nevertheless it is worth noticing that
could be achieved thanks to a careful, model dependent choice of
(see Eqs.(
4,
8)). Though
is independent from the two-electron coupling,
will prove below quite instrumental in assigning the
values.
But before doing that, it is in order to derive the expression of the energy
, valid for
, by replacing electrons by holes. To that end, we begin with recasting
as
Any
n-hole state
reads
, with
characterised by each of
N sites accommodating 2 electrons. Then substituting
to
in Eqs.(
4,
8) and
to
in Eqs.(
4,
6,
8,
9), and proceeding as done above for electrons yields the following expression for the energy per site of a
n-hole state
with
and
being the energy per site of
for
and
, respectively.
3. 2-Two-body scattering
The
coefficients will be assessed by analysing the two-body scattering, embodied by
. As a matter of fact, the latter turns out to be model dependent and to differ according to whether both electrons, partaking in the scattering process, have parallel or anti-parallel spin, which is referred to below as triplet or singlet case, respectively, for the sake of comparison with a previous work[
7]. The general procedure, used to calculate the
’s, is to be sketched now. First a
n-electron state
is culled in order to illustrate the two-body scattering. It can be seen in Fig.2 to differ from
in Fig.1, merely by both electrons, involved in the scattering process and located on sites
. Accordingly it is convenient to make use of transpositions
, defined by
, which enables us to recast
, with
given by Eq.(
3), as
wherein
Q stands for
and the sum over
P is to be carried out on
of pairs
, whereas the subset of
n-electron states
is characterised by
. Then
in Eq.(
11) will be fulfilled by requiring
Figure 2.
Sketch of
in Eqs.(
13,
16,
18) and
in Eq.(
20); the various symbols have the same meaning as in Fig.1.
Figure 2.
Sketch of
in Eqs.(
13,
16,
18) and
in Eq.(
20); the various symbols have the same meaning as in Fig.1.
4. 3-Boundary condition
The hereabove results will be taken advantage of, in order to show that the
’s are related to one another through a boundary condition, ensueing from Eq.(
5). To that end, the permutations
and
n-electron states
are needed
with
(
) for
(
). Besides, it is convenient to take the sequence
, such that
. Then the sought boundary condition reads by substituting
to
in Eq.(
22) and applying it further to any
with
and the integers
. At last taking the limit
yields
for which
stand for the whole electron concentration,
triplet or
singlet energy per site and corresponding one-electron density of states, respectively. Though each eigenvalue
is seen to read as a sum over one-particle energies
, as is the case for independent electrons[
9], there is a one to one correspondence between
and its associated,
-dependent
, whereas the density of states is
unique for all many independent electron states. Moreover, since each eigenvector
is defined by a
unique sequence
, two different eigenvectors do not even belong in the same vector space, because their respective
-dependent bases
are thence linearly independent from each other.
Solving Eq.(
24) for the groundstate energy will be done below as follows
with
. Besides Eqs.(
15,
17,
19,
21) imply
whence it is inferred
, so that Eq.(
25) can be recast as
with
. By discretising
and calculating the integral in Eq.(
26) with help of Simpson’s rule, Eq.(
26) will be eventually solved below for
, as a Cramer system, comprising
m equations in terms of the unknowns
, while
are taken from Eq.(
15),Eq.(
17) (Eq.(
19),Eq.(
21)), respectively, in case of
(
). Finally the groundstate energy
is achieved by reckoning
as follows
Assigning the values of thanks to the constraint gives finally access to as the lower of .
5. 4-Groundstate energy for
being constant in Eq.(
15) entails via Eq.(
26)
, which further implies thanks to Eq.(
27)
which shows up independent from
. The
data have been plotted in Fig.3. The inequality
can be seen to hold for any
value. A comparison with groundstate energies
, reckoned with a previous version of Bethe’s wave-function[
6], reveals that the inequality
is found to hold for
, whereas the opposite one
is observed for
. Actually this discrepancy results from
being calculated for
singlet electrons only, whereas
has been worked out for a mixture of
singlet and
triplet electrons. Therefore it can be explained as follows: the smaller
is, the bigger
is with respect to
for
, so that refraining from mixing
singlet and
triplet electrons causes
. Contrarily,
is seen to merge into
for
, because of
as seen in Eqs.(
15,
17). Consequently,
entails that
, so that the
singlet-triplet mixture favours eventually the opposite conclusion
for
.
Figure 3.
Plots of
, as given in Eq.(
28) (white square),
(solid line, dashed line, dotted line, dashed-dotted line) and
(white triangle, white circle, white diamond, ×), calculated for
with
, as said in the text.
Figure 3.
Plots of
, as given in Eq.(
28) (white square),
(solid line, dashed line, dotted line, dashed-dotted line) and
(white triangle, white circle, white diamond, ×), calculated for
with
, as said in the text.
Furthermore it is in order to compare
with the electronic energy
, obtained with help of the correlated Fermi gas (
CFG) model[
7] in normal metals, the characteristic features of which will be recalled now for self-containedness. Each
independent-electron band of dispersion
, as given in Eq.(
1), accommodating at most 2 electrons of opposite spin direction per site, splits into one
singlet and one
triplet band, each of them accommodating at most 1 electron per site. The corresponding dispersion curves
read for the general Hamiltonian
H as follows
for which
has been assumed and
are one-electron creation and annihilation operators on the Bloch[
9] state
Besides, the calculation requires the expression of
H in momentum space, which will be taken from a previous work[
8] for illustration in case of
. Note that
are found in general to depend not only on
but also on the concentration of singlet or triplet electrons
and there is
. Then both
singlet and
triplet bands are populated in accordance with Fermi-Dirac statistics[
9], which yields in the
case at
with
standing for the Fermi energy[
9] and the partial
singlet and
triplet energy, respectively.
The expressions of
are recalled[
7] to read for the Hubbard model
Unlike
, the energies
have been found[
7] quite close to each other, namely there is
for all
-values. As a matter of fact this agreement is all the more baffling, since
is a true eigenvalue, whereas
comes out of a variational calculation, and furthermore the associated eigen- and variational many-electron states belong in quite different vector spaces. Hence this unexpected feature is likely to stem from both states comprising
singlet and
triplet electrons.
6. 5-Groundstate energy for
is inferred from Eq.(
21) to be constant, which entails owing to Eq.(
26) that
is constant too, and eventually thanks to Eq.(
27)
Figure 4.
Plots of
, as given in Eq.(
31) (white square),
(solid line, dashed line, dotted line, dashed-dotted line) and
(white triangle, white circle, white diamond, ×), calculated for
with
and
, as said in the text.
Figure 4.
Plots of
, as given in Eq.(
31) (white square),
(solid line, dashed line, dotted line, dashed-dotted line) and
(white triangle, white circle, white diamond, ×), calculated for
with
and
, as said in the text.
It results from Eq.(
31) that
and
shows up independent from
. As a matter of fact, there is no meaningful
solution either, i.e. with
, for
. The
data, plotted in Fig.4, show that the inequality
holds for all
-values.
Implementing Eqs.(
29,
30) for
yields
The data have been plotted in Fig.4. The inequality is seen to hold for all values and is likely to ensue again from the state, including triplet and singlet electrons with , by contrast with Bethe’s wave-function of eigenvalue comprising only triplet electrons.
7. 6-Discussion
Since the groundstate is widely believed to describe the properties of any physical system at
, it is of significance to sort, out of the various many electron states, discussed hereabove, that one, likely to account at best for the observed properties of interacting electrons. To that end, it should be noticed that all of them share a common property, namely the total energy
consists in a sum over one-
fermion energy either
or
. Thus, the groundstate can be built[
9] by populating every one-electron state from the bottom of the one-electron band, corresponding to
, up to a
dependent upper bound, designated as the Fermi energy
. The reader can check that this requirement is met by all kinds of many-electron states of concern hereabove, i.e. Bethe’s wave-function and
states. However Fermi-Dirac statistics requires in addition at
that the relationship
hold[
9,
10]. Obviously solely the
CFG solution meets successfully this constraint, because it obeys Fermi-Dirac statistics by definition (see
in Eq.(
30)), whereas there is
for all kinds of Bethe’s wave-functions, studied here and elsewhere[
3,
4,
6]. Yet, as seen above, the
CFG solution is
not the groundstate for
in case of
.
Buttressing the claim that the
CFG state is observable would help validate this analysis. To that end, let us discuss electron spin resonance (
ESR) in case of
. Applying an external magnetic field
H lifts[
11] the degeneracy between the respective energies of one-electron states
of
for
triplet electrons as follows
for which
stands for the electron magnetic moment and there is
, taken from Eq.(
29) with
and
instead of
in case of
, which thence implies
. Then the experiment consists of measuring the absorption of a resonant electromagnetic field of frequency
, such that
, which implies
Remarkably the
singlet electrons are seen not to contribute to the
ESR signal, because their associated many-electron state is
not degenerate at
, due to the electrons of spin
keeping
same concentration
even for
. Actually the
ESR signal has been observed[
11] but in a paucity of cases, namely alcali, noble metals and
Al at
GHz/KG. This might stem from
in Eq.(
32), which would shift
upward at fixed
H. Hopefully this work might kindle attempts at seeking the
ESR signal in the
IF rather than microwave range, as done usually[
11] with
.
8. Conclusion
Bethe’s wave-function has been shown to subtend the subspace of scattering eigenstates of the
Hubbard Hamiltonian. Though this analysis is independent of a previous one[
3,
4], both pieces of work lead to the same conclusions, summarised by Eqs.(
3,
7). The two-body scattering plays a key role. Likewise, though the Coulomb force does not depend on the spin of both electrons, involved in the scattering process, its outcome
does indeed (see Eqs.(
15,
17,
19,
21)), as a consequence of Pauli’s principle. This analysis can be applied to any realistic
Hamiltonian, which has been exemplified on
. Nevertheless Bethe’s wave-function turns out to be of limited significance in condensed matter physics, because it has been shown not to be
observable. Actually, the groundstate can be observed at
solely for atoms, molecules and insulators. Conversely it cannot in superconducting and magnetic compounds and in normal metals, because, as argued elsewhere[
7], there are two kinds of conduction electrons at thermal equilibrium with one another in each case, namely normal versus superconducting, normal versus magnetic and
triplet versus
singlet electrons, respectively.