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

Finite and Infinite Matrix Product States
for Gutzwiller Projected Mean-Field Wavefunctions

Gabriel Petrica [email protected] Institute for Condensed Matter Theory and IQUIST and Department of Physics, University of Illinois at Urbana-Champaign, Urbana, IL 61801, USA    Bo-Xiao Zheng Division of Chemistry and Chemical Engineering, California Institute of Technology, Pasadena, California 91125, USA AxiomQuant Investment Management LLC, Shanghai 200120, China    Garnet Kin-Lic Chan Division of Chemistry and Chemical Engineering, California Institute of Technology, Pasadena, California 91125, USA    Bryan K. Clark [email protected] Institute for Condensed Matter Theory and IQUIST and Department of Physics, University of Illinois at Urbana-Champaign, Urbana, IL 61801, USA
Abstract

Matrix product states (MPS) and ‘dressed’ ground states of quadratic mean fields (e.g. Gutzwiller projected Slater Determinants) are both important classes of variational wave-functions. This latter class has played important roles in understanding superconductivity and quantum spin-liquids. We present a novel method to obtain both the finite and infinite MPS (iMPS) representation of the ground state of an arbitrary fermionic quadratic mean-field Hamiltonian, (which in the simplest case is a Slater determinant and in the most general case is a Pfaffian). We also show how to represent products of such states (e.g. determinants times Pfaffians). From this representation one can project to single occupancy and evaluate the entanglement spectra after Gutzwiller projection. We then obtain the MPS and iMPS representation of Gutzwiller projected mean-field states that arise from the variational slave-fermion approach to the S=1S=1 Bilinear-Biquadratic (BLBQ) quantum spin chain. To accomplish this, we develop an approach to orthogonalize degenerate iMPS to find all the states in the degenerate ground-state manifold. We find the energies of the MPS and iMPS states match the variational energies closely indicating the method is accurate and there is minimal loss due to truncation error. We then present the first exploration of the entanglement spectra of projected slave-fermion states exploring their qualitative features and finding good qualitative agreement with the respective exact ground state spectra found from DMRG.

I Introduction

Variational wave-functions are frequently used to understand quantum many-body systems. Two important classes of variational wave-functions are dressed Slater determinants and tensor networks. Dressed Slater determinants introduce correlation on top of a mean-field ground state. On the other hand, a tensor network is represented as a network of connected tensors providing a natural framework in which to understand and represent low-entangled quantum states (see fig. 1).

Slater determinants (SDs) (and other generalized quadratic ground states such as Bogoliubov-de Gennes (BdG) leggett2006quantum and Pfaffian states read2000paired ) have played a key role in the understanding of physical systems ranging from their use as the Hartree-Fock solution in quantum chemistry to being applied as a starting mean-field ansatz for strongly-correlated systems. These latter ansatz are then dressed in various ways: Slater-Jastrow wave-functions are the de-facto standard for simulating material systems in quantum Monte Carlo; many prototypical quantum Hall states are represented as powers or products of Slater Determinants and Pfaffians; and projected mean-field states are an important starting point for probing the physics of high temperature superconductivity as well as quantum spin-liquids.

While dressed mean-field states are often easy to represent in variational Monte Carlo (VMC), it is also often difficult to extract certain properties from VMC. Foremost among these is the entanglement spectra which is an important metric used for understanding topological phases of matter. Even properties which can be extracted easily, such as the energy, can be statistically noisy making aspects such as optimization difficult. Moreover, evaluating dressed mean-field states in Monte Carlo scales cubically with the system size making the approach to the thermodynamic limit costly. Matrix product states avoid many of these problems in one dimensional systems and ladders: they are ideally suited for extracting entanglement spectra, computing observables exactly without any statistical noise, and directly representing (gapped) physical systems in the thermodynamic limit.

Our main contribution in this paper is to describe a series of efficient and highly parallel algorithms which take (projected) mean field (i.e. quadratic) eigenstates and generate both finite (fMPS) and infinite (iMPS) matrix product states from them. We will also show how to generate fMPS and iMPS for products of mean-field wave-functions. We will then apply our approach to compute the MPS and entanglement spectra of a series of projected slave-fermion wave-functions of the bilinear biquadratic model. This example will bring to light a number of interesting aspects of generating multiple degenerate ground states from Gutzwiller projected slave-fermion systems in iMPS.

Beyond this particular application, being able to generate a MPS from a projected SD is generically useful. It allows for more faithful comparisons between slave-fermion and DMRG results which often disagree on the underlying phase of spin-liquids. It could be used to initialize DMRG with a good initial mean-field guess for certain Hamiltonians. This can be useful both for calculations on discrete lattices as well as DMRG in the continuum. Because there exist algorithms which build MPS on quantum computers, it immediately gives an additional approach to generate a dressed quadratic mean-field state on a quantum device.

We are aware of two other algorithms which convert Slater determinants to MPS fishman2015compression , wu2020tensor . Both these are based on the idea of applying quantum gates or matrix-product operators to a simpler quantum state to generate the MPS. Our approach differs from these techniques in two key ways: (1) we can generate the infinite MPS for a family of slater Determinants and (2) we generate our (i)MPS by directly generating the MPS coefficients without the application of any operators to the system. We also note that ref. schuch2019matrix, represents Slater Determinants in a MPS-like framework in a Gaussian fermionic representation.

In section II, we will describe our key algorithm for turning a SD into a (i)MPS. In section III, we will show a series of examples for how to use this basic procedure for generating more complicated mean-field states (i.e. pfaffians) as well as states which are products of mean-field states. Finally, in section IV we focus on computing (i)MPS for the slave-fermion states of the billinear-biquadratic model showing their entanglement spectra and energy.

II Slater Determinants to MPS

In this section we are going to show how to generate either a finite matrix-product state (fMPStoSD) or an infinite matrix product state (iMPStoSD) from a Slater determinant (SD). This will not only be useful in its own right but will be the key operation used in the rest of this work to produce MPS for both more complicated quadratic mean field states as well as dressed versions of these states.

fMPStoSD generates the matrix product state site by site in an approach that is highly reminiscent of the site-decimation canonical technique to convert a generic wave-function (i.e. a multi-site tensor) into a matrix product state schollwock2011density . The typical site-decimation procedure involves performing SVD’s over matrices generated by collecting different subsets of indices into the two matrix dimensions. This general approach will become efficient to use with Slater determinants because SVD’s of Slater determinants are efficient and generate sums of products of Slater determinants.

In fMPStoSD we perform a series of Schmidt decompositions over all bipartitions of our system. Each Schmidt decomposition generates a set of Schmidt vectors; each such Schmidt vector is a Slater determinant. The MPS is then generated by taking overlaps of these Slater determinant Schmidt vectors with each other in the correct way.

In iMPStoSD we can easily generate the bulk uniform iMPS tensor from just two Schmidt decompositions: one for each of two ground state Slater determinants of the same Hamiltonian defined on sufficiently large systems that differ in size by one unit cell. Again, these Schmidt decompositions will have Slater determinant Schmidt eigenvectors. After we appropriately fix the gauge of the two Schmidt decompositions, the uniform bulk MPS tensor will be generated from appropriate overlaps of these Schmidt eigenvectors.

II.1 Slater Determinant \rightarrow Finite MPS

In this section, we show in detail how to convert a Slater determinant into a finite matrix product state. The Schmidt decomposition of a Slater determinant |SD|SD\rangle on NN sites bi-partitioned into two regions cut between sites ii and i+1i+1 will be notated as

|SD=αλαi;N|Lαi;N|Rαi;N\ket{SD}=\sum_{\alpha}\lambda^{i;N}_{\alpha}\ket{L_{\alpha}^{i;N}}\ket{R_{\alpha}^{i;N}} (1)

where |Lα|L_{\alpha}\rangle and |Rα|R_{\alpha}\rangle are the α\alpha’th left and right Schmidt vectors respectively (with support in their respective subsystem) and λα\lambda_{\alpha} is the α\alpha’th Schmidt eigenvalue. Note that, for a Slater determinant, each of the individual left and right Schmidt vectors are also Slater determinants and efficiently computable cheong2004many ; klich2006lower ; knizia2012density ; peschel2012special . Slater determinants are specified by a set of single-particle orbitals and all the Slater determinants in the set of right Schmidt vectors {|Rα}\{|R_{\alpha}\rangle\} are specified by subsets of single particle orbitals from a set of (at most) NN single-particle orbitals {ϕ1RϕNR}\{\phi^{R}_{1}...\phi^{R}_{N}\} defined on the (inclusive) sites [(i+1),,N][(i+1),\ldots,N]. There are, at most, 2N2^{N} such subsets. Analogous statements hold for the left Schmidt vectors.

Refer to caption
(a)
Refer to caption
(b)
Figure 1: a) Graphical representation of the left-boundary, bulk and right boundary AA tensors forming an open-boundary matrix product state. The α\alpha’s are the virtual indices; the σ\sigma’s are the physical indices. b) Graphical representation of an 88-site matrix product state

A general matrix product state can be written as

|MPS={σ},{a}A1,a1[1]σ1AaN1,1[N]σN|σ1σN\ket{MPS}=\sum_{\{\sigma\},\{a\}}A^{[1]\sigma_{1}}_{1,a_{1}}\cdots A^{[N]\sigma_{N}}_{a_{N-1},1}\ket{\sigma_{1}\cdots\sigma_{N}} (2)

where A[k]A^{[k]} is the kk’th three-tensor specified by the physical index σk\sigma_{k} (e.g. occupancy or spin) and the virtual indices (αk1,αk)\left(\alpha_{k-1},\alpha_{k}\right) schollwock2011density . To generate the MPS of a Slater determinant, we compute each three-tensor A[i+1]A^{[i+1]} as

Aαkαk+1[i+1]σi+1=(σi+1|Rαk+1i+1;N|)|Rαki;NA^{[i+1]\sigma_{i+1}}_{\alpha_{k}\alpha_{k+1}}=\left(\bra{\sigma_{i+1}}\otimes\bra{R^{i+1;N}_{\alpha_{k+1}}}\right)\ket{R^{i;N}_{\alpha_{k}}} (3)

giving a matrix which is in right canonical form, i.e. σA[i+1]σ(A[i+1]σ)=I\sum_{\sigma}A^{[i+1]\sigma}\left(A^{[i+1]\sigma}\right)^{\dagger}=I. Note that this procedure is very similar to the one which transforms a vector into a MPS schollwock2011density and works for the same reason: the sets {|Rαi;N}\{|R^{i;N}_{\alpha}\rangle\} and {|σi|Rβi+1;N}\{|\sigma_{i}\rangle\otimes|R^{i+1;N}_{\beta}\rangle\} span the same space and therefore there is a transformation AA which rotates between them. In practice, we keep the bond-dimension of AA controlled by only computing the Schmidt vectors whose Schmidt values are above a certain threshold ϵ\epsilon. This can be done without computing any Schmidt eigenvector with eigenvalue less than ϵ\epsilon. Here we have focused on the bulk tensors and slight modifications need to be made for the boundary tensors A[1]σ1A^{[1]\sigma_{1}} and A[N]σNA^{[N]\sigma_{N}} (see supplement  S1).

Refer to caption
Figure 2: Top: Graphical representation of the Schmidt decomposition of the wavefunction |Ψ8=βλβ4;8|Lβ4;8|Rβ4;8\ket{\Psi^{8}}=\sum_{\beta}\lambda_{\beta}^{4;8}\ket{L_{\beta}^{4;8}}\ket{R_{\beta}^{4;8}} over a bipartition [1,,4]×[5,,8][1,\ldots,4]\times[5,\ldots,8] of an 88-site system. Middle, Bottom: Two additional ways of representing the quantum state |Ψ8|\Psi_{8}\rangle. The tensor A[4]A^{[4]} is constructed by having the overlap of the right five sites of the bottom two figures equal one.

We now describe how to efficiently evaluate the matrix elements of each AA. We start by noting that |σi+1||Ri+1;N|\sigma_{i+1}|\rangle\otimes|R^{i+1;N}\rangle is also a Slater determinant. It is specified by the single particle orbitals

{[0ϕa],[0ϕb],,[0ϕc],ϕi+1}\{[0\phi_{a}],[0\phi_{b}],\cdots,[0\phi_{c}],\phi^{i+1}\} (4)

where [0ϕa][0\phi_{a}] is the single particle orbital with coefficients in the lattice basis [0,ϕa(i+2),ϕa(i+3),,ϕa(N)][0,\phi_{a}(i+2),\phi_{a}(i+3),\cdots,\phi_{a}(N)] and ϕi+1\phi^{i+1} is the single-particle orbital in analogous notation, [1,0,,0][1,0,\cdots,0]. Eqn. 3 then reduces to the overlap of two Slater determinants of size (Ni)×(Ni)(N-i)\times(N-i) which can be computed in O(N3)O(N^{3}) time.

While naively each element of AA requires such a computation, there is a significant overlap in these different computations which reduce the naive compuational complexity of the tensor computation. There are two steps in computing the overlap of two Slater determinants: evaluating the overlap matrix between all pairs of single particle orbitals that make up the two determinants and computing the determinant of this overlap matrix. All the Slater determinants used in the ket (respectively bra) of eqn. 3 (over different terms in A) come from a subset of single-particle orbitals of the N-orbital set {ϕ1R,..,ϕNR}\{\phi_{1}^{R},..,\phi_{N}^{R}\}. We can compute the overlap matrix of all these respective single-particle orbitals once per three-tensor AA at a cost of O(N3).O(N^{3}). The entries of AA are then determinants of submatrices of this overlap matrix. While naively each determinant also costs O(N3)O(N^{3}) to compute, the submatrices differ only in the bottom log2D\log_{2}D columns and right log2D\log_{2}D rows where DD is the bond-dimension of AA; determinant update formulas can then be used to accelerate this computation letting each determinant be computed in time O(N2log2D)O(N^{2}\log_{2}D) after an initial O(N3)O(N^{3}) operation to evaluate the inverse of the upper-left (Nlog2D)×(Nlog2D)(N-\log_{2}D)\times(N-\log_{2}D) block of the overlap matrix. The whole evaluation of each tensor AA can be done in O(N3)+O(D2N2log2D)O(N^{3})+O(D^{2}N^{2}\log_{2}D) time. This can be further attenuated somewhat by more aggressive use of determinant update formulas harville1998matrix .

Notice that there are significant parts of this algorithm that can be run in parallel. Each three-tensor AA can be computed separately. Within each A,A, the Schmidt decomposition can be partially parallelized; each element of the overlap matrix can be computed in parallel; and, after the initial evaluation of the inverse of the upper-left block of the overlap matrix, each determinant can then be computed in parallel.

II.2 Gapped Slater Determinant \rightarrow Infinite MPS

The above procedure generates a finite MPS approximation (the accuracy of the representation is given as an input to the algorithm) of any Slater determinant. In this section, we describe how to generate an infinite MPS from the Slater determinant ground state of a gapped mean-field Hamiltonian. This infinite MPS can be described by left LL and right RR boundary tensors which sandwich the bulk tensor AA giving us an infinite matrix product state of the form,

|iMPS=σLσLAσn1AσnAσn+1RσR×|σLσn1σnσn+1σR\ket{\textrm{iMPS}}=\sum_{\sigma}L^{\sigma_{L}}\ldots A^{\sigma_{n-1}}A^{\sigma_{n}}A^{\sigma_{n+1}}\ldots R^{\sigma_{R}}\times\\ \ket{\sigma_{L}\ldots\sigma_{n-1}\sigma_{n}\sigma_{n+1}\ldots\sigma_{R}} (5)

with an arbitrary number of bulk tensors AA. LL and RR are tensors which span a fixed number kk of sites. Note that any thermodynamic observable can be computed directly in the thermodynamic limit of the Slater Determinant using only the bulk tensor AA. In addition, we can compute the amplitude for the Slater determinant on any (large enough) system size, by inserting the corresponding number of bulk tensors between the boundary tensors LL and RR (i.e. to generate the MPS for a NN site Slater determinant from the infinite MPS, we therefore use N2kN-2k bulk tensors AA); see fig.3.

Refer to caption
Figure 3: We obtain the finite MPS |ψ2N+n\ket{\psi^{2N+n}} by inserting nn (in the figure n=3n=3) AA iMPS bulk tensors between the left |LN/2;N\ket{L^{N/2;N}} and right |RN/2;N\ket{R^{N/2;N}} Schmidt vectors obtained from |ψ2N\ket{\psi^{2N}}.

To generate the iMPS, we start off by producing two Slater determinants defined on 2N2N and 2N+12N+1 sites, where NN is sufficiently large such that the entanglement spectrum is constant over cuts in the “bulk” of the wave-functions. For gapped systems, we generically expect the entanglement spectrum over the bulk to be constant; see fig.  S3 in supplement  S5 for an example of this. We then generate the Schmidt decompositions

|Ψ2N\displaystyle|\Psi^{2N}\rangle =\displaystyle= αλαN;2N|LαN;2N|RαN;2N\displaystyle\sum_{\alpha}\lambda^{N;2N}_{\alpha}|L_{\alpha}^{N;2N}\rangle|R_{\alpha}^{N;2N}\rangle (6)
|Ψ2N+1\displaystyle|\Psi^{2N+1}\rangle =\displaystyle= αλαN;2N+1|LαN;2N+1|RαN;2N+1\displaystyle\sum_{\alpha}\lambda^{N;2N+1}_{\alpha}|L_{\alpha}^{N;2N+1}\rangle|R_{\alpha}^{N;2N+1}\rangle (7)

Both |LαN;2N|L_{\alpha}^{N;2N}\rangle and |LαN;2N+1|L_{\alpha}^{N;2N+1}\rangle are going to be the same up to a gauge freedom. We fix this gauge freedom by defining a unitary

CαβN;2N+1=\displaystyle C_{\alpha\beta}^{N;2N+1}= 0\displaystyle 0 if λαλβ\displaystyle\textrm{ if }\lambda_{\alpha}\neq\lambda_{\beta} (8)
=\displaystyle= LN;2N|α|LN;2N+1β\displaystyle\bra{L^{N;2N}}_{\alpha}\ket{L^{N;2N+1}}_{\beta} otherwise

which rotates between Schmidt eigenvectors with the same Schmidt eigenvalue allowing the state on 2N+12N+1 sites to be defined as

|Ψ2N+1=αγ|LN;2NαλαN;2N+1CαγN;2N+1|RN;2N+1γ\ket{\Psi^{2N+1}}=\sum_{\alpha\gamma}\ket{L^{N;2N}}_{\alpha}\lambda^{N;2N+1}_{\alpha}C^{N;2N+1}_{\alpha\gamma}\ket{R^{N;2N+1}}_{\gamma} (9)

Then the tensor AA for the iMPS is

AαβσN+1=γCαγ[2N+1]σN+1|RN;2N|β||RN,2N+1γA^{\sigma_{N+1}}_{\alpha\beta}=\sum_{\gamma}C^{[2N+1]}_{\alpha\gamma}\bra{\sigma_{N+1}|\bra{R^{N;2N}}_{\beta}}\ket{R^{N,2N+1}}_{\gamma} (10)
Refer to caption
Figure 4: Illustration of the algorithm for generating the infinite MPS representation of a Slater determinant. Lines 1 and 2 correspond to the standard Schmidt decomposition after site NN of wavefunctions defined on 2N2N and 2N+12N+1 sites. For line three, we use our gauge freedom to replace the left-Schmidt eigenvalues λN;2N+1\lambda^{N;2N+1} and eigenvectors |LN;2N+1\ket{L^{N;2N+1}} with eigenvalues λN;2N\lambda^{N;2N} and eigenvectors |LN;2N\ket{L^{N;2N}} rotated by CC where CC is defined in eqn. II.2. Finally, the tensor AA is constructed by having the overlap of the right N+1N+1 sites (including CC) of the bottom two lines equal one.

As in the finite MPS case, we have that the single-particle orbitals of the Slater determinant |σN+1|RN;2N|\sigma_{N+1}\rangle|R^{N;2N}\rangle are shifted to the right with an initial zero as their first element. The overlap of this tensor can be computed in exactly the same way as for the finite MPS case. Here, though, we only need to evaluate one tensor AA instead of a tensor per site, with the assumption that we are using an iMPS defined by a single tensor (i.e. single site unit cell) AA. This process can be generalized to multi-site unit cells as well (see supplement S2). Note that by directly applying the finite MPS algorithm to large systems to try to find the bulk tensor AA will fail because the gauge freedom available in the tensors will prevent a single identical bulk tensor from being produced at each step.

II.3 Numerical Validation

We numerically validate our algorithms by applying fMPStoSD and iMPStoSD on the ground state of the Su-Schrieffer-Heeger(SSH) model su1979solitons ,

HSSH=vn(cn,1cn,2+h.c.)+wn(cn+1,1cn,2+h.c.)H_{SSH}=v\sum_{n}\left(c^{\dagger}_{n,1}c_{n,2}+\text{h.c.}\right)+\\ w\sum_{n}\left(c^{\dagger}_{n+1,1}c_{n,2}+\text{h.c.}\right) (11)

The model describes spinless fermions on a 1D lattice, with a 2 site unit cell made up of AA,BB sites, with different (real) parameters for intra-cell hopping (vv) and intercell hopping (ww). It admits two different quantum ground states, distinct in their topological properties: a trivially gapped phase for v>wv>w and a (symmetry protected) topological gapped ground state, characterized by the presence of two zero-energy edge modes inside the gap, for v<wv<w, separated by a quantum critical point at v=wv=w.

We will discuss here the trivial ground state. A small subtlety related to choosing the same gapless boundary mode in the Slater determinant wave-functions used for generating the uniform tensor in the iMPS procedure is delegated to the supplement S4. For the finite Slater Determinant, we compare the MPS we generate using two different truncation values against the exact Slater Determinant by comparing all of the amplitudes (see 1st column in fig. 5). For the infinite case, we generate the iMPS and then use the bulk tensor we have computed along with the boundary tensors to compute amplitudes for a much larger system and again compare amplitudes against the exact solution for that much larger system (see 2nd column in fig. 5). In both cases, we find that the amplitudes are in very good agreement for all amplitudes down to the Schmidt eigenvalue cutoff.

Refer to caption
Figure 5: Comparison of amplitudes (normalized by the largest amplitude seen) between our fMPS/iMPS wave-functions and the exact SD. The largest (normalized) amplitude is at the “origin” of the graphs with smaller amplitudes toward both edges. Orange triangles are values at which the fMPS/iMPS gives zero amplitude; for these the ”y” coordinate is arbitrarily set to 1. The amplitudes for the top row are less accurate as they are generated with larger MPS thresholds ϵ\epsilon. (a) and (c) compare all amplitudes for N=8N=8 on (v=1.0v=1.0; w=0.6w=0.6; eqn. 11) and (t=1t=1; μ=3\mu=3; Δ=1\Delta=1; eqn. 16) respectively. (b) We compare 459428 random configuration (top and bottom are different configurations) between the N=24N=24 Slater determinant with (v=1.0v=1.0; w=0.6w=0.6) of eqn. 11 and a MPS generated from 8 uniform iMPS bulk tensors (generated from SD on N=16,17N=16,17 sites) sandwiched between the 88-left and 88-right tensors from the 1616-site Slater determinant. (d) We compare 4997249972 (top) and 3493334933 (bottom) random configuration between the N=32N=32 p+ip Pfaffian groundstate with (t=1.0t=1.0; Δ=1\Delta=-1 μ=2.2\mu=-2.2) of eqn. 17 and a MPS generated from 8 uniform iMPS bulk tensors (generated from SD on N=24,25N=24,25 sites) sandwiched between the 1212-left and 1212-right tensors from the 2424-site Pfaffian.

III General (Dressed) Quadratic Mean Fields

In sec. II we showed how to generate a matrix product state from a Slater determinant. In this section, we show that this machinery gives us the means to generate the matrix-product state representations of ground states of arbitrary quadratic mean-field Hamiltonians.

All quadratic Hamiltonians can be easily diagonalized using a canonical transformation ripka1986quantum . Without loss of generality, in our derivations, we will use translation invariant systems for ease of presentation. We will first go through two canonical examples. In III.1 we will show how to produce the MPS representation of groundstates of BdG Hamiltonians which are Slater determinants in disguise. In III.2, we show how to compute the MPS representation of the p-wave pairing ground state of the Kitaev p+ip chain kitaev2001unpaired . We then generalize this result to general Pfaffian wave-functions which are the most general quadratic mean field ground states. Finally, we show how to take products (or powers) of quadratic mean-field Hamiltonians and turn them into (i)MPS.

III.1 BdG \rightarrow MPS

The key trick to convert a BdG wave-function into a MPS will be to 1) convert it to a Slater determinant through a particle hole transformation, 2) convert this Slater determinant to a MPS, and 3) then undo the particle-hole transformation in the MPS language.

Consider a BdG Hamiltonian:

HBdG=ij,σtij(ciσcjσ+h.c.)ijΔij(cicj+h.c.)μiσciσciσH_{BdG}=-\sum_{\langle ij\rangle,\sigma}t_{ij}\left(c^{\dagger}_{i\sigma}c_{j\sigma}+\text{h.c.}\right)-\\ \sum_{\langle ij\rangle}\Delta_{ij}\left(c^{\dagger}_{i\uparrow}c^{\dagger}_{j\downarrow}+\text{h.c.}\right)-\mu\sum_{i\sigma}c^{\dagger}_{i\sigma}c_{i\sigma} (12)

Under a canonical particle-hole transformation in the \downarrow sector:

f2i1=cif2i=ci\begin{split}f^{\dagger}_{2i-1}=c^{\dagger}_{i\uparrow}\\ f^{\dagger}_{2i}=c_{i\downarrow}\end{split} (13)

the BdG Hamiltonian becomes:

HBdGph=ijtij(f2i1f2j1f2if2j+h.c.)ijΔij(f2i1f2j+h.c.)μiσ(f2i1f2i1f2if2i)H^{ph}_{\text{BdG}}=-\sum_{\langle ij\rangle}t_{ij}\left(f^{\dagger}_{2i-1}f_{2j-1}-f^{\dagger}_{2i}f_{2j}+\text{h.c.}\right)-\\ \sum_{\langle ij\rangle}\Delta_{ij}\left(f^{\dagger}_{2i-1}f_{2j}+\text{h.c.}\right)-\mu\sum_{i\sigma}\left(f^{\dagger}_{2i-1}f_{2i-1}-f^{\dagger}_{2i}f_{2i}\right) (14)

and the new vacuum is |0ph=c1c2cN|0\ket{0^{\text{ph}}}=c^{\dagger}_{1\downarrow}c^{\dagger}_{2\downarrow}\ldots c^{\dagger}_{N\downarrow}\ket{0}, where |0\ket{0} is the vacuum of the original theory. The ground state of HBdGH_{\text{BdG}} is thus the Slater determinant ground state of HBdGphH^{\text{ph}}_{\text{BdG}} on top of the new vacuum |0ph\ket{0^{\text{ph}}}.

Using the results from section II, we convert the Slater determinant ground state of HBdGphH^{\text{ph}}_{\text{BdG}} into a MPS |ΨMPS={σ}A[1]σ1A[2N]σ2N|σ1σ2N\ket{\Psi_{MPS}}=\sum_{\{\sigma\}}A^{[1]\sigma_{1}}\ldots A^{[2N]\sigma_{2N}}\ket{\sigma_{1}\ldots\sigma_{2N}} where σ2i1={0,1}\sigma_{2i-1}=\{0,1\} (i[1,N])i\in[1,N]) indicates the absence/presence of a \uparrow particle and σ2i={0,1}\sigma_{2i}=\{0,1\} indicates the absence/presence of a hole on top of the filled \downarrow Fermi sea at site ii.

To ‘undo’ the particle-hole transformation, we need to deal with the fact that the fif_{i} act on the false vacuum |0ph\ket{0^{ph}} (and not the real vacuum) by swapping, for all ii, the matrices A[2i]1A^{[2i]1} and A[2i]0A^{[2i]0}. Moreover, by ordering the fermionic operators by site, and then spin, the matrices A[2i1]1A^{[2i-1]1}, A[2i]1A^{[2i]1} will pick up factors of (1)i1(-1)^{i-1}. We can now combine these transformations giving us our final MPS for the BdG ground state of the form |ΦGS={σ=0,,,}B[1]σ1B[N]σN|σ1σN|\Phi_{GS}\rangle=\sum_{\{\sigma={0,\uparrow,\downarrow,\uparrow\downarrow\}}}B^{[1]\sigma_{1}}\ldots B^{[N]\sigma_{N}}|\sigma_{1}\ldots\sigma_{N}\rangle where

B[i]0\displaystyle B^{[i]0} =(1)i1×A[2i1]0A[2i]1\displaystyle=(-1)^{i-1}\times A^{[2i-1]0}A^{[2i]1}
B[i]\displaystyle B^{[i]\uparrow} =(1)2i2×A[2i1]1A[2i]1\displaystyle=(-1)^{2i-2}\times A^{[2i-1]1}A^{[2i]1}
B[i]\displaystyle B^{[i]\downarrow} =(1)0×A[2i1]0A[2i]0\displaystyle=(-1)^{0}\times A^{[2i-1]0}A^{[2i]0}
B[i]\displaystyle B^{[i]\uparrow\downarrow} =(1)i1×A[2i1]1A[2i]0\displaystyle=(-1)^{i-1}\times A^{[2i-1]1}A^{[2i]0} (15)

This approach works both for the finite and infinite MPS as we just used our (i)MPS \rightarrow Slater determinant approach as a subroutine. For the infinite MPS it produces a unit cell of size 2 as every other BB differs by a sign.

As a check of our algorithm, we consider the ground state of the BdG Hamiltonian of the form:

HBdG=ij,σtij(ciσcjσ+h.c.)ijΔij(cicj+cjci+h.c.)μiσciσciσH_{BdG}=-\sum_{\langle ij\rangle,\sigma}t_{ij}\left(c^{\dagger}_{i\sigma}c_{j\sigma}+\text{h.c.}\right)-\\ \sum_{\langle ij\rangle}\Delta_{ij}\left(c^{\dagger}_{i\uparrow}c^{\dagger}_{j\downarrow}+c^{\dagger}_{j\uparrow}c^{\dagger}_{i\downarrow}+\text{h.c.}\right)-\mu\sum_{i\sigma}c^{\dagger}_{i\sigma}c_{i\sigma} (16)

and compare the amplitudes of the exact ground state with the MPS generated (see 3rd column in fig. 5).

III.2 Pfaffian \rightarrow MPS

We will show how to generate a MPS representa-tion of the Pfaffian ground state of the Kitaev p+ip chain:

Hc=tn(cncn+1+h.c.)+Δn(cncn+1+h.c.)+μncncnH_{c}=-t\sum_{n}\left(c^{\dagger}_{n}c_{n+1}+\text{h.c.}\right)+\Delta\sum_{n}\left(c^{\dagger}_{n}c^{\dagger}_{n+1}+\text{h.c.}\right)\\ +\mu\sum_{n}c^{\dagger}_{n}c_{n} (17)

We first consider Hext=HcHdH^{\text{ext}}=H_{c}\bigoplus H_{d} whose ground state is given by the tensor product of two identical pfaffians:

|GS=σc,σdPf(Mσc)|σcPf(Mσd)|σd\ket{GS}=\sum_{\sigma_{c},\sigma_{d}}Pf(M_{\sigma_{c}})\ket{\sigma_{c}}\bigotimes Pf(M_{\sigma_{d}})\ket{\sigma_{d}} (18)

where MM is a N×NN\times N matrix built from parameters of the model and MσcM_{\sigma_{c}} is a submatrix of MM obtained by selecting indices as given by the σc\sigma_{c} configuration.

Using the local canonical transformation of fermions,

cn=(c¯n,+c¯n,)2dn=i(c¯n,c¯n,)2\begin{split}c^{\dagger}_{n}=\frac{\left(\bar{c}^{\dagger}_{n,\uparrow}+\bar{c}^{\dagger}_{n,\downarrow}\right)}{\sqrt{2}}\\ d^{\dagger}_{n}=i\frac{\left(\bar{c}^{\dagger}_{n,\uparrow}-\bar{c}^{\dagger}_{n,\downarrow}\right)}{\sqrt{2}}\end{split} (19)

converts HextH^{\text{ext}} to

HBdG=tn,σ(c¯n,σc¯n+1,σ+h.c.)+Δn(c¯nc¯n+1+c¯n+1,c¯n,+h.c.)+μnσc¯n,σc¯n,σH_{BdG}=-t\sum_{n,\sigma}\left(\bar{c}^{\dagger}_{n,\sigma}\bar{c}_{n+1,\sigma}+\text{h.c.}\right)\\ +\Delta\sum_{n}\left(\bar{c}^{\dagger}_{n\uparrow}\bar{c}^{\dagger}_{n+1\downarrow}+\bar{c}^{\dagger}_{n+1,\downarrow}\bar{c}^{\dagger}_{n,\uparrow}+\text{h.c.}\right)\\ +\mu\sum_{n\sigma}\bar{c}^{\dagger}_{n,\sigma}\bar{c}_{n,\sigma} (20)

The transformation leaves the vacuum unchanged. Given a BdG Hamiltonian, we can obtain the ground-state as an MPS as done in section III.1.

|GS=σA2i1σ2i1A2iσ2i|σ1σ2σ2N\ket{GS}=\sum_{\sigma}\ldots A^{2i-1\sigma_{2i-1}}A^{2i\sigma_{2i}}\ldots\ket{\sigma_{1}\sigma_{2}\ldots\sigma_{2N}} (21)

where A2i1σ2i1A^{2i-1\sigma_{2i-1}} is the tensor on site ii for the \uparrow local physical sector and A2iσ2iA^{2i\sigma_{2i}} is the tensor on site ii for the \downarrow local physical sector.

Notice that the canonical transformation given in eqn. 19 mixes the \uparrow, \downarrow physical sectors on site ii. Hence we can obtain the MPS for the GS in the c,dc,d space by choosing the on-site tensor in the following way:

|GS=σC1σ1C2σ2CNσN|σ1σ2σN\ket{GS}=\sum_{\sigma}C^{1\sigma_{1}}C^{2\sigma_{2}}\ldots C^{N\sigma_{N}}\ket{\sigma_{1}\sigma_{2}\ldots\sigma_{N}} (22)

with σ={0,c,d,cd}\sigma=\{0,c,d,cd\} where:

Cn,0=A2n1,0A2n,0Cn,c=A2n1,1A2n,0+A2n1,1A2n,02Cn,d=iA2n1,1A2n,0A2n1,1A2n,02Cn,cd=iA2n1,1A2n,1\begin{split}C^{n,0}=A^{2n-1,0}A^{2n,0}\\ C^{n,c}=\frac{A^{2n-1,1}A^{2n,0}+A^{2n-1,1}A^{2n,0}}{\sqrt{2}}\\ C^{n,d}=i\frac{A^{2n-1,1}A^{2n,0}-A^{2n-1,1}A^{2n,0}}{\sqrt{2}}\\ C^{n,cd}=iA^{2n-1,1}A^{2n,1}\end{split} (23)

By projecting out the dd particles in the |GS\ket{GS} wavefunction, we obtain a Pfaffian wavefunction in the cc-particle sector: |GS=constant×σcPf(Mσc)|σc\ket{GS}=\text{constant}\times\sum_{\sigma_{c}}Pf(M_{\sigma_{c}})\ket{\sigma_{c}}. At the level of the MPS this projection is realized by eliminating the sectors σ={d,cd}\sigma=\{d,cd\}.

|Pf=σ={0,c}C1σ1C2σ2CNσN|σ1σ2σN\ket{\text{Pf}}=\sum_{\sigma=\{0,c\}}C^{1\sigma_{1}}C^{2\sigma_{2}}\ldots C^{N\sigma_{N}}\ket{\sigma_{1}\sigma_{2}\ldots\sigma_{N}} (24)

III.3 Pfaffian \rightarrow MPS Generalization

While we focused in the previous section on a specific example, here we consider a generic quadratic Hamiltonian H=n,mCnhn,mCmH=\sum_{n,m}C^{\dagger}_{n}h_{n,m}C_{m} with gg species of fermions per unit cell where the vector Cn=(cn,1,cn,1,cn,2,cn,2cn,g,cn,g)C^{\dagger}_{n}=\left(c^{\dagger}_{n,1},c_{n,1},c^{\dagger}_{n,2},c_{n,2}\ldots c^{\dagger}_{n,g},c_{n,g}\right).

We form an extended Hamiltonian HextH^{\text{ext}} which is a sum of two copies of HH:

Hext=hiα,jβhop(ci,αcj,β+di,αdj,β+h.c.)+hiα,jβpair(ci,αcj,β+di,αdj,β+h.c.)H^{ext}=\sum h^{\text{hop}}_{i\alpha,j\beta}\left(c^{\dagger}_{i,\alpha}c_{j,\beta}+d^{\dagger}_{i,\alpha}d_{j,\beta}+\text{h.c.}\right)\\ +\sum h^{\text{pair}}_{i\alpha,j\beta}\left(c^{\dagger}_{i,\alpha}c^{\dagger}_{j,\beta}+d^{\dagger}_{i,\alpha}d^{\dagger}_{j,\beta}+\text{h.c.}\right) (25)

where α,β(1,,g)\alpha,\beta\in(1,\ldots,g). As before, its ground state |GSext=σc,σdPf(Mσc)|σcPf(Mσd)|σd\ket{\text{GS}}^{\text{ext}}=\sum_{\sigma_{c},\sigma_{d}}Pf(M_{\sigma_{c}})\ket{\sigma_{c}}\bigotimes Pf(M_{\sigma_{d}})\ket{\sigma_{d}} is a tensor product of two identical Pfaffian wavefunctions. We then obtain the Pfaffian ground state of HH by projecting out all the dd sectors. Under the following linear canonical transformation

c¯n,α,=cn,α+idn,α2c¯n,α,=cn,αidn,α2\begin{split}\bar{c}^{\dagger}_{n,\alpha,\uparrow}=\frac{c^{\dagger}_{n,\alpha}+id^{\dagger}_{n,\alpha}}{\sqrt{2}}\\ \bar{c}^{\dagger}_{n,\alpha,\downarrow}=\frac{c^{\dagger}_{n,\alpha}-id^{\dagger}_{n,\alpha}}{\sqrt{2}}\end{split} (26)

HextH^{\text{ext}} becomes a BdG-like hamiltonian when expressed in terms of cc_{\uparrow} and cc_{\downarrow}:

HBdGext=hiα,jβhop(c¯i,α,c¯j,β,+c¯i,α,c¯j,β,+h.c.)++hiα,jβpair(c¯i,α,c¯j,β,+c¯i,α,c¯j,β,+h.c.)\begin{split}H^{ext}_{\text{BdG}}=\sum h^{\text{hop}}_{i\alpha,j\beta}\left(\bar{c}^{\dagger}_{i,\alpha,\uparrow}\bar{c}_{j,\beta,\uparrow}+\bar{c}^{\dagger}_{i,\alpha,\downarrow}\bar{c}_{j,\beta,\downarrow}+\text{h.c.}\right)+\\ +\sum h^{\text{pair}}_{i\alpha,j\beta}\left(\bar{c}^{\dagger}_{i,\alpha,\uparrow}\bar{c}^{\dagger}_{j,\beta,\downarrow}+\bar{c}^{\dagger}_{i,\alpha,\downarrow}\bar{c}^{\dagger}_{j,\beta,\uparrow}+\text{h.c.}\right)\end{split} (27)

We can then solve for the MPS representation of the groundstate of the above BdG-like Hamiltonian using the methods described in section  III.1.

We obtain the Slater determinant ground state Ψext|\bra{\Psi_{ext}} by diagonalizing the particle-hole transformed HBdGextH^{\text{ext}}_{\text{BdG}} and then computing its MPS representation. Each unit cell MM is described by 2g2g tensors A[M,p]σA^{[M,p]\sigma} with σ=0,1\sigma=0,1 signifying the absence/presence of a particle of type p[0,1,2g1]p\in[0,1,\ldots 2g-1]. A particle of type p=2kp=2k corresponds to the flavor k,k,\uparrow; a particle of type p=2k+1p=2k+1 corresponds to the flavor k,k,\downarrow. From the above MPS (which is in c¯\bar{c}_{\uparrow}, c¯\bar{c}_{\downarrow} local physical space) we construct the MPS tensors in cc, dd space. In particular, the matrices describing the absence/presence of a particle of type cic_{i} on site MM are given by:

B[M,i]0\displaystyle B^{[M,i]0} =A[M,2i1][0]A[M,2i][1]\displaystyle=A^{[M,2i-1][0]}A^{[M,2i][1]} (28)
B[M,i]1\displaystyle B^{[M,i]1} =(1)g(M1)+i12×\displaystyle=\frac{(-1)^{g(M-1)+i-1}}{\sqrt{2}}\times
[A[M,2i1][1]A[M,2i][1]+A[M,2i1]0A[M,2i][0]]\displaystyle\left[A^{[M,2i-1][1]}A^{[M,2i][1]}+A^{[M,2i-1]0}A^{[M,2i][0]}\right]

where (1)g(M1)+i1(-1)^{g(M-1)+i-1} takes care of fermionic ordering. and the MPS representation of |ΨGS\ket{\Psi_{GS}} defined on NgNg sites is given by:

|ΨGS={σ}(B[1,1]σ11B[1,2]σ12B[1,g]σ1g)×(B[N,1]σN1B[N,2]σN2B[N,g]σNg)×|(σ11σ12σ1g)(σN1σN2σNg)\ket{\Psi_{GS}}=\sum_{\{\sigma\}}\left(B^{[1,1]\sigma_{1_{1}}}B^{[1,2]\sigma_{1_{2}}}\ldots B^{[1,g]\sigma_{1_{g}}}\right)\ldots\times\\ \left(B^{[N,1]\sigma_{N_{1}}}B^{[N,2]\sigma_{N_{2}}}\ldots B^{[N,g]\sigma_{N_{g}}}\right)\times\\ \ket{\left(\sigma_{1_{1}}\sigma_{1_{2}}\ldots\sigma_{1_{g}}\right)\ldots\left(\sigma_{N_{1}}\sigma_{N_{2}}\ldots\sigma_{N_{g}}\right)} (29)

By suitably contracting tensors we can obtain a NN-tensor MPS representation with physical dimension 2g2^{g}: |Pf=σC1σ1C2σ2CNσN|σ1σ2σN\ket{Pf}=\sum_{\sigma}C^{1\sigma_{1}}C^{2\sigma_{2}}\ldots C^{N\sigma_{N}}\ket{\sigma_{1}\sigma_{2}\ldots\sigma_{N}}. For instance, the tensor corresponding to the presence of particles of type t1,t2,tst_{1},t_{2},\ldots t_{s} on site MM is

C[M](t1,t2,,ts)=B[M,1][0]×B[M,t1][1]B[M,t1+1][0]B[M,ts][1]B[M,g][0]C^{[M](t_{1},t_{2},\ldots,t_{s})}=B^{[M,1][0]}\ldots\times\\ B^{[M,t_{1}][1]}B^{[M,t_{1}+1][0]}\ldots B^{[M,t_{s}][1]}\ldots B^{[M,g][0]} (30)

III.4 Power of Slater determinants

In this section we describe how to obtain the MPS representation of a wavefunction

|ψ1/n=r1,r2,rnr1,r2,rn|ψn|r1,r2,,rn\ket{\psi_{1/n}}=\sum_{r_{1},r_{2},\cdots r_{n}}\braket{r_{1},r_{2},\ldots r_{n}}{\psi}^{n}|r_{1},r_{2},\ldots,r_{n}\rangle (31)

where |ψ\ket{\psi} is a Slater determinant. We will use as an example n = 3. Products of other mean-field wave-functions can be obtained similarly.

We extend our N-site system to a 3N3N-site system for which we label the sites as: {111213212223N1N2N3}\{1_{1}1_{2}1_{3}2_{1}2_{2}2_{3}\ldots N_{1}N_{2}N_{3}\}. We then write a single Slater determinant (by padding and interlacing the orbitals to keep the above ordering) of the form |ψ|ψ|ψ\ket{\psi}\bigotimes\ket{\psi}\bigotimes\ket{\psi} for which we then convert into an MPS given by

|MPS=ip(B[11]i11B[12]i12B[13]i13)×(B[N1]iN1B[N2]iN2B[N3]iN3)×|i11i12i13iN1iN2iN3\ket{MPS}=\sum_{{i_{p}}}\left(B^{[1_{1}]i_{1_{1}}}B^{[1_{2}]i_{1_{2}}}B^{[1_{3}]i_{1_{3}}}\right)\ldots\times\\ \left(B^{[N_{1}]i_{N_{1}}}B^{[N_{2}]i_{N_{2}}}B^{[N_{3}]i_{N_{3}}}\right)\times\\ \ket{i_{1_{1}}i_{1_{2}}i_{1_{3}}\ldots i_{N_{1}}i_{N_{2}}i_{N_{3}}} (32)

Projecting on the sector in1=in2=in3i_{n_{1}}=i_{n_{2}}=i_{n_{3}} gives us the desired results of

|ψ1/3=A[1]i1A[2]i2A[N]iN|i1i2iN\ket{\psi_{1/3}}=A^{[1]i_{1}}A^{[2]i_{2}}\ldots A^{[N]i_{N}}\ket{i_{1}i_{2}\ldots i_{N}} (33)

where we define

A[n]in=B[n1]in1B[n2]in2B[n3]in3A^{[n]i_{n}}=B^{[n_{1}]i_{n_{1}}}B^{[n_{2}]i_{n_{2}}}B^{[n_{3}]i_{n_{3}}} (34)

IV Bilinear-Biquadratic S=1 model

In this section, we use our approach to compute the MPS representation and entanglement spectra of the Gutzwiller projected slave-fermion mean-field states liu2012gutzwiller of the Bilinear-Biquadratic S=1 model,

H=J12+J22<i,j>(cosθ𝐒i𝐒j+sinθ(𝐒i𝐒j)2),H=\sqrt{J_{1}^{2}+J_{2}^{2}}\sum_{<i,j>}\left(\cos{\theta}\mathbf{S}_{i}\cdot\mathbf{S}_{j}+\sin{\theta}\left(\mathbf{S}_{i}\cdot\mathbf{S}_{j}\right)^{2}\right), (35)

The physics of the 1D quantum Heisenberg spin-chain is qualitatively different for different spin representationshaldane1983continuum ; half-integer spins have a gapless ground state and power-law spin correlations; integer spins have a gapped ground state with exponentially decaying correlations, the Haldane/AKLT phaseaffleck1988valence . This latter phase is robust due to a combination of symmetries which protect its topological propertiesgu2009tensor ; pollmann2012symmetry . This symmetry protection can be understood in terms of “fractionalization”: a S=1S=1 spin effectively splits into two S=1/2S=1/2 edge modes that transform under non-trivial projective representations of the symmetries (the product of the symmetry representations differs from the representation of the product). These features are reflected by non-trivial degeneracies in the entanglement spectrum turner2011topological ; pollmann2010entanglement - i.e. the eigenvalues of HentH_{\textrm{ent}} in ρA=eHent\rho_{A}=\text{e}^{-H_{\textrm{ent}}} where ρA=TrB|ψψ|\rho_{A}=Tr_{B}\ket{\psi}\bra{\psi} is the reduced density matrix on an A subsystem li2008entanglement ; levin2006detecting ; kitaev2006topological . The BLBQ model has four phases as shown in fig. 6. This includes the Haldane phase (at the Heisenberg point), as well as a dimerized and critical phase.

One can derive the relevant projected mean field state from the slave-fermion construction by fractionalizing the spin operators S^\hat{\textbf{S}} in terms of fermionic parton operators:

S^i=fi;αSαβfi;β\hat{\textbf{S}}_{i}=f^{\dagger}_{i;\alpha}\textbf{S}_{\alpha\beta}f_{i;\beta} (36)

where fi;αf^{\dagger}_{i;\alpha} is the α\alpha-flavor fermionic parton creation operator at site ii and Sαβ\textbf{S}_{\alpha\beta} are the matrix elements of the spin operators in a given representation SS liu2010fermionic . Substituting these expressions into the original Hamiltonian gives a quartic fermionic Hamiltonian HfH_{f} which can be decoupled through a mean-field. The resulting mean-field ground state must then be projected back into the original Hilbert space essentially ‘glueing’ together the fractionalized degrees of freedom.

The slave-fermion construction of the Bilinear-Biquadratic model was studied in ref. liu2012gutzwiller, where the authors used VMC to optimize the (χ,δ,μ)(\chi,\delta,\mu) parameters of the projected wavefunction and studied the energy of this slave-fermion state compared to the exact energy achieved by TEBD. In this section, we will take the same points as studied in ref. liu2012gutzwiller, ; liu2014gutzwiller, , convert the slave-fermion states to MPS, and compute the entanglement spectra and the energies.

At the level of the Gutzwiller projected groundstates, the two gapped phases, the dimer phase and the AKLT phase, are distinguished by their “fingerprint” in the low-lying structure of the entanglement spectrum: the lowest level of the entanglement spectrum of the dimerized topologically trivial phase is singly degenerate (for between dimer cuts); the lowest level of the entanglement spectrum of the Haldane phase ground state is doubly degenerate, corresponding to the presence of two boundary S=1/2S=1/2 edge modes.

Refer to caption
Figure 6: Phase diagram of the bilinear-biquadratic S=1S=1 model. Figure reproduced from ref. liu2012gutzwiller, .

IV.1 Generating the MPS

The relevant mean-field Hamiltonian that arises from the parton construction of the bilinear-biquadratic S=1 model liu2010fermionic ,liu2014gutzwiller is

Hmf=Jχi,α=1,0,1[ci,αci,α+h.c.]+(JK)Δi,j[ci,1cj,1c0ic0j+c1ic1j+h.c.]+λi,αciαciαH_{mf}=-J\chi\sum_{i,\alpha=-1,0,1}\left[c^{\dagger}_{i,\alpha}c_{i,\alpha}+\text{h.c.}\right]\\ +(J-K)\Delta\sum_{i,j}\left[c^{\dagger}_{i,-1}c^{\dagger}_{j,1}-c^{\dagger}_{0i}c^{\dagger}_{0j}+c^{\dagger}_{1i}c^{\dagger}_{-1j}+\text{h.c.}\right]\\ +\lambda\sum_{i,\alpha}c^{\dagger}_{i\alpha}c_{i\alpha} (37)

where the c1c^{\dagger}_{-1}, c0c^{\dagger}_{0} and c1c^{\dagger}_{1} are the on-site fermion parton flavors corresponding to Sz=1,0,1S_{z}=-1,0,1.

This could be converted into a MPS by treating it as a general Pfaffian and then applying the techniques in section  III.2. In models such as this, though, where the mean-field Hamiltonian in the parton basis has a tensor sum structure where one or more of the Hilbert subspaces can be treated with a simpler mean-field (i.e. with a SD or BdG ground state) it makes computational sense to obtain the MPS representation in each sector and then “glue” the two MPS together; we exemplify this approach here.

Introducing a Nambu spinor, in the kk-basis the Hamiltonian is block-diagonal:

Hmfk=12[ck,1ck,1ck,0ck,0]×\displaystyle H^{k}_{mf}=\frac{1}{2}\begin{bmatrix}c^{\dagger}_{k,1}&c_{-k,-1}&c^{\dagger}_{k,0}&c_{-k,0}\end{bmatrix}\times
[χkΔk00Δkχk0000χkΔk00Δkχk][ck,1ck,1ck,0ck,0]\displaystyle\begin{bmatrix}\chi_{k}&\Delta_{k}&0&0\\ \Delta_{k}^{*}&-\chi_{k}&0&0\\ 0&0&\chi_{k}&-\Delta_{k}\\ 0&0&-\Delta_{k}^{*}&-\chi_{k}\end{bmatrix}\begin{bmatrix}c_{k,1}\\ c^{\dagger}_{-k,-1}\\ c_{k,0}\\ c^{\dagger}_{-k,0}\end{bmatrix} (38)

The one-body hamiltonian is a tensor sum of BdG-like Hamiltonian HBdGH_{BdG} and a p+ipp+ip Hamiltonian Hp+ipH_{p+ip}. The mean field ground state is (we consider anti-periodic boundary conditions and an even number of sites):

|ΨGS=Π0<k<2π(uk+vkck,1ck,1)×Π0<q<π(uqvqcq,0cq,0)|0\ket{\Psi_{GS}}=\Pi_{0<k<2\pi}\left(u_{k}+v_{k}c^{\dagger}_{k,1}c^{\dagger}_{-k,-1}\right)\times\\ \Pi_{0<q<\pi}\left(u_{q}-v_{q}c^{\dagger}_{q,0}c^{\dagger}_{-q,0}\right)\ket{0} (39)

where uku_{k} and vkv_{k} are given in terms of the parameters of the Hamiltonian.

By performing a particle-hole transformation in the Sz={,}S_{z}=\{\uparrow,\downarrow\} sector (see section III.1) and the Pfaffian artificial extension in the Sz=0S_{z}=0 sector (see section III.2),

f1,k\displaystyle f^{\dagger}_{1,k} =ck,1\displaystyle=c^{\dagger}_{k,1} (40)
f2,k\displaystyle f^{\dagger}_{2,k} =ck,1\displaystyle=c_{-k,-1}
f3,k\displaystyle f^{\dagger}_{3,k} =ck,0\displaystyle=c^{\dagger}_{k,0}
f4,k\displaystyle f^{\dagger}_{4,k} =ck,0\displaystyle=c_{-k,0}

where {fk,α,fq,β}=δαβδkq\{f_{k,\alpha},f_{q,\beta}\}=\delta_{\alpha\beta}\delta_{kq}, giving us

|ΨGSext=Πk(ukfk,2+vkfk,1)Π0<q<Π(uqfq,4vqfk,3)|vac\ket{\Psi_{GS}^{ext}}=\Pi_{k}\left(u_{k}f^{\dagger}_{k,2}+v_{k}f^{\dagger}_{k,1}\right)\\ \Pi_{0<q<\Pi}\left(u_{q}f^{\dagger}_{q,4}-v_{q}f^{\dagger}_{k,3}\right)\ket{\text{vac}} (41)

with |vac=Π0<k<2πck|0Π0<q<πck,0|0\ket{\text{vac}}=\Pi_{0<k<2\pi}c_{k\downarrow}\ket{0}\Pi_{0<q<\pi}c_{k,0}\ket{0}. Since |ΨGSext\ket{\Psi^{ext}_{GS}} is a Slater Determinant, we can obtain the MPS representation using the methods in section II. We now “undo” the transformation (see again section III.1 and section III.2) and write the MPS in the following form:

|MPS={i}(C[1]i1C[1]i1C[1,]i1,)×(C[N]iNC[N]iNC[N,]iN)|(i1i1i1)(iNiNiN)\ket{MPS}=\sum_{\{i\}}\left(C^{[1_{\uparrow}]i_{1\uparrow}}C^{[1_{\downarrow}]i_{1\downarrow}}C^{[1,\rightarrow]i_{1,\rightarrow}}\right)\times\\ \ldots\left(C^{[N_{\uparrow}]i_{N\uparrow}}C^{[N_{\downarrow}]i_{N\downarrow}}C^{[N,\rightarrow]i_{N_{\rightarrow}}}\right)\\ \ket{\left(i_{1_{\uparrow}}i_{1_{\downarrow}}i_{1_{\rightarrow}}\right)\ldots\left(i_{N_{\uparrow}}i_{N_{\downarrow}}i_{N_{\rightarrow}}\right)} (42)

where in,α{0,1}i_{n,\alpha}\in\{0,1\} indicates the absence/presence of a particle of type α\alpha on site nn.

To obtain the MPS with onsite tensors A[n]σn,σn{,,,,,,}A^{[n]\sigma_{n}},\sigma_{n}\in\{\uparrow,\downarrow,\rightarrow,\uparrow\rightarrow,\downarrow\rightarrow,\uparrow\downarrow,\uparrow\downarrow\rightarrow\}, we ”glue” together appropriate sectors. For example

A[n]=C[n]1C[n]0C[n]0A^{[n]\uparrow}=C^{[n\uparrow]1}C^{[n\downarrow]0}C^{[n\rightarrow]0} (43)

Gutzwiller projection is realized by summing only over the one-particle per site physical indices σn{,,}\sigma_{n}\in\{\uparrow,\downarrow,\rightarrow\}:

PG|ΨGS=σn{,,}A[1]σ1A[N]σN|σ1σNP_{G}\ket{\Psi_{GS}}=\sum_{\sigma_{n}\in\{\uparrow,\downarrow,\rightarrow\}}A^{[1]\sigma_{1}}\ldots A^{[N]\sigma_{N}}\ket{\sigma_{1}\ldots\sigma_{N}} (44)

IV.2 iMPS orthogonalization

In this section, we will discuss orthogonalizing our iMPS states. This includes a brief overview of the standard iMPS orthogonalization as well as a detailed description of how we address the degeneracies that appear when Gutzwiller projecting slave-fermion mean-field states onto degenerate ground state manifolds.

The orthogonalization procedure for a typical iMPS is standard (see ref. orus2008infinite, and supplement  S6 for more details). The method relies on obtaining the leading right/left eigenvectors of the transfer matrix operator E=σAσAσE=\sum_{\sigma}A^{\sigma}\otimes A^{\sigma*} where σ\sigma runs over the on-site physical index. EE admits the following decomposition:

E=iλi|RiL|iE=\sum_{i}\lambda_{i}\ket{R}_{i}\bra{L}_{i} (45)

where |Li\ket{L}_{i} and |Ri\ket{R}_{i} are left/right eigenvectors of EE and Li|Ri=0\braket{L_{i}}{R_{i}}=0 for λi\lambda_{i} non-degenerate. In the infinite limit only the leading left/right eigenvectors of EE survive. If the dominant eigenvalue is non-degenerate, the transfer matrix is given by

limNEN=|RL|\lim_{N\rightarrow\infty}E^{N}=\ket{R}\bra{L} (46)

where |R(L)\ket{R(L)} are by definition the eigenvectors corresponding to the dominant eigenvalue. Thus, the dominant left and right eigenvectors correspond to a pure state. The Implicitly Restarted Arnoldi Method can be efficiently used for this purpose by noting that (σAσAσ)vec(v)=σvec(AσvAT)\left(\sum_{\sigma}A^{\sigma}\otimes A^{\sigma*}\right)\mathrm{vec}(v)=\sum_{\sigma}\mathrm{vec}\left(A^{\sigma*}vA^{T}\right), where the vec(v)\mathrm{vec}(v) operation takes the square matrix vv and stacks the columns together. The entanglement spectrum and observables are then easily obtained.

When the leading eigenvalues of the transfer matrix are degenerate in magnitude

limNEN=|R1L1|+|R2L2|\lim_{N\rightarrow\infty}E^{N}=\ket{R_{1}}\bra{L_{1}}+\ket{R_{2}}\bra{L_{2}} (47)

ENE^{N} is in mixed form. In general the output of the Arnoldi method gives Li|Ri0\braket{L_{i}}{R_{i}}\neq 0. Thus, additional steps are required to obtain the canonical form of the iMPS. The degeneracy of the leading eigenvalues signals the presence of degenerate states. This is indeed what happens for the 2-fold degenerate dimer phase and the 4-fold degenerate Haldane phase (in the thermodynamic limit). In order to access all the states in the ground state manifold, we need to obtain the proper set of pure iMPS states. The transfer matrix of a pure iMPS has unique left/right leading eigenvectors.

Here we consider the case of two-fold degeneracy present in the dimer phase states (other states and higher degeneracies can be dealt with using a similar procedure). For a two-fold degenerate iMPS, we need to find two pure iMPS generated by bulk tensors A1A_{1} and A2A_{2}. Any iMPS within the degenerate manifold is then able to be written as a linear superposition of these pure states: |ψB=α1|ψ(A1)+α2|ψ(A2)\ket{\psi_{B}}=\alpha_{1}\ket{\psi(A_{1})}+\alpha_{2}\ket{\psi(A_{2})} where the notation |ψ(B)\ket{\psi(B)} indicates the iMPS generated by bulk tensor BB. Note that the entanglement spectrum of the reduced density matrix ρB=|α1|2ρA1+|α2|2ρA2\rho_{B}=|\alpha_{1}|^{2}\rho_{A_{1}}+|\alpha_{2}|^{2}\rho_{A_{2}} is given by the combined spectra of α1ρ1\alpha_{1}\rho_{1} and α2ρ2\alpha_{2}\rho_{2}.

To generate these pure iMPS, we start from a non-canonical bulk tensor AA iMPS with a single site unit cell (typically generated by projection). In the case of the dimer phase, this bulk tensor has two leading right (respectively left) eigenvectors, v1v_{1} and v2v_{2}, with equal magnitude eigenvalues |η1|=|η2||\eta_{1}|=|\eta_{2}|, but different signs (i.e. η1=η2\eta_{1}=-\eta_{2}).

According to Theorem 5 in ref. perez2006matrix, and Theorem 11 in ref. ruiz2011tensor, , there is a unitary that transforms each of the matrices Bσσ=AσAσB^{\sigma\sigma^{\prime}}=A^{\sigma}A^{\sigma^{\prime}} into block diagonal form, with two blocks; the two blocks are the 2-site uniform tensors corresponding to the two pure states.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 7: Illustration of decomposition of a mixed transfer matrix into pure components. In step 1 (first line) we find the dominant left-eigenvector of the transfer matrix; In step 4 (second line), we find ULU_{L} and S~1\tilde{S}_{1} and S~2\tilde{S}_{2}. In step 5 (third, fourth line and fifth line), we form two new tensors, B~i\tilde{B}_{i} and find their right leading eigenvectors. In step 6 (other lines), we find the entanglement spectrum Λi\sqrt{\Lambda_{i}} and the right/left canonical matrices, BiRB^{R}_{i} and BiLB^{L}_{i}, corresponding to the two pure states.

Based on the mathematical theorems in refs. perez2006matrix, and ruiz2011tensor, , we use the following procedure to compute the pure states (see fig. 7):

  1. 1.

    Start with the D×DD\times D (DD is the bond dimension of the bulk tensors AA) left leading eigenvectors (with the same eigenvalue), V1LV^{L}_{1} and V2LV^{L}_{2}, of the completely positive map E2E^{2}, i.e. σ,σBσσViLBσ=ViL\sum_{\sigma,\sigma^{\prime}}B^{\dagger\sigma\sigma^{\prime}}V^{L}_{i}B^{\sigma^{\prime}}=V^{L}_{i}.

  2. 2.

    ViLV^{L}_{i} are transformed into hermitian matrices: ViL:=1/2(ViL+(ViL))V^{L}_{i}:=1/2(V^{L}_{i}+(V^{L}_{i})^{\dagger}); this is possible because if ViLV^{L}_{i} is an eigenvector of E2E^{2}, then (ViL)(V^{L}_{i})^{\dagger} is also an eigenvector and so their sum is Hermitian. If ViL=UiDiUiV^{L}_{i}=U_{i}D_{i}U^{\dagger}_{i}, then we can write ViL=YiYiV^{L}_{i}=Y^{\dagger}_{i}Y_{i} with Yi=DiUiY_{i}=\sqrt{D_{i}}U^{\dagger}_{i}.

  3. 3.

    Diagonalize V1LV^{L}_{1} and V2LV^{L}_{2} together; this can be done since [V1L,V2L]=0[V^{L}_{1},V^{L}_{2}]=0 so that UViLU=SiU^{\dagger}V^{L}_{i}U=S_{i}, with SiS_{i} being a diagonal matrix.

  4. 4.

    Form two linear combinations ViL~=V1LαiV2L\tilde{V^{L}_{i}}=V^{L}_{1}-\alpha_{i}V^{L}_{2} where αi\alpha_{i} is one of the two non-zero values obtained by the elementwise division of S1S_{1} and S2S_{2}. Then ULV1L~ULU^{\dagger}_{L}\tilde{V^{L}_{1}}U_{L} will be a diagonal matrix S1~\tilde{S_{1}} with entries (d~11,d~12,,d~1p,0,0,0)\left(\tilde{d}^{1}_{1},\tilde{d}^{2}_{1},\ldots,\tilde{d}^{p}_{1},0,0,\ldots 0\right) and ULV2L~ULU^{\dagger}_{L}\tilde{V^{L}_{2}}U_{L} a diagonal matrix S2~\tilde{S_{2}} with entries (0,0,,0,d~2Dk,d~2Dk1,,d~2D,)\left(0,0,\ldots,0,\tilde{d}^{D-k}_{2},\tilde{d}^{D-k-1}_{2},\ldots,\tilde{d}^{D}_{2},\right) and DkpD-k\geq p; in fact, it will almost always be the case that Dk>pD-k>p, since the bond dimension of the canonical bulk tensor decreases after projection; this decomposition is guaranteed by Theorem 5 in ref. perez2006matrix, .

  5. 5.

    Form two new 22-site bulk tensors B~i=S~iULBULS~i1\tilde{B}_{i}=\sqrt{\tilde{S}_{i}}U^{\dagger}_{L}BU_{L}\sqrt{\tilde{S}_{i}}^{-1} and obtain their transfer matrix right leading eigenvectors; they will each have a unique leading hermitian semi-positive definite diagonal eigenvector, ViR=UR,iΛi(UR,i)V^{R}_{i}=U_{R,i}\Lambda_{i}\left(U_{R,i}\right)^{\dagger}; we can write ViR=XiXiV^{R}_{i}=X_{i}X^{\dagger}_{i} with Xi=UR,iΛX_{i}=U_{R,i}\sqrt{\Lambda}; then Λi\sqrt{\Lambda_{i}} is the entanglement spectrum of the corresponding pure state.

  6. 6.

    BiR=Λi1(UR,i)B~iUR,iΛiB^{R}_{i}=\sqrt{\Lambda_{i}}^{-1}\left(U_{R,i}\right)^{\dagger}\tilde{B}_{i}U_{R,i}\sqrt{\Lambda_{i}} are the right canonical tensors and BiL=(UR,i)B~iUR,iB^{L}_{i}=\left(U_{R,i}\right)^{\dagger}\tilde{B}_{i}U_{R,i} are the left canonical tensors; since BiLΛi=ΛiBiRB^{L}_{i}\sqrt{\Lambda_{i}}=\sqrt{\Lambda_{i}}B^{R}_{i} the uniform 22-site translationally invariant bulk tensors can be written as Ai=ΛiBiRΛi1=Λi1BiLΛiA_{i}=\sqrt{\sqrt{\Lambda_{i}}}B^{R}_{i}\sqrt{\sqrt{\Lambda_{i}}}^{-1}=\sqrt{\sqrt{\Lambda_{i}}}^{-1}B^{L}_{i}\sqrt{\sqrt{\Lambda_{i}}}.

(1,1)ULS (1,13\frac{1}{3})AKLT (1,0)Heisenberg (1,-1)TB (1,-2) (1,-3) (0,-1) (-1,-3) (-1,-2)
itebd [liu2012gutzwiller, ] 0.2971 23-\frac{2}{3} -1.4015 -4 -6.7531 -9.5330 -2.7969 -7.3518 -4.5939
dmrg 0.2978 23-\frac{2}{3} -1.4015 -3.9999 -6.7526 -9.5314 -2.7969 -7.3516 -4.5939
VMC [liu2012gutzwiller, ] 0.2997 23-\frac{2}{3} -1.4001 -3.9917 -6.7372 -9.5103 -2.7953 -7.2901 -4.4946
±0.0004\pm 0.0004 ±7×1015\pm 7\times 10^{-15} ±0.0004\pm 0.0004 ±0.0012\pm 0.0012 ±0.0023\pm 0.0023 ±0.0034\pm 0.0034 ±0.0005\pm 0.0005 ±0.0038\pm 0.0038 ±0.0028\pm 0.0028
fMPS 0.2995 23-\frac{2}{3} -1.3999 -3.9895 -6.7369 -9.5073 -2.7948 -7.2877 -4.4935
iMPS 23-\frac{2}{3} -1.3999 -6.7368 -9.5071 -2.7947 -7.2877 -4.4934
χ\chi 1 1 1 1 1 1 0 0 0
Δ\Delta 0 32\frac{3}{2} 0.98 1.11 1.15 1.79 1 1 1
λ\lambda 1 0 1.78 2.00 2.07 2.22 0.14 0.21 0.12
Table 1: Comparison of energies per site between the exact ground state (iTEBD/DMRG), variational Monte Carlo (VMC) and the fMPS and iMPS generated from the projected slave-fermion states. fMPS are computed with N=64N=64 or N=96N=96 and 104ϵ3×10410^{-4}\leq\epsilon\leq 3\times 10^{-4}. The bulk tensors used in iMPS have bond dimension D1000D\approx 1000. Column headings correspond to (J,K)(J,K)

IV.3 Energy of BLBQ Slave-Fermion Wavefunctions

We compute both the MPS and iMPS (except at the critical points) for the variational Gutzwiller projected wavefunctions corresponding (as found in ref. liu2012gutzwiller, by minimizing the variational energy) to the points in fig. 6. We directly compare the energy for all of these points (see table 1) and find that the energies are all within the error bars reported for the VMC calculationliu2012gutzwiller .

IV.4 Entanglement Spectra of BLBQ Slave-Fermion Wavefunctions

IV.4.1 Dimer phase

In this section, we will consider entanglement spectra of the dimerized phase of the BLBQ model. The ground state of the dimerized phase is two-fold degenerate depending on whether the dimer covering spans even or odd bonds; the entanglement spectra also depends on whether the entanglement cut is made through or between dimers. For the fMPS, we can obtain both the even and odd cut entanglement spectra of the dimerized states by choosing two consecutive cuts whereas for the iMPS we use the procedure described in sec. IV.2 to find the two pure states which corresponds respectively to the even and odd cuts.

We start by considering a generic slave-fermion point in the BLBQ model; see fig. 8 for the entanglement spectrum. The iMPS and fMPS slave-fermion point agrees well both with each other and the exact ES from DMRG.

The low-lying level of the entanglement spectrum cycles between a singlet and a triplet as we move the location of the entanglement cut within the chain. This is indicative of translation invariance breaking and the dimerized structure of the ground state: namely the low-level singlet is associated with a cut between dimers, whereas the low-level triplet is associated with a cut inside dimers.

We can also further understand the higher states in the entanglement spectra. A generic point in the dimer phase of the BLBQ model is SU(2)SU(2) symmetric. Consequently, the entanglement levels transform under SU(2)SU(2) representation and therefore we expect that degeneracies should go as the dimension of SU(2)SU(2) representations (i.e. 2n+12n+1 for non-negative integer nn); this can be seen in the multiplet structure of both the entanglement spectra in fig. 8(bottom-left). From 8(bottom-left) where the lowest three degeneracies between dimers form the singlet(S=0S=0), triplet(S=1S=1) and quintuplet(S=2S=2).

Refer to caption
Refer to caption
Refer to caption
Figure 8: Entanglement spectra of the projected mean-field state from iMPS and fMPS at (J,K)=(1,2)(J,K)=(1,-2) showing (a) the comparison with DMRG between dimers; and the fMPS spin-resolved ES (b) between dimers and (c) within dimers. Note that the single and triple degeneracy seen in (b) and (c) are the expected low-level structure seen in a pure VBS state. The lowest three entanglement levels of (b) are representations of SU(2): singlet, triplet and quintet.

Beyond considering a generic point within the dimer phase, we now consider the exactly solvable point where (J,K)=(0,1)(J,K)=(0,-1) (the so-called KBB point klumper1989new , barber1989spectrum ), which is invariant under a larger symmetry group, SU(3)SU(3) (as opposed to SU(2)SU(2)). This larger symmetry group forces the triplet and quintet to form an octet (the adjoint representation of SU(3)SU(3)). Variationally, the vanishing of the hopping parameter in the slave-fermion mean-field hamiltonian forces this larger symmetry group at the level of the variational Gutzwiller projected wavefunction. See fig. 9.

Refer to caption
Figure 9: Spin resolved entanglement spectrum (between dimers) for the (J,K)=(0,1)(J,K)=(0,-1) variational point generated from fMPS. The SU(3)SU(3) symmetry forces the S=1S=1 and S=2S=2 into a degenerate octet.

The points (J,K)=(1,2)(J,K)=(-1,-2) and (J,K)=(1,3)(J,K)=(-1,-3) which are found at a variational minima with t=0t=0 in the parent Hamiltonian by ref. liu2012gutzwiller, also have SU(3)SU(3) symmetry. This symmetry is not present in the true DMRG ground state which transforms only under SU(2)SU(2) symmetry; therefore, a better agreement is obtained by perturbing slightly away from this point (see  S7).

IV.4.2 Haldane phase

In this section we compute the entanglement spectra of the AKLT and Heisenberg points belonging to the Haldane phase of BLBQ. The slave-fermion mean-field for this model has a 4-fold degeneracy at the fermi level that allows for choosing six orthogonal pre-projected mean-field states (at half filling). Projecting each of these states, generates (post-projection) a space of MPS which span 4 degenerate ground states which correspond to the representations of the sum of the two fractionalized S=1/2S=1/2 edge modes of the Haldane phase.

Here we start by considering the AKLT point which has an exact analytic solution. The AKLT point (J,K)=(1,1/3)(J,K)=(1,1/3) is exactly mapped under the above slave fermion projective construction to (χ,δ,μ)=(1,3/2,0)(\chi,\delta,\mu)=(1,3/2,0). To find the AKLT state which (for example) has the edge modes \uparrow\downarrow we can either search in the four-fold projected degenerate space or choose the correct orbitals at the fermi-level pre-projection. We find the entanglement spectra for each of the four fMPS which correspond to \downarrow\downarrow, \downarrow\uparrow, \uparrow\downarrow, \uparrow\uparrow edge spin configurations is equal to ln(2)\ln(2).

For the iMPS, unlike the dimer phase, where there was 22-fold degeneracy in the leading eigenvalues of the transfer matrix operator for points in the Haldane-phase, we find 44-fold degeneracy. We obtain 22-negative and 22-positive (equal in magnitude) leading eigenvalues. Taking the space spanned by the two eigenvectors with positive eigenvalues, we apply the iMPS orthogonalization procedure from sec. IV.2. From this process, for the AKLT slave-fermion point we find after orthonormalization the iMPS

A0=(1001)A=2(0eiθ00)A=2(00eiθ0)\begin{split}A^{0}&=\begin{pmatrix}-1&0\\ 0&1\end{pmatrix}\\ A^{\downarrow}&=\sqrt{2}\begin{pmatrix}0&-e^{-i\theta}\\ 0&0\end{pmatrix}\\ A^{\uparrow}&=\sqrt{2}\begin{pmatrix}0&0\\ e^{i\theta}&0\end{pmatrix}\end{split} (48)

associated with two pure states (i.e |\ket{\uparrow\downarrow} and |\ket{\downarrow\uparrow} of the edge modes). In supplement  S8, we also obtain the iMPS representation of the S=1/2S=1/2 VBS groundstate of the Majumdar–Ghosh (MG) chain.

In fig. 10 we also present the entanglement spectrum of the Heisenberg point (J,K)=(1,0)(J,K)=(1,0) obtained using both fMPS and iMPS. We see that the lower levels match well the entanglement spectrum levels obtained from DMRG of the true Heisenberg ground state. Discrepancies occur naturally higher-up in the spectrum as the variational wavefunction is not the exact groundstate. However, the entanglement spectrum of the variational groundstate (qualitatively) captures the symmetries and degeneracies of the true entanglement spectrum. Note the fact that every level has even degeneracy comes from the topological nature of the phase. Moreover, notice that the lowest part of the entanglement spectra match the AKLT state in the same sector.

Refer to caption
Refer to caption
Figure 10: (top) Comparison of entanglement spectrum obtained from fMPS, iMPS and DMRG at the (J,K)=(1,0)(J,K)=(1,0) Heisenberg point. (bottom) Spin resolved entanglement spectrum from fMPS (J,K)=(1,0)(J,K)=(1,0) Heisenberg point in the (S=1,Sz=1)(S=1,S_{z}=1) sector.

IV.4.3 Critical points

We compute the two critical points at (J,K)=(1,1)(J,K)=(1,-1) (the Takhtajan-Babujian(TB) point takhtajan1982picture ,babujian1982exact ) and at (J,K)=(1,1)(J,K)=(1,1)(the Uimin-Lai-Sutherland (ULS) uimin1970one ,lai1974lattice , sutherland1994model ); they are gapless and hence we analyze them only in the framework of our fMPS method. The TB ground state is unique; the associated effective conformal field theory is SU(2)|k=2SU(2)|_{k=2}.

The ULS ground state is also unique. However, it has an enlarged SU(3)SU(3) symmetry group; the associated effective conformal field theory is SU(3)|k=1SU(3)|_{k=1}. In particular it can be mapped to the SU(3)SU(3) nearest-neighbor Heisenberg model thomale2015entanglement . This enforces an equal number of “quark” particle constraints (and hence global equal number of spins 1,1,01,-1,0). At the mean-field level, the pairing parameter vanishes since JK=0J-K=0. The Hamiltonian is then a tensor sum of 3 identical hopping Hamiltonians acting independently on the fermions of flavour “up”, “down” and “zero”. The particle number constraint of c1,c1,c0c_{1},c_{-1},c_{0} is naturally enforced at the mean-field level if the the number of sites NN is a multiple of 33.

The SU(3)SU(3) symmetry of the ULS point is reflected in the degeneracies of the entanglement spectrum where the S=1S=1 and S=2S=2 levels combine together to form SU(3)SU(3) octets as can be seen in fig. 11. For the TB point the S=1S=1 and S=2S=2 entanglement levels remain separated.

The central charge of SU(N)|kSU(N)|_{k} CFTs is given by: c=k(N21)/(N+k)c=k\left(N^{2}-1\right)/(N+k). Hence, analytically c=1.5c=1.5 for the TB point and c=2c=2 for ULS.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 11: (a) Comparison of entanglement spectra at ULS point for fMPS and DMRG. (b) Comparison of entanglement spectra at TB point for fMPS and DMRG. (c) Spin resolved entanglement spectrum for (left) TB point and (right) ULS point

Calabrese and Cardy calabrese2004j obtained the following expression for the entanglement entropy scaling for a 1D critical gapless point of finite size LL with open-boundary conditions and partition size xx:

S(x,L)=c6ln2Lπsin(πxL)+lng+s1/2S(x,L)=\frac{c}{6}\ln{\frac{2L}{\pi}\sin{(\frac{\pi x}{L})}}+\ln{g}+s_{1}/2 (49)

where lng\ln{g} is a boundary entropy term and S(x,L)S(x,L) is the von-Neumann entanglement entropy.

It was found in ref. laflorencie2006boundary, that there is an additional alternating term in S(x,L)S(x,L) which decays away from the boundaries. In fig. 12 we plot the entanglement entropy against c6ln2Lπsin(πxL)\frac{c}{6}\ln{\frac{2L}{\pi}\sin{(\frac{\pi x}{L})}} for both TB and ULS models. We work on system of size L=300L=300 and plot the region x[L/4,3L/8]x\in[L/4,3L/8] (so [75,111][75,111] for TB and [76,112][76,112] for ULS). We picked the lower bound at L/4L/4 as it is far enough from the boundary and the upper boundary at 3L/83L/8 to work in the region of the sine curve with xx away from L/2L/2 where the curve becomes very flat. For the SU(3)SU(3) ULS point, when xx is a multiple of 33, the highest eigenvalue Schmidt vector contains equal numbers of “quarks” and hence is dominant. For cuts at x=3k+1,3k+2x=3k+1,3k+2 (for integer kk), the Schmidt vectors cannot satisfy the particle conservation constraint and hence the highest eigenvalue Schmidt vectors are degenerate. A similar situation occurs at the SU(2)SU(2) TB point where the 22-periodicity is easily explained in the dimer picture: for even cuts we cut between dimers, whereas for odd cuts we break dimers and hence split the singlet apart. The alternating term is still significant for the parameters we chose. Hence, it is difficult to reliably extract the central charge. We overlay the lines obtained from least square fitting for both models. For the TBTB point we obtain the central charge c1.505c\approx 1.505 which is remarkably close to its true value of 1.51.5. For the ULSULS point we obtain c=1.307c=1.307 which is substantially away from the true value at the ULSULS point of c=2c=2.

Refer to caption
Refer to caption
Figure 12: Von-Neumann entanglement entropy scaling of the slave-fermion wave-functions describing the ULS (top) and TB (bottom) critical points for a system of size L=300L=300 computed with fMPS. Shown also is the central charge obtained from least squares fitting.

V Discussion and Future Work

We have developed a series of efficient and highly parallel algorithms to obtain the finite and infinite (for gapped states) MPS representation of fermionic mean-field states. Gutzwiller projection is easily implemented by eliminating the doubly-occupied and unoccupied physical sectors of the mean-field slave-fermion MPS tensors. We have used these methods to obtain the (i)MPS representation of Gutzwiller projected mean-field states that arise from the variational slave-fermion approach to the S=1S=1 Bilinear-Biquadratic (BLBQ) quantum spin chain introduced in ref. liu2012gutzwiller, . We first verify that the energies we obtain via both finite MPS and infinite MPS (not applicable to the critical points) for the points considered are within the error bars of their VMC calculations liu2012gutzwiller .

Additionally, we obtain the entanglement spectra at two critical points (ULS and TB) and several generic points in the dimer and Haldane phases of the BLBQ model. We find good qualitative (and quantitative) agreement with results obtained directly from DMRG. We briefly discuss the salient structural features of the entanglement spectrum in all the phases (but see ref. thomale2015entanglement, for a more detailed analysis). Extracting the central charges of the conformal field theories describing the two gapless critical points from numerical computation of the entanglement spectrum on finite open-boundary systems is made difficult by a slowly decaying oscillatory term in the entanglement entropy. However, we do obtain very good agreement for the central charge at the TB point, 1.505 as compared to the exact analytical value of 1.51.5. At the ULS point, we obtain a larger discrepancy, c=1.331.33 compared to the analytical value of 22.

We also introduce an algorithmic procedure that orthogonalizes an iMPS by breaking it down into its pure states. This is essential when dealing with degenerate ground states that appear upon Gutzwiller projection as is the case with points in the dimer phase of the BLBQ model. Having obtained the pure states, we can compute the entanglement spectrum for any state in the groundstate manifold. We check that the entanglement spectrum obtained from iMPS matches the one we obtain from the finite MPS procedure. Discrepancies naturally appear as we approach values close to the thresholds used to generate the finite MPS and infinite MPS.

The methods can be easily adapted to the study of systems on 2D ladders (infinite in length but with finite width). The iMPS unit cell is now formed by the tensors sitting on the width of the cylinder. We will explore the applications of Gutzwiller projected variational wavefunctions to the study of 2D quantum spin liquids in future publications. This method may also be applicable to topological states such as quantum Hall and fractional Chern insulators that are represented as products of mean-field wavefunctions.

Acknowledgements.
BKC acknowledges support from the Department of Energy grant DOE de-sc0020165. This project is part of the Blue Waters sustained petascale computing project, which is supported by the National Science Foundation (awards OCI-0725070 and ACI-1238993) and the State of Illinois. Blue Waters is a joint effort of the University of Illinois at Urbana-Champaign and its National Center for Supercomputing Applications. GKC was supported by the US National Science Foundation via grant no. 1839204. GKC also acknowledges support from the Simons Foundation via the Investigator Award and the Many-Electron Collaboration.

References

  • (1) A. J. Leggett et al., Quantum liquids: Bose condensation and Cooper pairing in condensed-matter systems.   Oxford university press, 2006.
  • (2) N. Read and D. Green, “Paired states of fermions in two dimensions with breaking of parity and time-reversal symmetries and the fractional quantum hall effect,” Physical Review B, vol. 61, no. 15, p. 10267, 2000.
  • (3) M. T. Fishman and S. R. White, “Compression of correlation matrices and an efficient method for forming matrix product states of fermionic gaussian states,” Physical Review B, vol. 92, no. 7, p. 075132, 2015.
  • (4) Y.-H. Wu, L. Wang, and H.-H. Tu, “Tensor network representations of parton wave functions,” Physical Review Letters, vol. 124, no. 24, p. 246401, 2020.
  • (5) N. Schuch and B. Bauer, “Matrix product state algorithms for gaussian fermionic states,” Physical Review B, vol. 100, no. 24, p. 245121, 2019.
  • (6) U. Schollwöck, “The density-matrix renormalization group in the age of matrix product states,” Annals of Physics, vol. 326, no. 1, pp. 96–192, 2011.
  • (7) S.-A. Cheong and C. L. Henley, “Many-body density matrices for free fermions,” Physical Review B, vol. 69, no. 7, p. 075111, 2004.
  • (8) I. Klich, “Lower entropy bounds and particle number fluctuations in a fermi sea,” Journal of Physics A: Mathematical and General, vol. 39, no. 4, p. L85, 2006.
  • (9) G. Knizia and G. K.-L. Chan, “Density matrix embedding: A simple alternative to dynamical mean-field theory,” Physical review letters, vol. 109, no. 18, p. 186404, 2012.
  • (10) I. Peschel, “Special review: Entanglement in solvable many-particle models,” Brazilian Journal of Physics, vol. 42, no. 3-4, pp. 267–291, 2012.
  • (11) D. A. Harville, “Matrix algebra from a statistician’s perspective,” 1998.
  • (12) W. Su, J. Schrieffer, and A. J. Heeger, “Solitons in polyacetylene,” Physical review letters, vol. 42, no. 25, p. 1698, 1979.
  • (13) S. Ripka, J. Blaizot, and G. Ripka, Quantum Theory of Finite Systems.   MIT Press, 1986. [Online]. Available: https://books.google.com/books?id=s_xlQgAACAAJ
  • (14) A. Y. Kitaev, “Unpaired majorana fermions in quantum wires,” Physics-Uspekhi, vol. 44, no. 10S, p. 131, 2001.
  • (15) Z.-X. Liu, Y. Zhou, H.-H. Tu, X.-G. Wen, and T.-K. Ng, “Gutzwiller projected wave functions in the fermionic theory of s= 1 spin chains,” Physical Review B, vol. 85, no. 19, p. 195144, 2012.
  • (16) F. D. M. Haldane, “Continuum dynamics of the 1-d heisenberg antiferromagnet: Identification with the o (3) nonlinear sigma model,” Physics Letters A, vol. 93, no. 9, pp. 464–468, 1983.
  • (17) I. Affleck, T. Kennedy, E. H. Lieb, and H. Tasaki, “Valence bond ground states in isotropic quantum antiferromagnets,” in Condensed matter physics and exactly soluble models.   Springer, 1988, pp. 253–304.
  • (18) Z.-C. Gu and X.-G. Wen, “Tensor-entanglement-filtering renormalization approach and symmetry-protected topological order,” Physical Review B, vol. 80, no. 15, p. 155131, 2009.
  • (19) F. Pollmann, E. Berg, A. M. Turner, and M. Oshikawa, “Symmetry protection of topological phases in one-dimensional quantum spin systems,” Physical review b, vol. 85, no. 7, p. 075125, 2012.
  • (20) A. M. Turner, F. Pollmann, and E. Berg, “Topological phases of one-dimensional fermions: An entanglement point of view,” Physical review b, vol. 83, no. 7, p. 075102, 2011.
  • (21) F. Pollmann, A. M. Turner, E. Berg, and M. Oshikawa, “Entanglement spectrum of a topological phase in one dimension,” Physical review b, vol. 81, no. 6, p. 064439, 2010.
  • (22) H. Li and F. D. M. Haldane, “Entanglement spectrum as a generalization of entanglement entropy: Identification of topological order in non-abelian fractional quantum hall effect states,” Physical review letters, vol. 101, no. 1, p. 010504, 2008.
  • (23) M. Levin and X.-G. Wen, “Detecting topological order in a ground state wave function,” Physical review letters, vol. 96, no. 11, p. 110405, 2006.
  • (24) A. Kitaev and J. Preskill, “Topological entanglement entropy,” Physical review letters, vol. 96, no. 11, p. 110404, 2006.
  • (25) Z.-X. Liu, Y. Zhou, and T.-K. Ng, “Fermionic theory for quantum antiferromagnets with spin s¿ 1 2,” Physical Review B, vol. 82, no. 14, p. 144422, 2010.
  • (26) ——, “Gutzwiller approach for elementary excitations in s= 1 antiferromagnetic chains,” New Journal of Physics, vol. 16, no. 8, p. 083031, 2014.
  • (27) R. Orus and G. Vidal, “Infinite time-evolving block decimation algorithm beyond unitary evolution,” Physical Review B, vol. 78, no. 15, p. 155117, 2008.
  • (28) D. Perez-Garcia, F. Verstraete, M. M. Wolf, and J. I. Cirac, “Matrix product state representations,” arXiv preprint quant-ph/0608197, 2006.
  • (29) S. Ruiz et al., “Tensor networks in condensed matter,” Ph.D. dissertation, Technische Universität München, 2011.
  • (30) A. Klümper, “New results for q-state vertex models and the pure biquadratic spin-1 hamiltonian,” EPL (Europhysics Letters), vol. 9, no. 8, p. 815, 1989.
  • (31) M. N. Barber and M. T. Batchelor, “Spectrum of the biquadratic spin-1 antiferromagnetic chain,” Physical Review B, vol. 40, no. 7, p. 4621, 1989.
  • (32) L. Takhtajan, “The picture of low-lying excitations in the isotropic heisenberg chain of arbitrary spins,” Physics Letters A, vol. 87, no. 9, pp. 479–482, 1982.
  • (33) H. Babujian, “Exact solution of the one-dimensional isotropic heisenberg chain with arbitrary spins s,” Physics Letters A, vol. 90, no. 9, pp. 479–482, 1982.
  • (34) G. Uimin, “One-dimensional problem for s= 1 with modified antiferromagnetic hamiltonian,” JETPL, vol. 12, p. 225, 1970.
  • (35) C. Lai, “Lattice gas with nearest-neighbor interaction in one dimension with arbitrary statistics,” Journal of Mathematical Physics, vol. 15, no. 10, pp. 1675–1676, 1974.
  • (36) B. Sutherland, “Model for a multicomponent quantum system,” in Exactly Solvable Models Of Strongly Correlated Electrons.   World Scientific, 1994, pp. 287–297.
  • (37) R. Thomale, S. Rachel, B. A. Bernevig, and D. P. Arovas, “Entanglement analysis of isotropic spin-1 chains,” Journal of Statistical Mechanics: Theory and Experiment, vol. 2015, no. 7, p. P07017, 2015.
  • (38) P. Calabrese and J. Cardy, “J. stat. mech. p06002,” 2004.
  • (39) N. Laflorencie, E. S. Sørensen, M.-S. Chang, and I. Affleck, “Boundary effects in the critical scaling of entanglement entropy in 1d systems,” Physical review letters, vol. 96, no. 10, p. 100603, 2006.
  • (40) V. Karimipour and L. Memarzadeh, “Matrix product representations for all valence bond states,” Physical Review B, vol. 77, no. 9, p. 094416, 2008.

Supplementary Material

Appendix S1 Finite MPS boundary tensors

We can produce the left/right boundary tensors in either left/right canonical form. Naturally, most of the time we will be concerned with producing the left boundary tensor in left canonical form and the right boundary tensor in right canonical form. Employing the notation used in the main text:

A1,a1[1]σ1=σ1|La11,NA^{[1]\sigma_{1}}_{1,a_{1}}=\braket{\sigma_{1}}{L^{1,N}_{a_{1}}} (S1)
AaN1,1[N]σN=σN1|RaN1N1,NA^{[N]\sigma_{N}}_{a_{N-1},1}=\braket{\sigma_{N-1}}{R^{N-1,N}_{a_{N-1}}} (S2)

Similar expressions exist for the left boundary tensor in right-canonical form and the right boundary tensor in left canonical form respectively:

A1,a1[1]σ1=σ1|Λa11,N1La11,NA^{[1]\sigma_{1}}_{1,a_{1}}=\braket{\sigma_{1}}{\Lambda^{1,N-1}_{a_{1}}L^{1,N}_{a_{1}}} (S3)
AaN1,1[N]σN=σN1|ΛaN1N1,NRaN1N1,NA^{[N]\sigma_{N}}_{a_{N-1},1}=\braket{\sigma_{N-1}}{\Lambda^{N-1,N}_{a_{N-1}}R^{N-1,N}_{a_{N-1}}} (S4)

To produce a finite MPS in mixed-canonical form at site ll:

A1,a1[1]σ1=σ1|La11,NAαkαk+1[i+1]σi+1=σi+1|Lαk+1i;N|Lαki+1;Nif1<i+1lAαkαk+1[i+1]σi+1=σi+1Rαk+1i+1;N|Rαki;Nifl<i+1<NAaN1,1[N]σN=σN1|RaN1N1,N\begin{split}A^{[1]\sigma_{1}}_{1,a_{1}}&=\braket{\sigma_{1}}{L^{1,N}_{a_{1}}}\\ A^{[i+1]\sigma_{i+1}}_{\alpha_{k}\alpha_{k+1}}&=\braket{\sigma_{i+1}}{\otimes L^{i;N}_{\alpha_{k+1}}}{L^{i+1;N}_{\alpha_{k}}}\quad\text{if}\qquad 1<i+1\leq l\\ A^{[i+1]\sigma_{i+1}}_{\alpha_{k}\alpha_{k+1}}&=\braket{\sigma_{i+1}\otimes R^{i+1;N}_{\alpha_{k+1}}}{R^{i;N}_{\alpha_{k}}}\quad\text{if}\qquad l<i+1<N\\ A^{[N]\sigma_{N}}_{a_{N-1},1}&=\braket{\sigma_{N-1}}{R^{N-1,N}_{a_{N-1}}}\end{split} (S5)

Appendix S2 iMPS Multi-site unit cell

We can also consider iMPS with multiple sites unit cells - i.e.

|ΨiMPS=σ(ABCD)(ABCD)\ket{\Psi_{iMPS}}=\sum_{\sigma}\ldots\left(ABCD\right)\left(ABCD\right)\ldots (S6)

Our method will generate the uniform unit-cell tensor:

TσN+1σN+2σN+3σN+4=AσN+1BσN+2CσN+3DσN+4T^{\sigma_{N+1}\sigma_{N+2}\sigma_{N+3}\sigma_{N+4}}=A^{\sigma_{N+1}}B^{\sigma_{N+2}}C^{\sigma_{N+3}}D^{\sigma_{N+4}} (S7)
TαβσN+1σN+2σN+3σN+4=γCαγ[2N+4](σN+1σN+2σN+3σN+4|RβN;2N|)|RγN,2N+4T^{\sigma_{N+1}\sigma_{N+2}\sigma_{N+3}\sigma_{N+4}}_{\alpha\beta}=\sum_{\gamma}C^{[2N+4]}_{\alpha\gamma}\left(\bra{\sigma_{N+1}\sigma_{N+2}\sigma_{N+3}\sigma_{N+4}}\otimes\bra{R^{N;2N}_{\beta}}\right)\ket{R^{N,2N+4}_{\gamma}} (S8)

We can obtain a more compact representation of TT by decimating the unit-cell tensor (i.e. inserting complete basis |RN+1,2N+4|R^{N+1,2N+4}\rangle,|RN+2,2N+4|R^{N+2,2N+4}\rangle, |RN+1,2N+3|R^{N+1,2N+3}\rangle inside the inner product) and expressing it in terms of the following 4 on-site tensors:

A~α,j1σN+1=γCαγ[2N+4](σN+1|Rj1N+1;2N+4|)|RN,2N+4γ\tilde{A}^{\sigma_{N+1}}_{\alpha,j_{1}}=\sum_{\gamma}C^{[2N+4]}_{\alpha\gamma}\left(\bra{\sigma_{N+1}}\otimes\bra{R^{N+1;2N+4}_{j_{1}}}\right)\ket{R^{N,2N+4}}_{\gamma} (S9)
B~j1j2σN+2=(σN+2|Rj2N+2;2N+4|)|Rj1N+1;2N+4\tilde{B}^{\sigma_{N+2}}_{j_{1}j_{2}}=\left(\bra{\sigma_{N+2}}\otimes\bra{R^{N+2;2N+4}_{j_{2}}}\right)\ket{R^{N+1;2N+4}_{j_{1}}} (S10)
C~j2j3σN+2=(σN+3||Rj3N+3;2N+4|)|Rj2N+2;2N+4\tilde{C}^{\sigma_{N+2}}_{j_{2}j_{3}}=\left(\bra{\sigma_{N+3}}|\otimes\bra{R^{N+3;2N+4}_{j_{3}}}\right)\ket{R^{N+2;2N+4}_{j_{2}}} (S11)
D~j3βσN+2=(σN+4|RβN;2N|)|Rj3N+3;2N+4\tilde{D}^{\sigma_{N+2}}_{j_{3}\beta}=\left(\bra{\sigma_{N+4}}\otimes\bra{R^{N;2N}_{\beta}}\right)\ket{R^{N+3;2N+4}_{j_{3}}} (S12)

So that:

TσN+1σN+2σN+3σN+4=A~σN+1B~σN+1C~σN+2D~σN+3T^{\sigma_{N+1}\sigma_{N+2}\sigma_{N+3}\sigma_{N+4}}=\tilde{A}^{\sigma_{N+1}}\tilde{B}^{\sigma_{N+1}}\tilde{C}^{\sigma_{N+2}}\tilde{D}^{\sigma_{N+3}} (S13)

Note that obtained in this way the tensors A~\tilde{A},B~\tilde{B},C~\tilde{C} and D~\tilde{D} are not necessarily uniform as they depend on the gauge choice of the Schmidt decompositions that generate |RN+1,2N+4|R^{N+1,2N+4}\rangle,|RN+2,2N+4|R^{N+2,2N+4}\rangle, |RN+1,2N+3|R^{N+1,2N+3}\rangle.

Appendix S3 Schmidt decomposition of Slater Determinants and MPS tensor computation

Consider an NN-particle Slater determinant |Ψ|\Psi\rangle with support on an (indexed) MM-site system. For a bipartition [L]x[R]=[1,i]x[i+1,M][L]x[R]=[1,\ldots i]x[i+1,\ldots M], where ii is any site index 1iM11\leq i\leq M-1, |Ψ|\Psi\rangle can be easily and efficiently brought to the following form peschel2012special :

|Ψ=Πk=1N[ϵkϕk;L+1ϵkϕk;R]|\Psi\rangle=\Pi^{N}_{k=1}\left[\sqrt{\epsilon_{k}}\phi^{\dagger}_{k;L}+\sqrt{1-\epsilon_{k}}\phi^{\dagger}_{k;R}\right] (S14)

Here, 0ϵk10\leq\epsilon_{k}\leq 1; ϕk;L\phi^{\dagger}_{k;L} and ϕk;R\phi^{\dagger}_{k;R} are operators that create orthogonal single orbital states with support in the left, respectively right, bipartitions.

The Schmidt decomposition of |Ψ|\Psi\rangle is obtained by expanding the above product:

|Ψ={l},{r}λ{l1,l2,lp;r1,r2,rNp}|ϕl1;L,.ϕlp;L|ϕr1;R,.ϕrNp;R|\Psi\rangle=\sum_{\{l\},\{r\}}\lambda_{\{l_{1},l_{2},\ldots l_{p};r_{1},r_{2},\ldots r_{N-p}\}}|\phi_{{l_{1}};L},\ldots.\phi_{{l_{p}};L}\rangle|\phi_{{r_{1}};R},\ldots.\phi_{{r_{N-p}};R}\rangle (S15)

where

λ{l1,l2,lp;r1,r2,rNp}=ϵl1ϵlp1ϵr11ϵrp\lambda_{\{l_{1},l_{2},\ldots l_{p};r_{1},r_{2},\ldots r_{N-p}\}}=\sqrt{\epsilon_{l_{1}}}\cdots\sqrt{\epsilon_{l_{p}}}\cdots\sqrt{1-\epsilon_{r_{1}}}\cdots\sqrt{1-\epsilon_{r_{p}}}

are the Schmidt values and:

|L{l}=|ϕl1;L,.ϕlp;L|L_{\{l\}}\rangle=|\phi_{{l_{1}};L},\ldots.\phi_{{l_{p}};L}\rangle
|R{r}=|ϕr1;R,.ϕrNp;R|R_{\{r\}}\rangle=|\phi_{{r_{1}};R},\ldots.\phi_{{r_{N-p}};R}\rangle

are the left/right Schmidt vectors. These are Slater determinants made up of subsets of the NN-orbital sets {ϕ;L}\{\phi_{;L}\} ({ϕ;R}\{\phi_{;R}\}).

The exact Schmidt decomposition is given by an exponential number, 2N2^{N} of tuples (λ,|L,|R)\left(\lambda,|L\rangle,|R\rangle\right). In practice, we will select only the tuples that have λ\lambda greater than a threshold value. Due to the exponentially decaying nature of the Schmidt values, this will actually give us a very good approximation of the Slater determinant.

S3.1 Overlaps of Schmidt vectors

Consider two right Schmidt vectors generated by pp-orbital subsets {r1}\{r^{1}\} and {r2}\{r^{2}\}. Their overlap is given by:

Rr1|Rr2=ϕr11;R,.ϕrp2;R|ϕr12;R,.ϕrp2;R\langle R_{r^{1}}|R_{r^{2}}\rangle=\langle\phi_{{r^{1}_{1}};R},\ldots.\phi_{{r^{2}_{p}};R}|\phi_{{r^{2}_{1}};R},\ldots.\phi_{{r^{2}_{p}};R}\rangle (S16)

So

Rr1|Rr2=detM{r1}{r2}\langle R_{r^{1}}|R_{r^{2}}\rangle=\det{M_{\{r^{1}\}\{r^{2}\}}} (S17)
M{r1}{r2}=[ϕr11;R|ϕr12;Rϕr11;R|ϕr22;Rϕr11;R|ϕrp2;Rϕr21;R|ϕr12;Rϕr21;R|ϕr22;Rϕr21;R|ϕrp2;Rϕrp1;R|ϕr12;Rϕrp1;R|ϕr22;Rϕrp1;R|ϕrp2;R]M_{\{r^{1}\}\{r^{2}\}}=\begin{bmatrix}\braket{\phi_{{r^{1}_{1}};R}}{\phi_{{r^{2}_{1}};R}}&\braket{\phi_{{r^{1}_{1}};R}}{\phi_{{r^{2}_{2}};R}}&\ldots&\braket{\phi_{{r^{1}_{1}};R}}{\phi_{{r^{2}_{p}};R}}\\ \braket{\phi_{{r^{1}_{2}};R}}{\phi_{{r^{2}_{1}};R}}&\braket{\phi_{{r^{1}_{2}};R}}{\phi_{{r^{2}_{2}};R}}&\ldots&\braket{\phi_{{r^{1}_{2}};R}}{\phi_{{r^{2}_{p}};R}}\\ \vdots&\vdots&\ddots&\vdots\\ \braket{\phi_{{r^{1}_{p}};R}}{\phi_{{r^{2}_{1}};R}}&\braket{\phi_{{r^{1}_{p}};R}}{\phi_{{r^{2}_{2}};R}}&\ldots&\braket{\phi_{{r^{1}_{p}};R}}{\phi_{{r^{2}_{p}};R}}\par\par\end{bmatrix} (S18)

Each of these matrices can be formed by selecting rows and columns from the matrix OO which needs to be computed only once.

O=[ϕ1;R|ϕ1;Rϕ1;R|ϕ2;Rϕ1;R|ϕN;Rϕ2;R|ϕ1;Rϕ2;R|ϕ2;Rϕ2;R|ϕN;RϕN;R|ϕ1;RϕN;R|ϕ2;RϕN;R|ϕN;R]O=\begin{bmatrix}\braket{\phi_{1;R}}{\phi_{1;R}}&\braket{\phi_{1;R}}{\phi_{2;R}}&\ldots&\braket{\phi_{1;R}}{\phi_{N;R}}\\ \braket{\phi_{2;R}}{\phi_{1;R}}&\braket{\phi_{2;R}}{\phi_{2;R}}&\ldots&\braket{\phi_{2;R}}{\phi_{N;R}}\\ \vdots&\vdots&\ddots&\vdots\\ \braket{\phi_{N;R}}{\phi_{1;R}}&\braket{\phi_{N;R}}{\phi_{2;R}}&\ldots&\braket{\phi_{N;R}}{\phi_{N;R}}\par\par\end{bmatrix} (S19)

Appendix S4 IMPS with edge modes

Obtaining the iMPS involves producing mean-field ground states on two systems of different sizes. If the mean-field Hamiltonian has multiple zero energy orbitals, we need to make sure we form the two ground states using the “same” zero energy orbitals. We next show how this works in the context of the SSH model.

HSSH=vn(cn,1cn,2+h.c.)+wn(cn+1,1cn,2+h.c.)H_{SSH}=v\sum_{n}\left(c^{\dagger}_{n,1}c_{n,2}+\text{h.c.}\right)+w\sum_{n}\left(c^{\dagger}_{n+1,1}c_{n,2}+\text{h.c.}\right)

In the topological phase v<wv<w, there will be two zero-energy modes (and hence 2-fold ground state degeneracy).

Refer to caption
Figure S1: Single particle eigenvalues for the SSH model (eqn. 11) for N=24N=24, v=0.5v=0.5, w=1w=1.
Refer to caption
Figure S2: Edge mode eigenvector in real-space for N=24N=24, v=0.5v=0.5, w=1w=1 for the SSH model (eqn. 11) .

Suppose the ground state produced in the first step of iMPS contains the zero-energy mode depicted in the figure. We picture the insertion step as introducing an unit cell in the middle of the chain. We generate the zero energy mode of the N+1=49 unit cell ground state by having the amplitudes on the first 24 unit cells and last 24 unit cells the same as the amplitudes of the zero energy mode on the N=48 unit cell system. In addition, the amplitudes on the 2 new sites introduced in the middle will be zero.

Appendix S5 Constant entanglement spectrum in the bulk for area-law entangled states

We obtain the entanglement spectrum for the Slater Determinant ground state of the SSH model of eqn. 11 on a finite OBC chain with N=24N=24 unit-cells, v=1v=1, w=0.99w=0.99 and Schmidt value threshold ϵ=1012\epsilon=10^{-12}.

Refer to caption
Refer to caption
Figure S3: Top: Largest 16 entanglement levels (some levels are degenerate) for the SSH groundstate obtained on N=24N=24 unit cells, v=1v=1, w=0.99w=0.99. We are considering here the even cuts, i.e. cuts between unit cells (and not inside unit cells). Similar results hold for the odd cuts, i.e entanglement cuts inside the unit cells. Bottom: We show here the largest 16 entanglement levels (some levels are degenerate) for entanglement cuts around the middle of the chain (i.e “bulk”). Note that the values have converged in bulk and hence our iMPS algorithm can be used.

Appendix S6 iMPS orthogonalization procedure

In this subsection we present an intuitive derivation of the iMPS orthogonalization procedure given in ref. orus2008infinite, .

Upon constructing the left/right and uniform bulk tensors for our area-law entangled wavefunction, and performing a local operation (i.e Gutzwiller projection), we end up with non-orthonormal MPS of the following form (contraction of virtual indices implied):

|Ψ={σ}L1σ1L2σ2LNσNAσN+1AσN+2AσN+2MRNσN+2M+1R2σ2N+2M1R1σ2N+M|σ1σ2N+2M\ket{\Psi}=\sum_{\{\sigma\}}L^{\sigma_{1}}_{1}L^{\sigma_{2}}_{2}\ldots L^{\sigma_{N}}_{N}A^{\sigma_{N+1}}A^{\sigma_{N+2}}\ldots A^{\sigma_{N+2M}}R^{\sigma_{N+2M+1}}_{N}\ldots R^{\sigma_{2N+2M-1}}_{2}R^{\sigma_{2N+M}}_{1}\ket{\sigma_{1}\ldots\sigma_{2N+2M}} (S20)

We are now interested in putting this MPS into a mixed-canonical form with a cut between sites N+MN+M and N+M+1N+M+1.

We rewrite the above MPS (and drop the physical indices for ease of presentation although they are implied) as:

|Ψ=(L1L2LNAAA)1,α(AAARNR2R1)α,1=|ΨLα|ΨRα\begin{split}\ket{\Psi}&=\sum\left(L_{1}L_{2}\ldots L_{N}AA\ldots A\right)_{1,\alpha}\left(AA\ldots AR_{N}\ldots R_{2}R_{1}\right)_{\alpha,1}\\ &=\sum\ket{\Psi_{L}}_{\alpha}\ket{\Psi_{R}}_{\alpha}\end{split} (S21)

with MM uniform AA tensors inserted after LNL_{N} and MM uniform AA tensors inserted before RNR_{N}; |ΨL(R)α\ket{\Psi_{L}(R)}_{\alpha} are non-orthonormal vectors on the left(right) system bipartitions.

To orthogonalize the (say) left vectors we obtain their overlap matrix MLM^{L} (which is hermitian). Upon diagonalization of MLM^{L} we are able to obtain an orthogonal set of vectors on the left bipartition. We proceed in the same way for the set of right vectors.

MαβL=βΨL|ΨLαM^{L}_{\alpha\beta}=_{\beta}\braket{\Psi_{L}}{\Psi_{L}}_{\alpha} (S22)

Since MLM_{L} is hermitian,

ML=ULDLULM_{L}=U_{L}D_{L}U^{\dagger}_{L} (S23)

It then follows that

1=DL1/2UL|ψL~|ψL~ULDL1/21=D_{L}^{-1/2}U^{\dagger}_{L}\tilde{\ket{\psi_{L}}}^{\dagger}\tilde{\ket{\psi_{L}}}U_{L}D_{L}^{-1/2} (S24)

with |ψL~\tilde{\ket{\psi_{L}}} a vector formed from |ψLα\ket{\psi_{L}}_{\alpha} so that

|ϕLγ=|ψL1,αULαγDL1/2\ket{\phi_{L}}_{\gamma}=\ket{\psi_{L}}_{1,\alpha}U_{L\alpha\gamma}D_{L}^{-1/2} (S25)

is now a set of orthonormal (Schmidt) vectors.

Similarly, for the right vectors we form the overlap matrix:

MαβR=βψR|ψRαM^{R}_{\alpha\beta}=_{\beta}\braket{\psi_{R}}{\psi_{R}}_{\alpha} (S26)
(MαβR)T=RRRNR1R1TRNTRTRT(M^{R}_{\alpha\beta})^{T}=\sum R^{*}R^{*}\ldots R^{*}_{N}\ldots R^{*}_{1}R^{T}_{1}\ldots R^{T}_{N}R^{T}\ldots R^{T} (S27)

Again MRM_{R} is hermitian and can be diagonalized:

MRT=URDRURM^{T}_{R}=U_{R}D_{R}U^{\dagger}_{R} (S28)
(MR)=UDRURT(M_{R})=U^{*}D_{R}U^{T}_{R} (S29)

and then

1=DR1URT(|ψR~)|ψR~URDR1/21=D_{R}^{-1}U^{T}_{R}\left(\tilde{\ket{\psi_{R}}}\right)^{\dagger}\tilde{\ket{\psi_{R}}}U^{*}_{R}D_{R}^{-1/2} (S30)

and so

|ϕRγ,1=(|ψRα,1URαγ(DR1/2)γγ\ket{\phi_{R}}_{\gamma,1}=(\ket{\psi_{R}}_{\alpha,1}U^{*}_{R\alpha\gamma}(D^{-1/2}_{R})_{\gamma\gamma} (S31)
|ϕRγ,1=DR1/2γγURγαT|ψRα,1\ket{\phi_{R}}_{\gamma,1}={D_{R}^{-1/2}}_{\gamma\gamma}{U^{*}_{R}}^{T}_{\gamma\alpha}\ket{\psi_{R}}_{\alpha,1} (S32)
|ϕRγ,1=DR1/2γγ(UR)γα|ψRα,1\ket{\phi_{R}}_{\gamma,1}={D_{R}^{-1/2}}_{\gamma\gamma}(U^{\dagger}_{R})_{\gamma\alpha}\ket{\psi_{R}}_{\alpha,1} (S33)

Going back to our original MPS representation and introducing the following resolutions of identity

I=ULDL1/2DL1/2ULI=U_{L}D_{L}^{-1/2}D^{1/2}_{L}U^{\dagger}_{L} (S34)

and

I=UDR1/2DR1/2UI=UD_{R}^{1/2}D^{-1/2}_{R}U^{\dagger} (S35)

we have:

|Ψ=|ΨLα|ΨRα=(L1L2LNAAA)1,α(AAARNR2R1)α,1=(L1L2LNAAA)1,α×(ULDL1/2DL1/2UL)α,α×(UDR1/2DR1/2U)α,α×(AAARNR2R1)α,1\begin{split}\ket{\Psi}&=\sum\ket{\Psi_{L}}_{\alpha}\ket{\Psi_{R}}_{\alpha}\\ &=\sum\left(L_{1}L_{2}\ldots L_{N}AA\ldots A\right)_{1,\alpha}\left(AA\ldots AR_{N}\ldots R_{2}R_{1}\right)_{\alpha,1}\\ &=\sum\left(L_{1}L_{2}\ldots L_{N}AA\ldots A\right)_{1,\alpha}\\ &\times\left(U_{L}D_{L}^{-1/2}D^{1/2}_{L}U^{\dagger}_{L}\right)_{\alpha,\alpha}\\ &\times\left(UD_{R}^{1/2}D^{-1/2}_{R}U^{\dagger}\right)_{\alpha,\alpha}\\ &\times\left(AA\ldots AR_{N}\ldots R_{2}R_{1}\right)_{\alpha,1}\end{split} (S36)

We isolate the orthogonal vectors by collecting the terms:

|ψ=(|ψαULαγDL1/2γγ)×(DL1/2ULURDR1/2)γδ×(DR1/2δδURδα|ψRα)\begin{split}\ket{\psi}&=\sum\left(\ket{\psi}_{\alpha}{U_{L}}_{\alpha\gamma}{D_{L}^{-1/2}}_{\gamma\gamma}\right)\\ &\times\left(D^{1/2}_{L}U^{\dagger}_{L}U_{R}D_{R}^{1/2}\right)_{\gamma\delta}\\ &\times\left({D_{R}^{-1/2}}_{\delta\delta}{U_{R}^{\dagger}}_{\delta\alpha}\ket{\psi_{R}}_{\alpha}\right)\end{split} (S37)

We thus obtain:

|ψ=|ϕL1,γ(DL1/2ULURDR1/2)γδ|ϕRδ,1\ket{\psi}=\ket{\phi_{L}}_{1,\gamma}\left(D^{1/2}_{L}U^{\dagger}_{L}U_{R}D_{R}^{1/2}\right)_{\gamma\delta}\ket{\phi_{R}}_{\delta,1} (S38)

By writing: ML=YYM_{L}=Y^{\dagger}Y and (MR)T=XX(M_{R})^{T}=XX^{\dagger} then YX=(DL1/2ULURDR1/2)YX=\left(D^{1/2}_{L}U^{\dagger}_{L}U_{R}D_{R}^{1/2}\right)

and obtain the SVD of YX=USVYX=USV^{\dagger} to finally obtain the mixed canonical representation of the MPS:

|ψ=(|ϕLU)S(V|ϕR)\ket{\psi}=\left(\ket{\phi_{L}}U\right)S\left(V^{\dagger}\ket{\phi_{R}}\right) (S39)

The presence of the uniform tensor in the bulk significantly simplifies obtaining MLM_{L} and MRM_{R} in the MNM\gg N limit, since the only terms that survives in the product of tensors solve the fixed point relations:

AMLA=ηMLA^{\dagger}M_{L}A=\eta M_{L} (S40)

where |η||\eta| is the largest such value(s). Similarly for MRM_{R}

AMRAT=ηMRA^{*}M_{R}A^{T}=\eta M_{R} (S41)

When viewed as vectors MLM_{L} and MRM_{R} solve the following eigenvalue equations

(AA)MRT=ηMRT\left(\sum A\bigotimes A^{*}\right)M^{T}_{R}=\eta M^{T}_{R} (S42)
(ATA)ML=ηML\left(\sum A^{T}\bigotimes A^{\dagger}\right)M_{L}=\eta M_{L} (S43)

which are the equations for finding the right(left) eigenvectors of the transfer matrix ε=(AA)\varepsilon=\left(\sum A\bigotimes A^{*}\right), hence reproducing the imps orthogonalization method used in ref. orus2008infinite, .

Appendix S7 Other entanglement spectra figures for BLBQ dimer phase points

S7.1 (J,K)=(1,2)(J,K)=(1,-2) entanglement spectrum comparison for odd cut

Refer to caption
Figure S4: Entanglement spectra comparison (inside dimers cut) for (J,K)=(1,2)(J,K)=(1,-2) of the BLQB model (eqn. 35) variational point between fMPS, iMPS and DMRG.

S7.2 (J,K)=(1,2)(J,K)=(-1,-2) point

Refer to caption
Refer to caption
Figure S5: Comparison of entanglement spectra for the point (J,K)=(1,2)(J,K)=(-1,-2) of the BLBQ model (eqn. 35) between fMPS and DMRG. Top: Hopping parameter χ=0\chi=0 forces SU(3)SU(3) degeneracies in the entanglement spectrum. Bottom: Taking χ=0.1\chi=0.1, we see that the degeneracies of the entanglement spectrum are now broken to SU(2)SU(2).
Refer to caption
Refer to caption
Figure S6: Spin resolved entanglement spectrum (between dimers cut) for the (J,K)=(1,2)(J,K)=(-1,-2) of the BLBQ model (eqn. 35) variational point obtained from fMPS. Top: hopping parameter χ=0\chi=0. Bottom: hopping parameter χ=0.1\chi=0.1

Appendix S8 iMPS tensors for spin S=1/2S=1/2 MG chain

The Majumdar-Ghosh (MG) model is a simple exactly solvable 1D quantum spin chain. The MG Hamiltonian is:

HMG=JnSnSn+1+J2nSnSn+2H_{MG}=J\sum_{n}S_{n}S_{n+1}+\frac{J}{2}\sum_{n}S_{n}S_{n+2} (S44)

In the thermodynamic limit the groundstate is 2-fold degenerate. Two pure orthogonal states in the groundstate manifold are given by products of dimer singlets covering neighboring sites. In one state the dimers cover sites (2n1,2n)(2n-1,2n); the other state is translationally shifted by one site and hence the dimers cover sites (2n,2n+1)(2n,2n+1) for integer nn.

On a finite open-boundary conditions (OBC) 2N2N sites chain, we can write the even ground state as:

|ψe=|S12|S34|S2N1,2N\ket{\psi_{e}}=\ket{S_{12}}\ket{S_{34}}\ldots\ket{S_{2N-1,2N}} (S45)

where |Sij=12(ijij)\ket{S_{ij}}=\frac{1}{\sqrt{2}}\left(\uparrow_{i}\downarrow_{j}-\downarrow_{i}\uparrow_{j}\right) is the singlet configuration of 2 1/2 spins.

The amplitude for a configuration |C{σ}=|σ1σ2σ2N\ket{C_{\{\sigma\}}}=\ket{\sigma_{1}\sigma_{2}\ldots\sigma_{2N}} where σi={,}\sigma_{i}=\{\uparrow,\downarrow\} is:

C|ψe=12N×(1)sC\braket{C}{\psi_{e}}=\frac{1}{2^{N}}\times(-1)^{s_{C}} (S46)

where sCs_{C} counts the number of \downarrow spins on odd sites.

We can obtain the groundstates of the MG chain within the Gutzwiller slave-fermion approach by using the following BdG slave-fermion mean-field Hamiltonian:

HBdG=ij,σtij(ciσcjσ+h.c.)ijΔij(cicj+cjci+h.c.)μiσciσciσH_{BdG}=-\sum_{ij,\sigma}t_{ij}\left(c^{\dagger}_{i\sigma}c_{j\sigma}+\text{h.c.}\right)-\sum_{ij}\Delta_{ij}\left(c^{\dagger}_{i\uparrow}c^{\dagger}_{j\downarrow}+c^{\dagger}_{j\uparrow}c^{\dagger}_{i\downarrow}+\text{h.c.}\right)-\mu\sum_{i\sigma}c^{\dagger}_{i\sigma}c_{i\sigma} (S47)

where we take tn,n+1=1t_{n,n+1}=1, Δn,n+1=1\Delta_{n,n+1}=-1 and μ=50000\mu=-50000.

First we produce the projected finite MPS on a small chain N=16N=16 and check the amplitudes for all configurations have the proper magnitude and sign. We can see in figure S7 that the relative error in the magnitude of amplitudes is at most 1.5×1051.5\times 10^{-5}.

Refer to caption
Figure S7: Histogram of the relative differences of amplitudes of the projected groundstate of eqn.  S47 with tn,n+1=1t_{n,n+1}=1, Δn,n+1=1\Delta_{n,n+1}=-1 and μ=50000\mu=-50000 found for all configurations on N=16N=16 sites.

We now obtain the iMPS representation using the procedures discussed in the main body of the text.

In particular we obtain the leading left/right eigenvectors of the 1-site transfer matrix and work with the positive left and positive right eigenvector and employ the standard iMPS orthogonalization. This produces the iMPS for the translationally invariant (NN\rightarrow\infty) state |ψe+T|ψe\ket{\psi_{e}}+T\ket{\psi_{e}} where TT translates the wave-function by one site .

The entanglement spectrum for this state is exactly given by: [log(2),log(4),log(4)][\log(2),\log(4),\log(4)]. We report here the difference between the entanglement spectrum we obtain from iMPS and the true entanglement spectrum:

δes=(6.59360277×109,5.50256990×106,5.48941290×106)\delta_{\text{es}}=(6.59360277\times 10^{-9},-5.50256990\times 10^{-6},5.48941290\times 10^{-6}) (S48)

We also compute the energy in the thermodynamic limit. The ground state energy of the MG chain is exactly given by 38-\frac{3}{8}. We obtain the following:

EMG=383.438739293315507×1010E_{MG}=-\frac{3}{8}-3.438739293315507\times 10^{-10} (S49)

We obtain the left-canonical tensors in an arbitrary gauge. After a suitable rotation ALUALUA_{L}\leftarrow U^{\dagger}A_{L}U, they are as follows:

AL=(0eiϕ100000.707103660eiϕ200)AL=(00eiϕ20.707104854eiϕ100000)A^{\downarrow}_{L}=\begin{pmatrix}0&e^{i\phi_{1}}&0\\ 0&0&0\\ 0.707103660e^{-i\phi_{2}}&0&0\end{pmatrix}A^{\uparrow}_{L}=\begin{pmatrix}0&0&e^{i\phi_{2}}\\ -0.707104854e^{-i\phi_{1}}&0&0\\ 0&0&0\end{pmatrix} (S50)

Note that since the ground state is a singlet, there will be the same numbers of \downarrow and \uparrow tensors and hence the phases ϕ1\phi_{1} and ϕ2\phi_{2} will cancel out. We can further rotate the matrices by a unitary that changes the sign of the first column:

U=(i00010001)U=\begin{pmatrix}i&0&0\\ 0&1&0\\ 0&0&1\end{pmatrix} (S51)

so that finally, we can put the 1-site left-canonical translationally invariant iMPS tensors in the form:

AL=(0100001/2+ϵ100)AL=(0011/2+ϵ200000)A^{\downarrow}_{L}=\begin{pmatrix}0&1&0\\ 0&0&0\\ -1/\sqrt{2}+\epsilon_{1}&0&0\end{pmatrix}A^{\uparrow}_{L}=\begin{pmatrix}0&0&1\\ 1/\sqrt{2}+\epsilon_{2}&0&0\\ 0&0&0\end{pmatrix} (S52)

where ϵ1=2.91581×106\epsilon_{1}=-2.91581\times 10^{-6} and ϵ2=2.91918×106\epsilon_{2}=2.91918\times 10^{-6}. A further rotation brings the above tensors in their Krauss operator form ruiz2011tensor . To obtain the uniform tensors, notice that since ΛAR=ALΛ\Lambda A_{R}=A_{L}\Lambda, with ALA_{L} and ARA_{R} left and right canonical matrices respectively, we can obtain a “uniform” iMPS tensors as A=Λ1ALΛ=ΛARΛ1A=\sqrt{\Lambda}^{-1}A_{L}\sqrt{\Lambda}=\sqrt{\Lambda}A_{R}\sqrt{\Lambda}^{-1}. In this way we regain the MPS representation found in ref. karimipour2008matrix, :

AL=(0100001+ϵ100)AL=(0011+ϵ200000)\begin{split}A^{\downarrow}_{L}=\begin{pmatrix}0&1&0\\ 0&0&0\\ -1+\epsilon_{1}&0&0\end{pmatrix}\\ A^{\uparrow}_{L}=\begin{pmatrix}0&0&1\\ 1+\epsilon_{2}&0&0\\ 0&0&0\end{pmatrix}\end{split} (S53)

and which by a further rotation can be brought to the form given in ref. perez2006matrix, .

Decomposing our iMPS into pure states, we can also obtain the 22-site translational invariant tensors (see ref. perez2006matrix, , ref. ruiz2011tensor, ). The 2-site unit cell uniform iMPS tensors for the even pure state (cuts between dimers) and odd pure state (cuts inside dimers) are given by:

Aevenσσ=BσCσAoddσσ=CσBσ\begin{split}A^{\sigma\sigma^{\prime}}_{\text{even}}=B^{\sigma}C^{\sigma^{\prime}}\\ A^{\sigma\sigma^{\prime}}_{\text{odd}}=C^{\sigma}B^{\sigma^{\prime}}\end{split} (S54)

with the BB and CC tensors numerically obtained as:

B=(1+ϵ10)B=(10)B^{\downarrow}=\begin{pmatrix}1+\epsilon_{1}&0\end{pmatrix}\qquad B^{\uparrow}=\begin{pmatrix}1&0\end{pmatrix} (S55)
C=(01)C=(1+ϵ20)C^{\downarrow}=\begin{pmatrix}0\\ 1\end{pmatrix}\qquad C^{\uparrow}=\begin{pmatrix}-1+\epsilon_{2}\\ 0\end{pmatrix} (S56)

where ϵ1=7.1400000000165775×106\epsilon_{1}=-7.1400000000165775\times 10^{6} and ϵ2=1.09999999997612×106\epsilon_{2}=1.09999999997612\times 10^{-6}. The bond dimension for the 1-site translationally invariant iMPS representation is naturally larger.