Slave Spins Mean Field Theory

Construction

Start from the observation that the possible occupancies of a real fermion(d) on a given site, n_d=0 and n_d=1, can be represented as two possible states of a spin-1/2 variable, S^z=-1/2 and S^z=+1/2. The idea then is to duplicate the original particle into a spin-like degree of freedom and a fermion charge degree of freedom by introducing a spin-field and an associated auxiliary fermion(f). In this manner vacuum state and occupied states are represented as:

\ket{\emptyset} & = \ket{n_f=0,S^z = -1/2} \\
\ket{1} & \equiv d^\dagger\ket{\emptyset} = \ket{n_f=1,S^z = +1/2}

In this context, the anti-commutation properties of the original fermion operator d are insured by the introduction of the auxiliary fermion operator f. Then, one formulates the constrain:

(1)f^\dagger f = S^z + \frac{1}{2}

which insures that the number of auxiliary fermions and the spin’s states are the same and, above all, it eliminates the non-physical states

\ket{n_f=1,S^z = -1/2} \\
\ket{n_f=0,S^z = +1/2}

It is important to note that the spin-1/2 variable has nothing to do with the actual spin of the particle. The slave spin-1/2 variable is introduced for every fermion species to give the presence of the particle. Therefore, the original particle d_\sigma is replicated into the slave variables with the same quantum numbers: n_{f\sigma} and S^z_\sigma. To exemplify this, the orbital base \{\ket{\emptyset}, \ket{\uparrow}, \ket{\downarrow}, \ket{\uparrow
\downarrow}\} is written in this extended Hilbert space as:

\ket{\emptyset} &= \ket{n_f = 0; S^z_\uparrow = -1/2}\ket{n_f = 0; S^z_\downarrow = -1/2} \\
\ket{\uparrow} &= \ket{n_f = 1; S^z_\uparrow = 1/2}\ket{n_f = 0; S^z_\downarrow = -1/2} \\
\ket{\downarrow} &= \ket{n_f = 0; S^z_\uparrow = -1/2}\ket{n_f = 1; S^z_\downarrow = 1/2} \\
\ket{\uparrow\downarrow} &= \ket{n_f = 1; S^z_\uparrow = 1/2}\ket{n_f = 1; S^z_\downarrow = 1/2}

Then when dealing with a multi orbital system, 2N new spin-1/2 variables S^z_{m\sigma} and 2N auxiliary fermions f_{m\sigma} are introduced, where m=1, \cdots, N is the number of orbitals.

Case: The isolated atom

In order to test the slave-spin method, we first consider the case of a degenerate multi-orbital atom, whose analytical solution is well known. The Hamiltonian in particle-hole symmetry formulation reads:

\mathcal{H} = \frac{U}{2} \left( \sum_{n} d_n^\dagger d_n - N \right)^2
 -\mu \sum_{n} d_n^\dagger d_n

The first step is to recast the Hamiltonian in terms of the new slave-spin operators. The first choice is to use the auxiliary fermions for the non-interacting term(d_n^\dagger d_n\rightarrow f_n^\dagger
f_n) and the slave spins for the interacting case(d_n^\dagger d_n \rightarrow
S^z_n + \frac{1}{2}). We seek a paramagnetic solution, then the constraint (1) which avoids non-physical states is included through a unique Lagrange multiplier \lambda, since all particles are indistinguishable in spin and orbital quantum numbers. After this treatment the Hamiltonian reads:

\mathcal{H} = \frac{U}{2} \left( \sum_{n} S_n^z \right)^2
  + \lambda \sum_{n} \left( S_n^z +\frac{1}{2} - f_n^\dagger f_n \right)
 -\mu \sum_{n} f_n^\dagger f_n

which is possible to separate into a fermion and a spin Hamiltonians:

\mathcal{H}_f &= -(\lambda + \mu) \sum_{n} f_n^\dagger f_n \\
\mathcal{H}_s &= \frac{U}{2} \left( \sum_{n} S_n^z \right)^2
                 +\lambda \sum_{n} (S_n^z + \frac{1}{2})

The spin Hamiltonian has a spectrum E_{Q \leftarrow S+1/2} = U/2(Q-N)^2 +\lambda Q, where Q is the charge present.The Lagrange multiplier \lambda is fixed at the mean-field level by determining the stationary point of the mean-field averaged Hamiltonian: 0=\frac{\partial \langle\mathcal{H}\rangle}{\partial \lambda}. Within this approach, the restriction (1) is therefore respected at the mean field level too.

0 &=-\sum_{n} < f_n^\dagger f_n>_f + \sum_{n} <S_n^z>_s \\
2Nn_F(-\lambda -\mu) &= 2N<S_n^z + \frac{1}{2}> = <S^z> + N = <Q> \\
2Nn_F(-\lambda -\mu) &=
\mathcal{Z}^{-1} \sum_{Q=0}^{2N} \binom{2N}{Q} Q \exp({\beta(U/2(Q-N)^2 +\lambda Q)})

Where every term is averaged with its corresponding Hamiltonian, n_F is the Fermi distribution and the Grand Canonical Partition function from the spin Hamiltonian is \mathcal{Z}= \sum_{Q=0}^{2N} \binom{2N}{Q} \exp({\beta(U/2(Q-N)^2 +\lambda Q)}). Then by numerical root finding the multiplier \bar{\lambda}(\mu,\beta) can be estimated and allows to describe the mean fermion occupation, which is 2Nn_F(-\mu - \bar{\lambda}(\mu,\beta)) . The solution that we find can be conveniently represented by the total fermion occupation as a function of the chemical potential \mu. The resulting curve displays the well known Coulomb ladder-shape, where the system acquires only integer fillings. The change from an integer occupation to an adjacent one takes place abruptly as a function of the chemical potential \mu.

It is necessary to compare the slave spins solution to the exact solution, given by:

2N<d^\dagger d> =  \mathcal{Z}^{-1} \sum_{Q=0}^{2N} \binom{2N}{Q} Q e^{\beta(U/2(Q-N)^2 -\mu Q)}

As shown in the next plot, the slave spin approximation is capable of recovering the coulomb occupation ladder, for the isolated atom with degenerate fermions in spin and orbital. The approximation works best around half-filling.

(Source code)

Case: The lattice model - The Hubbard Model

When in a lattice, atoms have overlapping orbitals and electrons are capable to move along this lattice. Then for the Hamiltonian this term needs to be included appearing in the Tight-Binding formulation. Then as simple extension of the previous isolated atom case and in a multi-orbital scenario, the Hamiltonian reads. The focus now for simplicity is the case of zero crystal-field splitting \epsilon_m=0 and half-filling of each band one electron per site in each orbital \mu=0.

(2)\mathcal{H} = -\sum_m t_m \sum_{<i,j>, \sigma} (d^\dagger_{im\sigma}d_{jm\sigma} +h.c.)
 + \sum_{im\sigma}(\epsilon_m - \mu)d^\dagger_{im\sigma}d_{im\sigma}
 + \frac{U}{2} \sum_i \left( \sum_{m\sigma} d_{im\sigma}^\dagger d_{im\sigma} - N \right)^2

Here it is needed to enforce the restriction:

(3)f_{im\sigma}^\dagger f_{im\sigma} = S_{im\sigma}^z + \frac{1}{2}

using the Lagrange multiplier \lambda_{im\sigma}, which can be used declaring specific constrains to lattice site, orbital, and spin.

When rewriting the Hamiltonian in terms of the auxiliary fermions and the slave spins the interaction term turns easily into:

\mathcal{H}_{int} = \frac{U}{2} \sum_i \left( \sum_{m\sigma} S^z_{im\sigma} \right)^2

For the non interacting part, an appropriate representation of the creation operator has to be chosen. The direct possibility d^\dagger \rightarrow S^+ f^\dagger, although correct leads to problems with the spectral weight conservation because S^+ and S^- don’t commute. Instead the representation d^\dagger \rightarrow
2S^xf^\dagger and d \rightarrow 2S^xf is chosen, which is identical on the physical Hilbert space and involves commuting slave spin operators.

The constrain is treated on average using a static and site, orbital and particle independent Lagrange multiplier \lambda_{im\sigma}. Then the Hamiltonian reads:

\mathcal{H} = &\frac{U}{2} \sum_i \left( \sum_{m\sigma} S^z_{im\sigma} \right)^2 \\
&-\sum_m t_m \sum_{<i,j>, \sigma} 4S^x_{im\sigma}S^x_{jm\sigma}(f^\dagger_{im\sigma}f_{jm\sigma} +h.c.) \\
&+ \sum_{im\sigma}(\epsilon_m - \mu)f^\dagger_{im\sigma}f_{im\sigma} \\
&+\sum_{im\sigma} \lambda_{im\sigma}\left( S_{im\sigma}^z + \frac{1}{2} - f_{im\sigma}^\dagger f_{im\sigma} \right)

Using a Hartree-Fock approximation for the operators S and f:

S^x_{im\sigma}S^x_{jm\sigma}(f^\dagger_{im\sigma}f_{jm\sigma} +h.c.)
\approx <S^x_{im\sigma}S^x_{jm\sigma}>(f^\dagger_{im\sigma}f_{jm\sigma}
+h.c.) \\
+S^x_{im\sigma}S^x_{jm\sigma}<f^\dagger_{im\sigma}f_{jm\sigma} +h.c.> \\
-<S^x_{im\sigma}S^x_{jm\sigma}(f^\dagger_{im\sigma}f_{jm\sigma} +h.c.)>

it is then possible to decouple the Hamiltonian into two effective ones:

(4)\mathcal{H}^f_{eff} = &-\sum_m t_m^{eff} \sum_{<i,j>, \sigma} (f^\dagger_{im\sigma}f_{jm\sigma} +h.c.) \\
&+\sum_{im\sigma} (\epsilon_m - \mu - \lambda_{im\sigma}) f_{im\sigma}^\dagger f_{im\sigma}

(5)\mathcal{H}^S_{eff} = &-\sum_m 4J^{eff}_m \sum_{<i,j>, \sigma} S^x_{im\sigma}S^x_{jm\sigma} \\
&+\sum_{im\sigma} \lambda_{im\sigma} \left( S_{im\sigma}^z + \frac{1}{2} \right)
+\frac{U}{2} \sum_i \left( \sum_{m\sigma} S^z_{im\sigma} \right)^2

Where the effective hopping and the effective exchange constants are determined self consistently from:

(6)t^{eff}_m &= 4t_m<S^x_{im\sigma}S^x_{jm\sigma}>

(7)J^{eff}_m &= t_m<f^\dagger_{im\sigma}f_{jm\sigma} +h.c.>

The fermion field Hamiltonian is a non-interacting one, and it’s analytical solution is well known. For the slave spin Hamiltonian, it can be treated in a single-site using the Weiss mean field approximation.

(8)\mathcal{H}_s = &\sum_{m\sigma} 2h_mS^x_{m\sigma}
+\sum_{m\sigma} \lambda_{m\sigma} \left( S_{m\sigma}^z + \frac{1}{2} \right)
+\frac{U}{2} \left( \sum_{m\sigma} S^z_{m\sigma} \right)^2

Here the mean field h_m has to be determined self-consistently from:

(9)h_m \equiv -2zJ^{eff}_m<S^x_{m\sigma}> = 4<S^x_{m\sigma}>\frac{1}{N_s}\sum_k \epsilon_{km}<f^\dagger_{km\sigma}f_{km\sigma}>

z is the coordination number, \epsilon_{km}=-t_m\sum_{\{\vec{a}\}}e^{-i\vec{k}\cdot\vec{a}} with \{\vec{a}\} the set of vectors to the nearest neighbours

The effective fermion Hamiltonian is

(10)\mathcal{H}^f_{eff} = &\sum_{km\sigma} (-t_m^{eff} \sum_{\{\vec{a}\}} e^{-i\vec{k}\cdot\vec{a}} - \lambda_m) f^\dagger_{km\sigma}f_{km\sigma} \\
&=\sum_{km\sigma} (Z_m\epsilon_{mk} + \epsilon_m - \mu - \lambda_{m\sigma}) f^\dagger_{km\sigma}f_{km\sigma}

where Z_m=4<S^x_{im\sigma}>^2 is the quasiparticle weight.

In the ordered spin basis \{\ket{\uparrow\uparrow}, \ket{\uparrow\downarrow}, \ket{\downarrow\uparrow}, \ket{\uparrow\downarrow}\}, where the spin labelling the operators are then

S^z_{\uparrow} = \frac{1}{2} \left[\begin{smallmatrix}1 & 0 & 0 & 0\\0 & 1 & 0 & 0\\0 & 0 & -1 & 0\\0 & 0 & 0 & -1\end{smallmatrix}\right]

S^z_{\downarrow} = \frac{1}{2} \left[\begin{smallmatrix}1 & 0 & 0 & 0\\0 & -1 & 0 & 0\\0 & 0 & 1 & 0\\0 & 0 & 0 & -1\end{smallmatrix}\right]

S^x_{\uparrow} = \frac{1}{2} \left[\begin{smallmatrix}0 & 0 & 1 & 0\\0 & 0 & 0 & 1\\1 & 0 & 0 & 0\\0 & 1 & 0 & 0\end{smallmatrix}\right]

S^x_{\downarrow} = \frac{1}{2} \left[\begin{smallmatrix}0 & 1 & 0 & 0\\1 & 0 & 0 & 0\\0 & 0 & 0 & 1\\0 & 0 & 1 & 0\end{smallmatrix}\right]

As seen in equations (2) there is no hybridization between bands, hopping preserves then the orbital quantum number. When treating the system within a local mean field, in absence of hybridization the \vec{k} dependence enters the problem only through each band dispersion as seen in equations (9), (10). Sums over momenta can thus be replaced by integrals over the energy weighted by the density of states D(\epsilon), which is specific to the lattice geometry and dimension. For this work, as commonly employed in the literature, the Bethe lattice will be used. It has a very simple semi-circular form of the density of states:

(11)D(\epsilon) = \frac{1}{2 \pi t^{2}} \sqrt{4 t^{2} - \epsilon^{2}}

and allows to simplify calculations in a great amount. Here t is the hopping amplitude and the half-bandwidth is D=2t, which is set as the energy unit throughout this work. It is known moreover that the Bethe lattice well portrays the salient physical properties of the Mott-Hubbard transition and it has immediate connection with the dynamical mean field theory [Georges1996], which is exact in the infinite dimensions limit and which we intend to implement in future work.

The mean field in equation (9) is then simplified into:

(12)h_{m\sigma} = \langle O_{m\sigma} \rangle \int_{-\infty}^\infty
\epsilon D(\epsilon)  n_F(Z_{m\sigma}\epsilon + \epsilon_m - \mu -
\lambda_{m\sigma}) d\epsilon

where n_F is the Fermi distribution function. In the same fashion to estimate the average particle number per site, orbital and spin, one easily uses the relation:

(13)\langle n_{im\sigma}\rangle  = \int_{-\infty}^\infty
D(\epsilon)  n_F(Z_{m\sigma}\epsilon + \epsilon_m - \mu -
\lambda_{m\sigma}) d\epsilon

In this work all calculations are done at zero temperature, where the Fermi distribution can be approximated into a step function. That implies for equations (13) and (12) that:

(14)Z_{m\sigma}\epsilon_{F_0}(n) = - \epsilon_m + \mu + \lambda_{m\sigma}

in which \epsilon_{F_0} is the Fermi energy at zero temperature for the non-interacting system such that

(15)\int_{-\infty}^{\epsilon_{F_0}} D(\epsilon)  d\epsilon=n

This procedure of defining a zero temperature non-interacting Fermi energy(U=0 and thus Z=1) allows to keep the particle population fixed when correlations are included into the problem [Yu2011] [Florens2004].

(Source code)

Introducing doping

The previous introduction to the slave spins treats the Hubbard Hamiltonian and deals with the quartic term that deals with the correlations. But it is only valid in the simplified case of degenerate orbitals at half-filling. When aiming to introduce doping, a more elaborate formulation is required. The fact is that the Spin Hamiltonian is unable to manage doping cases as previously formulated, since it is always populated by 2 spin which flip by the action of the operator S^x in a balanced manner. This is equivalent to the real electron hopping from one lattice site to the next.

In the case of electron or hole doping, the fermion Hamiltonian deals with it through the chemical potential, which doesn’t appear in the Spin Hamiltonian. Then a new generic Spin operator needs to be introduced, to treat this spin flip unbalance in the Spin system and deal with the doping cases. It is an mixture of the spin raising and lowering operators.

(16)O^\dagger = S^+ + cS^-

The inclusion of the gauge parameter c originates from the study of the action of the real creation and annihilation operators (d^\dagger, d) into the real states. And then how the new operators act into the extended Hilbert space. The reference case is such

\begin{align*}
  d\ket{\emptyset} &= 0 &   d\ket{1} &= \ket{0} \\
  d^\dagger\ket{1} &= 0 &   d^\dagger\ket{\emptyset} &= \ket{1}
\end{align*}

The first conditions on the left can be assured by the fermion operator.

f O \ket{n^f=0, S^z = -\frac{1}{2}} = 0 \\
f^\dagger O^\dagger \ket{n^f = 1, S^z = +\frac{1}{2}} = 0

the O operator does not play any role. For second set of conditions

f O \ket{n^f=1, S^z = +\frac{1}{2}} = \ket{n^f = 0, S^z = -\frac{1}{2}} \\
f^\dagger O^\dagger \ket{n^f = 0, S^z = -\frac{1}{2}} = \ket{n^f = 1, S^z =
+\frac{1}{2}}

only three out of four matrix elements of the O operator can be determined, which implies that

(17)O = \left( \begin{matrix} 0 & c \\ 1 & 0 \end{matrix} \right)

Which is in correspondence with the previous formulation of equation (16). c can be an arbitrary complex number and it is tuned in order to give rise to the most physical approximation scheme, by imposing that it correctly reproduces solvable limits of the problem such as the non-interacting limit.

The non-interacting limit

Starting with a Tight Binding Hamiltonian that only includes the kinetic energy term and a contribution from the chemical potential to control the doping, the new operators are introduced. Treating a single orbital case of atoms in a lattice.

\mathcal{H}_0 &= -t\sum_{<ij>, \sigma} (d^\dagger_{i\sigma} d_{j\sigma}
+h.c.)
- \mu\sum_{i\sigma} d^\dagger_{i\sigma}d_{i\sigma} \\
\rightarrow \mathcal{H}_0 &= -t\sum_{<ij>, \sigma}
(O^\dagger_{i\sigma}O_{i\sigma} f^\dagger_{i\sigma}f_{i\sigma} +h.c.) -
\mu\sum_{i\sigma} f^\dagger_{i\sigma} f_{i\sigma} \\
&+ \sum_{i\sigma} \lambda_i(S^z_{i\sigma} + \frac{1}{2} -
  f^\dagger_{i\sigma}f_{i\sigma})

Here the Lagrange multiplier is treated as individual to every site. Using first a Hartree-Fock approximation in the tight binding term and respecting the restriction set by the Lagrange multiplier, it is possible to separate the Hamiltonian into 2 coupled effective ones.

\mathcal{H}_f &= -t\sum_{<ij>,\sigma}( Q_{ij}f^\dagger_{i\sigma}f_{i\sigma}
+h.c.) - \sum_{i\sigma}(\mu + \lambda_i) f^\dagger_{i\sigma}f_{i\sigma} \\
\mathcal{H}_s &= -\sum_{<ij>,\sigma} ( J_{ij}O^\dagger_{i\sigma}O_{j\sigma}
+h.c.) + \sum_{i\sigma} \lambda_i(S^z_{i\sigma} + \frac{1}{2})

The parameters Q_{ij} hopping renormalization factor, J_{ij} slave-spin exchange constant and \lambda_i in these expressions are determined from the coupled self-consistency equations:

(18)Q_{ij} = \langle O^\dagger_{i\sigma}O_{i\sigma} \rangle

(19)J_{ij} = t \langle f^\dagger_{i\sigma}f_{j\sigma} \rangle

(20)\langle n^f_{i\sigma} \rangle_f = \langle S^z_{i\sigma} \rangle_s +
\frac{1}{2}

One further approximation needs to be applied, and it is to treat the spin Hamiltonian within a Weiss mean field, in which a single site is embedded in an effective field of its surroundings. The spin Hamiltonian becomes:

\mathcal{H}_S = \sum_\sigma (h_\sigma O^\dagger_\sigma + h.c.)
+\sum_{\sigma} \lambda(S^z_{\sigma} + \frac{1}{2})

Where the mean field h_\sigma

h_\sigma = \langle O_\sigma \rangle \frac{1}{N_s} \sum_k \epsilon_k
\langle f^\dagger_{k\sigma}f_{k\sigma} \rangle

Choosing the gauge c

The condition set, to reproduce the non-interacting case is Q_{ij} = Z = 1, where the quasiparticle residum Z=\langle O_\sigma \rangle. In the single site approximation Q_{ij} = Z by construction, and only it’s unitary value remains to be enforced Z=1. The non-interacting spin Hamiltonian is treated suppressing the spin index \sigma since in this case up-spin and down-spin fermions are decoupled.

\mathcal{H}_S &= hO^\dagger+\overline{h}O+\lambda(S^z_\sigma + \frac{1}{2})\\
&= \begin{pmatrix} \lambda & c \overline{h} + h\\h \overline{c} +
\overline{h} & 0 \end{pmatrix}

It is possible to diagonalize the Hamiltonian for one slave spin in the S^z =
\pm 1/2 basis. The ground state eigenvalue E_{GS} and the corresponding eigenstate are.

E_{GS} &= \frac{\lambda}{2} - \sqrt{\frac{\lambda^2}{4} + |a| ^2} \equiv
\frac{\lambda}{2} - R \\
\ket{GS} &= \begin{pmatrix} - \frac{a}{N} \\ \frac{\lambda /2+R}{N}
\end{pmatrix}

Where N=\sqrt{2R(\lambda /2 +R)} and a=c \overline{h} + h. The expected values of S^z and O are:

\langle S^z \rangle = -\frac{\lambda}{4R} \\
\langle O   \rangle = - \frac{a +\overline{a}c}{2R}

It is clearly seen that the Lagrange multiplier \lambda depends on the density n and it is adjusted to satisfy the constraint equation:

(21)n-\frac{1}{2} = \langle S^z \rangle = -\frac{\lambda}{4R}

and c needs to be tuned to match the condition Z=1

(22)Z = \langle O \rangle^2 = \frac{|a +\overline{a}c| ^2 }{4R^2}=1

It is possible to eliminate \lambda from the conditions by squaring (21) and using the relation (22), following the next derivation:

\frac{|a| ^2}{4R^2} +(n-\frac{1}{2})^2 &= -\frac{\lambda^2}{16R^2} +
\frac{|a| ^2}{4R^2} \\
\frac{|a| ^2}{|a+\overline{a}c| ^2} &= n - n^2

Then it is possible to choose c real, making also h and a real. The expression for c is found to be independent of the mean field h:

c = \frac{1}{\sqrt{n(1-n)}} -1

(Source code)

(Source code)

Hund’s coupling

The interaction Hamiltonian in this multiorbital systems then becomes:

\mathcal{H} =& U \sum_{i,m} n_{im\uparrow}n_{im\downarrow}
 + \left(U' - \frac{J}{2} \right) \sum_{i\sigma \atop m>m'} n_{im\sigma}n_{im'\sigma} \\
 &-J\sum_{i,m>m'}\left( 2\vec{S}_{im}\cdot\vec{S}_{im'}
 + ( d^\dagger_{im\uparrow}d^\dagger_{im\downarrow}d_{im'\uparrow}d_{im'\downarrow} + h.c.)  \right)

Where U is the intra orbital Hubbard interaction and U' = U - 2J is the interorbital interaction \vec{S}_{im} is the total spin in an orbital. Which is then traded into slave spin variables to give:

\mathcal{H} =& \frac{U'}{2} \sum_i \left(\sum_{m\sigma} S^z_{im\sigma} \right)^2
  + J \sum_{im}\left(\sum_\sigma S^z_{im\sigma} \right)^2
  - \frac{J}{2} \sum_{i\sigma} \left(\sum_m S^z_{im\sigma} \right)^2 \\
 & - J \sum_{i,m>m'}\left( S^+_{im\uparrow}S^-_{im\downarrow}S^+_{im'\downarrow}S^-_{im'\uparrow}
   + S^+_{im\uparrow}S^+_{im\downarrow}S^-_{im'\uparrow}S^-_{im'\downarrow}+ h.c. \right)