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

Disorder effects on the Topological Superconductor with Hubbard Interactions

Yiting Deng College of Physics, Sichuan University, Chengdu, Sichuan 610064, China heyan$˙[email protected]    Yan He College of Physics, Sichuan University, Chengdu, Sichuan 610064, China heyan$˙[email protected]
Abstract

We study the two-dimensional disordered topological superconductor with Hubbard interactions. When the magnitude of the pairing potential is tuned to special values, this interacting model is exactly solvable even when disorders are imposed on the potential term or coupling constants. The topology of this model is investigated in detail by the real space Chern number formula, which computes the topological index of disordered systems to high precisions. It is found that the disorders can drive the system from topological trivial phase to a non-trivial phase, which generalizes the topological Anderson phenomena to interacting models. The self-consistent Born approximation is also employed to understand the influence of the disorders on the parameters of the interacting topological superconductor. It provide an alternative way to understand the topological transitions at weak disordered region.

I Introduction

In the past decade, there are tremendous progress that have been achieved in the area of the topological properties of materials Hasan and Kane (2010); Qi and Zhang (2011). However, most of these studies are restricted to the free fermion systems or weakly interacting systems. It is well known that the understanding of strongly correlated systems such as Hubbard model Arovas et al. (2022); Hirsch (1985); Schulz (1990); Fradkin (2013) is very difficult since the perturbation theory is not reliable for these systems due to the lack of small expansion parameters. Recently, there emerges a series of topological models with Hubbard interactions which are exact solvable due to the existence of flat bands. This type of models are partly inspired by the Kitaev model Kitaev (2006) on the honeycomb lattice and are extensively studied by several authors Chen et al. (2018); Li et al. (2019); Miao et al. (2017, 2019); Ezawa (2018). They provide some interesting examples for understanding how the interactions affect the topological properties. Along this line, we also proposed an exactly solvable topological superconductor with Hubbard interaction Deng and He (2021), which enriches this class of models.

In the real materials, disorders or impurities are un-avoidable. In view of this, the studies of the disordered systems are also an important topic in condensed matter physics. A classical example of disorder effects is the Anderson localization Abrahams et al. (1979) which induces metal-insulator transitions in free fermion systems. An intriguing question is that how the disorders will affect the exact solvable interacting topological models we mentioned above. It is well-known that the topological phase can persist until the disorder strength reaches some critical values. After that, the system will return to the trivial phase. However, it is also found that the disorders can induce nontrivial topology in the system that is trivial in the clean limit. Ever since the pioneer work of the so-called topological Anderson insulator Li et al. (2009); Shen (2012) has been carried out, many other theoretical models have been proposed to understand these disorder induced topological phases Chen et al. (2017a, b); Klier et al. (2019); Guo et al. (2010); Guo (2010); Shapourian and Hughes (2016); Chen et al. (2015). The disorder induced topological phase transitions were also discussed in works Louvet et al. (2016); Ryu and Nomura (2012). In this paper, we will investigate how the disorders influence the topology of our interacting topological superconductor. The exact solvability allows us to perform a detailed study of the interplay between the interactions and disorders. Furthermore, it is also possible to make reliable calculation when the disorders are imposed on the coupling constants.

The rest of this paper is organized as follows. In section II, we present the TS model with Hubbard interactions and show that this model is exactly solvable when flat band condition is satisfied. In section III, we introduce the real space Chern number formula which is suitable for investigations of the topology of disordered systems. The self-consistent Born approximation is also discussed, which can provide more insights about the influence of disorders. The numerical results of the disordered interacting TS model based on the above mentioned two methods are presented in section IV. Finally, we briefly conclude in section V.

II The topological superconductor with Hubbard interaction

In this section, we introduce the model of topological superconductor which can support flat bands with a suitable choice of parameters. Owing to the presence of flat bands, this type of model is exactly solvable when a Hubbard interaction term is included Miao et al. (2017); Ezawa (2018). The Hamiltonian of the topological superconductor (TS) in momentum space can be expressed in the conventional Bogoliubov-de Gennes form as Deng and He (2021)

TS=𝐤a,bψa,𝐤(HTS(𝐤)abψb,𝐤,HTS(𝐤)=(ξ(𝐤)Δϕ(𝐤)Δϕ(𝐤)ξ(𝐤))\displaystyle\mathcal{H}_{TS}=\sum_{\mathbf{k}}\sum_{a,b}\psi_{a,\mathbb{\mathbf{k}}}^{\dagger}(H_{TS}(\mathbf{k})_{ab}\psi_{b,\mathbf{k}},\quad H_{TS}(\mathbf{k})=\left(\begin{array}[]{cc}\xi(\mathbf{k})&\Delta\phi(\mathbf{k})\\ \Delta\phi^{\dagger}(\mathbf{k})&-\xi(\mathbf{-k})^{*}\end{array}\right) (3)

Here we have defined ψ𝐤=(c1,𝐤,c2,𝐤,c1,𝐤,c2,𝐤)T\psi_{\mathbf{k}}=(c_{1,\mathbf{k}},c_{2,\mathbf{k}},c_{1,-\mathbf{k}}^{\dagger},c_{2,-\mathbf{k}}^{\dagger})^{T}. The first sub-indices denote the two different orbitals on each site. The lattice momentum 𝐤\mathbf{k} takes values from the Brillouin zone of a square lattice. The hopping and pairing terms are given by

ξ(𝐤)=j=13Rjσj,ϕ(𝐤)=i(R1σ2R2σ1)R3σ0\displaystyle\xi(\mathbf{k})=\sum_{j=1}^{3}R_{j}\sigma_{j},\quad\phi(\mathbf{k})=i(R_{1}\sigma_{2}-R_{2}\sigma_{1})-R_{3}\sigma_{0} (4)

Where σj\sigma_{j} with j=1,2,3j=1,2,3 are Pauli matrices and σ0\sigma_{0} is 2 by 2 identity matrix. The 3 coefficients RjR_{j} are chosen as follows

R1=m+coskx+cosky,R2=sinkx,R3=sinky\displaystyle R_{1}=m+\cos k_{x}+\cos k_{y},\quad R_{2}=\sin k_{x},\quad R_{3}=\sin k_{y} (5)

One can see that the hopping term alone is equivalent to the famous Qi-Wu-Zhang model of Chern insulator Qi et al. (2006).

It is also convenient to rewrite the topological superconductor model as

HTS=HCI+Hpair,\displaystyle H_{\textrm{TS}}=H_{\textrm{CI}}+H_{\textrm{pair}}, (6)
HCI=R1Γ31+R2Γ32+R3Γ03,\displaystyle H_{\textrm{CI}}=R_{1}\Gamma_{31}+R_{2}\Gamma_{32}+R_{3}\Gamma_{03}, (7)
Hpair=Δ(R1Γ22R2Γ21+R3Γ10)\displaystyle H_{\textrm{pair}}=-\Delta(R_{1}\Gamma_{22}-R_{2}\Gamma_{21}+R_{3}\Gamma_{10}) (8)

Here we defined Γij=τiσj\Gamma_{ij}=\tau_{i}\otimes\sigma_{j} with Pauli matrices τi\tau_{i} and σj\sigma_{j} acting on the Nambu spinor space and the orbital space respectively. One can see that the 4 energy bands of HTSH_{TS} is given by

E=±(Δ±1)R12+R22+R32\displaystyle E=\pm(\Delta\pm 1)\sqrt{R_{1}^{2}+R_{2}^{2}+R_{3}^{2}} (9)

Therefore, when Δ=±1\Delta=\pm 1, there will appear two flat bands E=0E=0. Note that the pairing term is chosen to share the same structure as the hopping term. This is the reason why the HTSH_{TS} can support two flat bands with special choice of parameters.

The topological properties of the model with Eq.(3) is the same as the Qi-Wu-Zhang model. One find that the Chern number C=±1C=\pm 1 when |m|<2|m|<2 and C=0C=0 otherwise. When C=±1C=\pm 1, the above model is topologically nontrivial, which also indicates that chiral edge modes will appear if we impose an open boundary condition to the system. One the other hand, for C=0C=0, the system is topologically trivial with no edge modes.

Up to now, the above topological superconductor is a model of spinless fermions. It is straightforward to include the spin degree of freedom which gives rise to the Hamiltonian of the first quantization form as

H=HTSs0\displaystyle H=H_{TS}\otimes s_{0} (10)

Here s0s_{0} is a 2×22\times 2 identity matrix acting on the spinor space and HTSH_{TS} is from Eq. (3). Now we can define the Hubbard interaction as the on-site repulsion between spin up and spin down fermions as follows

Hint=U𝐧,j(c𝐧,j,c𝐧,j,12)(c𝐧,j,c𝐧,j,12)\displaystyle H_{int}=U\sum_{\mathbf{n},j}(c_{\mathbf{n},j,\uparrow}^{\dagger}c_{\mathbf{n},j,\uparrow}-\frac{1}{2})(c_{\mathbf{n},j,\downarrow}^{\dagger}c_{\mathbf{n},j,\downarrow}-\frac{1}{2}) (11)

Since the Hubbard interaction term is expressed in the real space, it is more convenient to transfer the hopping and pairing terms of Eq.(10) to the real space. In the second quantization form, The resulting HCIH_{\textrm{CI}} and HpairH_{\textrm{pair}} terms can be written as

HCI=2𝐧,s(c𝐧,sσ1iσ22c𝐧+x^,s+c𝐧,sσ1iσ32c𝐧+y^,s+h.c.)+2𝐧,smc𝐧,sσ1c𝐧,s\displaystyle H_{\textrm{CI}}=2\sum_{\mathbf{n},s}\Big{(}c^{\dagger}_{\mathbf{n},s}\frac{\sigma_{1}-i\sigma_{2}}{2}c_{\mathbf{n}+\hat{x},s}+c^{\dagger}_{\mathbf{n},s}\frac{\sigma_{1}-i\sigma_{3}}{2}c_{\mathbf{n}+\hat{y},s}+h.c.\Big{)}+2\sum_{\mathbf{n},s}mc^{\dagger}_{\mathbf{n},s}\sigma_{1}c_{\mathbf{n},s} (12)
Hpair=2Δ𝐧,s(c𝐧,siσ2σ12c𝐧+x^,s+c𝐧,siσ2+iσ02c𝐧+y^,s+h.c.)\displaystyle H_{\textrm{pair}}=2\Delta\sum_{\mathbf{n},s}\Big{(}c^{\dagger}_{\mathbf{n},s}\frac{i\sigma_{2}-\sigma_{1}}{2}c^{\dagger}_{\mathbf{n}+\hat{x},s}+c^{\dagger}_{\mathbf{n},s}\frac{i\sigma_{2}+i\sigma_{0}}{2}c^{\dagger}_{\mathbf{n}+\hat{y},s}+h.c.\Big{)}
+Δ𝐧,s(mc𝐧,s(iσ2)c𝐧,s+h.c.)\displaystyle\qquad+\Delta\sum_{\mathbf{n},s}\Big{(}mc^{\dagger}_{\mathbf{n},s}(i\sigma_{2})c^{\dagger}_{\mathbf{n},s}+h.c.\Big{)} (13)

Here 𝐧=(nx,ny)\mathbf{n}=(n_{x},n_{y}) labels the lattice site on a square lattice and s=,s=\uparrow,\downarrow labels the spin up and down. The two-component fermion operator is defined as c𝐧,s=(c𝐧,1,s,c𝐧,2,s)Tc_{\mathbf{n},s}=(c_{\mathbf{n},1,s},c_{\mathbf{n},2,s})^{T}, where the second index labels the two orbitals. We also introduce x^\hat{x} and y^\hat{y} representing the unit vector along the xx and yy direction respectively. Now the full interacting topological superconductor can be summarized as

Hfull=HCI+Hpair+Hint\displaystyle H_{\textrm{full}}=H_{\textrm{CI}}+H_{\textrm{pair}}+H_{\textrm{int}} (14)

In order to understand the reason why the above interacting model is exactly solvable, we introduce the Majorana fermion operators as follows

c𝐧,1,s=a𝐧,1,s+ib𝐧,1,s,c𝐧,1,s=a𝐧,1,sib𝐧,1,s\displaystyle c_{\mathbf{n},1,s}=a_{\mathbf{n},1,s}+ib_{\mathbf{n},1,s},\quad c_{\mathbf{n},1,s}^{\dagger}=a_{\mathbf{n},1,s}-ib_{\mathbf{n},1,s} (15)
c𝐧,2,s=b𝐧,2,s+ia𝐧,2,s,c𝐧,2,s=b𝐧,2,sia𝐧,2,s\displaystyle c_{\mathbf{n},2,s}=b_{\mathbf{n},2,s}+ia_{\mathbf{n},2,s},\quad c_{\mathbf{n},2,s}^{\dagger}=b_{\mathbf{n},2,s}-ia_{\mathbf{n},2,s} (16)

If the Majorana fermion operators is substituted into the hopping and pairing terms of Eq.(12) and (13), we find that the terms involving a𝐧,j,sa_{\mathbf{n},j,s} are all vanished when the flat band condition Δ=1\Delta=1 is satisfied (Detailed calculations can be found in Deng and He (2021)). In terms of the Majorana fermions, the Hubbard interaction can be expressed as

Hint=U𝐧,j(2ia𝐧,j,b𝐧,j,)(2ia𝐧,j,b𝐧,j,)\displaystyle H_{\textrm{int}}=U\sum_{\mathbf{n},j}(2ia_{\mathbf{n},j,\uparrow}b_{\mathbf{n},j,\uparrow})(2ia_{\mathbf{n},j,\downarrow}b_{\mathbf{n},j,\downarrow}) (17)

If we define D𝐧,j=4ia𝐧,j,a𝐧,j,D_{\mathbf{n},j}=4ia_{\mathbf{n},j,\uparrow}a_{\mathbf{n},j,\downarrow}, it is easy to see that [D𝐧,j,H]=0[D_{\mathbf{n},j},H]=0 in all sites and orbitals. Therefore, all D𝐧,jD_{\mathbf{n},j} are actually constant cc-numbers. Due to the anti-commutation relations of Majorana fermions, one can see that D𝐧,j2=1D_{\mathbf{n},j}^{2}=1 and thus D𝐧,j=±1D_{\mathbf{n},j}=\pm 1. When the distribution of D𝐧,jD_{\mathbf{n},j} is given, the Hubbard interaction term becomes quadratic in terms of fermion operators, which makes the interacting model solvable.

Among all possible the D𝐧,jD_{\mathbf{n},j} configurations, there are two special ones with the potential to give rise the ground state of the interacting topological superconductor. The first one is the uniform distribution with D𝐧,j=1D_{\mathbf{n},j}=1 for all sites and orbitals, which is called the ferromagnetic (FM) configuration. The other one is the staggered distribution with D𝐧,1=1D_{\mathbf{n},1}=1 and D𝐧,2=1D_{\mathbf{n},2}=-1 for all sites, which is called anti-ferromagnetic (AFM) configuration.

If we transfer the Majorana fermions back to ordinary fermions, the Hubbard interaction term in the first quantized form for the FM and AFM configurations can be written as

HintFM=U4(Γ00Γ13)s2\displaystyle H_{\textrm{int}}^{\textrm{FM}}=-\frac{U}{4}(\Gamma_{00}-\Gamma_{13})\otimes s_{2} (18)
HintAFM=U4(Γ03Γ10)s2\displaystyle H_{\textrm{int}}^{\textrm{AFM}}=-\frac{U}{4}(\Gamma_{03}-\Gamma_{10})\otimes s_{2} (19)

With this result, the interacting topological superconductor of Eq.(14) can be converted into a simple free fermion model, for which the Chern number can be easily computed. Accordingly, we map out the phase diagram of the interacting topological superconductor in Fig.1. The topological non-trivial phase is roughly located inside the area of m(2,2)m\in(-2,2) and U(4,4)U\in(-4,4). The yellow and blue regions indicate the phase with Chern number C=±2C=\pm 2, while the green region indicate the topological trivial phase with C=0C=0. We would like to mention that the true ground state of the model is the AFM configuration.

Refer to caption
Refer to caption
Figure 1: The phase diagram of the interacting topological superconductor of Eq.(14) with AFM configurations (left panel) and FM configuration (right panel). The yellow and blue regions correspond the Chern number C=±2C=\pm 2. The green region corresponds to C=0C=0. The system size is Lx=Ly=20L_{x}=L_{y}=20.

In this paper, we will focus on the disorder effects of the interacting topological superconductor of Eq.(14). We always assume that the flat band condition Δ=1\Delta=1 is satisfied, such that the interacting model can be exactly solved. In order to keep the solvability of this interacting model, we impose the disorders either on the coefficient mm or the coupling constant UU. More explicitly, we will make the following replacement mm𝐧=m+ω𝐧m\to m_{\mathbf{n}}=m+\omega_{\mathbf{n}} or UU𝐧=U+ω𝐧U\to U_{\mathbf{n}}=U+\omega_{\mathbf{n}}. In other words, we consider two types of disorder models as follows

Hdis1=Hfull+ω𝐧(Γ31Γ22)s0,\displaystyle H_{\textrm{dis1}}=H_{\textrm{full}}+\omega_{\mathbf{n}}(\Gamma_{31}-\Gamma_{22})\otimes s_{0}, (20)
Hdis2=Hfullω𝐧4(Γ03Γ10)s2\displaystyle H_{\textrm{dis2}}=H_{\textrm{full}}-\frac{\omega_{\mathbf{n}}}{4}(\Gamma_{03}-\Gamma_{10})\otimes s_{2} (21)

Here ω𝐧\omega_{\mathbf{n}} located at each lattice site are random variables with uniform distribution inside the interval [W/2,W/2][-W/2,W/2]. WW is the disorder strength. We have numerically checked that the AFM configuration is also the true ground state even when the disorders are introduced. Therefore, the disorder potential term in Eq.(21) takes the form of Eq.(19).

III Numerical methods for disordered systems

In this section, we discuss two different methods to investigate the topology of the interacting TS model with disorders. Usually, the Chern number is defined as an integral in the momentum space, but this form is not suitable for disordered system. Here we introduce a Chern number formula in real space, which can compute this topological index to high precision. The other method is the so-called self-consistent Born approximation, which can convert the disorder system into a uniformed system with renormalized parameters. Then, it is convenient to understand how the disorder affects topological phase transitions.

The Chern number formula defined in 𝐤\mathbf{k}-space can be written as

C=12πid2kTr(P𝐤[xP𝐤,yP𝐤])\displaystyle C=\frac{1}{2\pi i}\int d^{2}k\,\textrm{Tr}\Big{(}P_{\mathbf{k}}[\partial_{x}P_{\mathbf{k}},\partial_{y}P_{\mathbf{k}}]\Big{)} (22)

Here P𝐤P_{\mathbf{k}} is the projection operator in the 𝐤\mathbf{k}-space, which is define as

P𝐤=En<Ef|ψn(𝐤)ψn(𝐤)|\displaystyle P_{\mathbf{k}}=\sum_{E_{n}<E_{f}}|\psi_{n}(\mathbf{k})\rangle\langle\psi_{n}(\mathbf{k})| (23)

where |ψn|\psi_{n}\rangle represents the nn-th eigenvector with the eigen-energy EnE_{n}, and EfE_{f} represents the Fermi energy. This operator P𝐤P_{\mathbf{k}} projects the wave-function into the subspace spanned by the filled energy bands.

For the disordered systems, it is not possible to directly make 𝐤\mathbf{k}-space calculations. Therefore, one must seek a method to compute the Chern number in real space. If we diagonalize the Hamiltonian in real space, we can also construct a projection operator PP in real space. Next, we can approximate the derivative by finite differences. Then we will arrived at a Chern number formula in real space. This approach was suggested a few years ago by Prodan Prodan et al. (2010); Prodan (2011). We will provide a detailed derivation of this formula in the appendix A. In this section, we simply give out the real space Chern number formula as follows

C=1πIm Tr[P(KxP)(KyP)]\displaystyle C=-\frac{1}{\pi}\textrm{Im Tr}\Big{[}P(K_{x}\cdot P)(K_{y}\cdot P)\Big{]} (24)

Here the dot means element-wise product of the two matrices. “Im” means taking the imaginary part of the following quantities. The matrices KxK_{x} and KyK_{y} are given in the appendix A.

Now we turn to the self-consistent Born approximation which is widely used in the many works such as Chen et al. (2017a, b); Klier et al. (2019). This method usually provide a useful way to understand topological phase transitions driven by disorders. The essence of this method is to incorporate the effects of the disorder into a self-energy. Then the self-energy will in turn modify the parameters of the original model. Now the disordered model is actually converted into a uniform model with renormalized parameters. Then we can determine the boundary of topological phase transition by the phase diagram of Fig. 1 with renormalized mm or UU. In other words, the way that renormalized mm or UU depends on the disorder strength WW directly reflect the influence of disorders on topological phase transitions.

In the self-consistent Born approximation, the self-energy Σ\Sigma is defined as Shen (2012)

(EfHfull(𝐤)Σ)1=(EfHfull(𝐤)V𝐧)1dis\displaystyle(E_{f}-H_{\textrm{full}}(\mathbf{k})-\Sigma)^{-1}=\langle(E_{f}-H_{\textrm{full}}(\mathbf{k})-V_{\mathbf{n}})^{-1}\rangle_{\textrm{dis}} (25)

Here Hfull(𝐤)H_{\textrm{full}}(\mathbf{k}) is the interacting TS model of Eq.(14) in the momentum space. V𝐧V_{\mathbf{n}} is the disorder potential in real space. The dis\langle\cdots\rangle_{\textrm{dis}} denotes the disorder average. For the two disorder models of Eq.(20) and (21), the disorder potential can be written as

V𝐧=ω𝐧Λs,(s=1,2)\displaystyle V_{\mathbf{n}}=\omega_{\mathbf{n}}\Lambda_{s},\quad(s=1,2)
Λ1=(Γ31Γ22)s0,Λ2=14(Γ03Γ10)s2\displaystyle\Lambda_{1}=(\Gamma_{31}-\Gamma_{22})\otimes s_{0},\quad\Lambda_{2}=-\frac{1}{4}(\Gamma_{03}-\Gamma_{10})\otimes s_{2} (26)

where ω𝐧\omega_{\mathbf{n}} is the random number we defined before.

If we expand Eq.(25) to the first order, we find that the disorder-induced self-energy is determined by the equation Groth et al. (2009); Hua et al. (2019) as follows

Σ=W2121(2π)2BZd2kΛs[EFHfull(𝐤)Σ]1Λs\displaystyle\Sigma=\frac{W^{2}}{12}\frac{1}{(2\pi)^{2}}\int_{\textrm{BZ}}d^{2}k\,\Lambda_{s}[E_{F}-H_{\textrm{full}}(\mathbf{k})-\Sigma]^{-1}\Lambda_{s} (27)

Here EFE_{F} is the Fermi energy which is close to zero in our model. The integration is over the whole Brillouin zone.

For the disorder model of Eq.(20), by iterating Eq.(27), we can solve the self-energy and find that Σ=δmΛ1\Sigma=\delta m\,\Lambda_{1} plus some constant matrices which can be ignored. Now the effective Hamiltonian is H=Hfull+ΣH=H_{\textrm{full}}+\Sigma. Therefore, one can see that the self-energy only renormalize the parameter mm to m~\widetilde{m}, which is given by

m~=m+δm\displaystyle\widetilde{m}=m+\delta m (28)

Similarly, if we consider the disorder model of Eq.(21), we find that the self-energy Σ=δUΛ2\Sigma=\delta U\,\Lambda_{2} and the coupling UU is renormalized to U~=U+δU\widetilde{U}=U+\delta U.

IV Numerical results

Refer to caption
Figure 2: The Chern number as a function of the disorder strength WW for m=1.2m=1.2 (blue curve) and m=2.2m=2.2 (red curve). The flat band condition Δ=1\Delta=1 is assumed. The system size is Lx=Ly=40L_{x}=L_{y}=40. The error bar indicates standard deviation of 30 random samples. The blue and red circles represent the values of Chern number.

In this section, we present the numerical results generated by the two methods that we discussed in the previous section. We start from with the real space Chern number for the disordered interacting topological superconductor. First, as a simple example, we consider the disorder model of Eq.(20) with U=0U=0. In other words, we are working with Eq.(10) without the Hubbard interaction and the disorders are imposed on the potential term as in Eq.(20). Here we study the disorder effects on two different cases with m=1.2m=1.2 or m=2.2m=2.2 corresponding to topological or trivial phases of the clean TS model respectively. The Chern number as a function of disorder strength WW is shown in Fig. 2.

For the case of m=1.2m=1.2 (blue curve), we have the Chern number C=2C=2 in the clean limit. As the disorder strength WW increases, one can see that CC stays at value 22 for a while and then start to drop at about W=4.8W=4.8. It gradually decreases to zero at about W=9W=9 and after that CC stays at zero. Therefore, there exists a critical disorder region where the topological phase transition takes place. This indicates that when the disorders are strong enough, they can kill the topological nontrivial phases.

For the other case of m=2.2m=2.2 (red curve), the model is topologically trivial when W=0W=0. However, when the disorder strength increased, one can see that the Chern number rises from 0 to 22 at around W=1.7W=1.7 and stays at the value 2 for a while. It then starts to drop at W=5.8W=5.8 and back to zero at around W=8W=8. Within a certain range of disorder strength, a plateau of non-zero Chern number is maintained, which indicates that disorder can drive the system into a topologically nontrivial phase. This is very similar to the so-called topological Anderson insulator. But here it is realized in a topological superconductor instead of insulator.

Refer to caption
Refer to caption
Figure 3: The Chern number as a function of the disorder strength WW. Left panel: The disorder model is Eq.(20) with m=1.2,U=1.2m=1.2,\,U=1.2 (blue curve) and m=2.2,U=1.2m=2.2,\,U=1.2 (red curve). Right panel: The disorder model is Eq.(21) with m=1.2,U=1.2m=1.2,\,U=1.2 (blue curve) and m=1.2,U=4.1m=1.2,\,U=4.1 (red curve). The system size is Lx=Ly=40L_{x}=L_{y}=40. The error bars indicate standard deviation of 30 random samples.

Next we turn on the Hubbard interaction in the disorder model of Eq.(20). Since we assume Δ=1\Delta=1, this mode is still exactly solvable. The result of its real space Chern number is presented in the left panel of Fig.3. Here we still consider two different cases with m=1.2m=1.2 (blue curve) and m=2.2m=2.2 (red curve). For both cases, we assume U=1.2U=1.2. The behavior is very similar to the non-interacting case of Fig.2. For m=1.2m=1.2, the model is already topological for W=0W=0. When WW increase to around W=5W=5, the Chern number CC gradually drop to zero. One the other hand, the model is trivial for m=2.2m=2.2 in the clean limit. But the disorders will drive CC to increase from zero to 22 and then return to zero again. The plateau of C=2C=2 within a certain range of WW suggests that the disorder can induce a topologically nontrivial phase even when the Hubbard interaction is presented. This result generalizes the phenomena of topological Anderson transition to interacting models.

Since the interacting TS model is exact solvable, we can also impose the disorders on the coupling constant, as the disorder model of Eq.(21). The Chern number of this model is shown in the right panel of Fig. 3. The blue and red curve corresponds to the U=1.2U=1.2 and U=4.1U=4.1 respectively. In both cases, we assume m=1.2m=1.2. One can see that the U=1.2U=1.2 system has C=2C=2 in the clean limit and CC gradually decreasing to zero inside the region of 20<W<3020<W<30. According to the phase diagram of Fig.1, the system with U=4.1U=4.1 is topologically trivial in the clean limit. Again, one can see that the disorders can generate a stable plateau of C=2C=2 inside the region of 10<W<2010<W<20. This is again the signature of topology driven by disorders. In the model of Eq.(21), the disorders come from the random coupling constant, which is usually difficult to analyze in other types of models.

We also would like to mention that the error bars in both Fig. 2 and 3 represent the standard deviations of 30 random samples. The error bars in the transition regions are larger than the ones in the stable regions.

Refer to caption
Refer to caption
Refer to caption
Figure 4: Top left: The renormalized m~\widetilde{m} as a function of the disorder strength WW with disorder of Eq.(20). We assume m=2.2,U=0m=2.2,\,U=0 (blue triangle) and m=2.2,U=1.2m=2.2,\,U=1.2 (yellow dot). Top right: The renormalized U~\widetilde{U} as a function of WW with disorder of Eq.(21). Here we assume m=1.2m=1.2. Bottom: The energy bands of Eq.(14) with renormalized U~\widetilde{U} as a function of kyk_{y}. The parameters are m=1.2,U=4.1,W=11m=1.2,U=4.1,W=11.

At last, we present the numerical results of effective medium theory based on the self-consistent Born approximation. We first consider the disorder model of Eq.(20) which renormalizes the parameter mm. The renormalized m~\widetilde{m} as a function of WW is displayed in the top left panel of Fig.4. One can see that for both m=2.2,U=0m=2.2,\,U=0 (blue triangle) and m=2.2,U=1.2m=2.2,\,U=1.2 (yellow dot) cases, the two renormalized m~\widetilde{m} are almost identical. The dashed line represents the topological phase boundary of m~=2\widetilde{m}=2. The corresponding critical disorder strength is W2W\approx 2, which is close the region where C=0C=0 rises to C=2C=2 in the left panel of Fig.3. For larger WW, we find that 2<m~<2-2<\widetilde{m}<2, which suggest the system stays in the topological phase even when WW increases up to 10. On the other hand the Chern number curve in the left panel of Fig.3 shows that CC drop to zero at W6W\approx 6. This contradiction shows that the self-consistent Born approximation only works for the relatively weak disorders. It cannot capture the correct topological features when WW is too strong.

Similarly, we can also study the disorder model of Eq.(21) which renormalizes the coupling UU. The resulting U~\widetilde{U} is shown in the right panel of Fig.4. The dashed line represents the phase boundary of U~=3.9\widetilde{U}=3.9. One can see the corresponding critical W10W\approx 10 which is close to the first transition in the right panel of Fig.3. Again, the renormalized U~\widetilde{U} fail to capture the second transition at strong WW.

Another advantage of the self-consistent Born approximation is that it can eliminate the disorders such that the model becomes a uniform one again. Because of this, we can put the model on a cylinder geometry with open boundary along xx-axes and closed boundary along yy-axes. For example, the disorder model of Eq.(21) is converted to Eq.(14) with renormalized U~\widetilde{U}. Now the Hamiltonian as a function of kyk_{y} is ready to be diagonalized. The resulting band structure is shown in the bottom panel of Fig.4. One can see the appears of the edge modes due to the non-trivial topology. It is also evident that the system preserves its flat bands even with the introduction of disorders, thereby maintaining its exact solvability. Similar conclusions can be drawn for the disorder model of Eq.(20).

V conclusion

We have studied the topological phase transitions of the disordered topological superconductor with Hubbard interaction where the disorders are either imposed on the potential term or the coupling constants. Throughout the whole paper, we stick to the flat-band condition which guarantees the exact solvability of this model even when disorders are included. It is found that the strong enough disorders are able to suppress the topological phases. For topological trivial system in the clean limit, the disorders also induce a phase with non-zero Chern number. This can be thought as a generalization of the topological Anderson transitions to interacting models. The results of real space Chern number is also confirmed by the self-consistent Born approximation. In the later method, we also demonstrated the influence of disorders on the system parameters and band structures.

Acknowledgements.
This work is supported by the National Natural Science Foundation of China under Grant No. 11874272.

Appendix A Derivation of Real space Chern Number

We start from the Chern number formula in momentum space, which we repeated here as follows

C=12πid2kTr(P𝐤[xP𝐤,yP𝐤])\displaystyle C=\frac{1}{2\pi i}\int d^{2}k\,\textrm{Tr}\Big{(}P_{\mathbf{k}}[\partial_{x}P_{\mathbf{k}},\partial_{y}P_{\mathbf{k}}]\Big{)} (29)

here μ=kμ\partial_{\mu}=\frac{\partial}{\partial k_{\mu}} with μ=x,y\mu=x,y. By discretizing the continuous model, the Brillouin zone of a 2D square lattice is divided into a N×NN\times N grid, with separation h=2π/Nh=2\pi/N. The derivative of kxk_{x} can be approximated as

xP𝐤δP𝐤m=1Qcm2h[P𝐤+mhx^P𝐤mhx^]\displaystyle\partial_{x}P_{\mathbf{k}}\approx\delta P_{\mathbf{k}}\equiv\sum_{m=1}^{Q}\frac{c_{m}}{2h}[P_{\mathbf{k}+mh\hat{x}}-P_{\mathbf{k}-mh\hat{x}}] (30)

Here x^\hat{x} is a unit vector along xx-axes. The coefficients cmc_{m} are some weights to be determined later. If we take Q=1Q=1 and c1=1c_{1}=1, we get back to a simple symmetric difference. In order to improve the precision, we assume Q=N/2Q=N/2 where NN is the number of lattice sites along the xx-axes. We can fourier transform P𝐤P_{\mathbf{k}} as

P𝐤=𝐫b𝐫ei𝐤𝐫\displaystyle P_{\mathbf{k}}=\sum_{\mathbf{r}}b_{\mathbf{r}}e^{i\mathbf{k}\cdot\mathbf{r}} (31)

Substituting Eq. (31) into Eq.(30) we find that

xP𝐤=i𝐫b𝐫xei𝐤𝐫\displaystyle\partial_{x}P_{\mathbf{k}}=i\sum_{\mathbf{r}}b_{\mathbf{r}}x\,e^{i\mathbf{k}\cdot\mathbf{r}} (32)
δP𝐤=𝐫b𝐫ei𝐤𝐫m=1Qcm2h(2i)sin(mhx)\displaystyle\delta P_{\mathbf{k}}=\sum_{\mathbf{r}}b_{\mathbf{r}}e^{i\mathbf{k}\cdot\mathbf{r}}\sum_{m=1}^{Q}\frac{c_{m}}{2h}(2i)\sin(mhx) (33)

The approximation in Eq.(30) implies the following condition

xm=1Qcm2h2sin(mhx)\displaystyle x\approx\sum_{m=1}^{Q}\frac{c_{m}}{2h}2\sin(mhx) (34)

Making a Taylor expansion of the right hand side of the above equation, we find that the above condition becomes a set of linear equations

mQcmm=1,mQcmm2j1=0\displaystyle\sum_{m}^{Q}c_{m}m=1,\quad\sum_{m}^{Q}c_{m}m^{2j-1}=0 (35)

from which we can numerically solve the weights cmc_{m} with m=1,,Qm=1,\cdots,Q.

In the real space representation, the projector P𝐤P_{\mathbf{k}} is replaced by its real space counterpart PP. The momentum shift of P𝐤P_{\mathbf{k}} can be achieved by the translational operator as

P𝐤+mhx^eimhXPeimhX\displaystyle P_{\mathbf{k}+mh\hat{x}}\to e^{-imhX}Pe^{imhX} (36)

Here X=diag{1,2,,N}INX=\textrm{diag}\{1,2,\,\cdots,\,N\}\otimes I_{N} with the identity matrix INI_{N} of dimension NN. Now the approximated derivative δP𝐤\delta P_{\mathbf{k}} in the real space representation can be written as

δP𝐤m=1Qcm2h(eimhXPeimhXeimhXPeimhX)ab\displaystyle\delta P_{\mathbf{k}}\to\sum_{m=1}^{Q}\frac{c_{m}}{2h}\Big{(}e^{-imhX}Pe^{imhX}-e^{imhX}Pe^{-imhX}\Big{)}_{ab}
=2im=1Qcm2hsinmh(xaxb)Pab\displaystyle\quad=2i\sum_{m=1}^{Q}\frac{c_{m}}{2h}\sin mh(x_{a}-x_{b})P_{ab} (37)

Here xax_{a} is the diagonal elements of matrix XX . In summary, we find that the

xP𝐤ihKxP,(Kx)ab=m=1Qcmsinmh(xaxb)\displaystyle\partial_{x}P_{\mathbf{k}}\to\frac{i}{h}K_{x}\cdot P,\quad(K_{x})_{ab}=\sum_{m=1}^{Q}c_{m}\sin mh(x_{a}-x_{b}) (38)

where the dot means element-wise matrix product. Similarly, we also have

yP𝐤ihKyP,(Ky)ab=m=1Qcmsinmh(yayb)\displaystyle\partial_{y}P_{\mathbf{k}}\to\frac{i}{h}K_{y}\cdot P,\quad(K_{y})_{ab}=\sum_{m=1}^{Q}c_{m}\sin mh(y_{a}-y_{b}) (39)

Here yay_{a} is the diagonal element of matrix Y=INdiag{1,2,,N}Y=I_{N}\otimes\textrm{diag}\{1,2,\,\cdots,\,N\}.

Plug the above two equations into Eq.(29), and notice that

Tr[P(KxP)(KyP)]=Tr[P(KyP)(KxP)]\displaystyle\textrm{Tr}\Big{[}P(K_{x}\cdot P)(K_{y}\cdot P)\Big{]}=\textrm{Tr}\Big{[}P(K_{y}\cdot P)(K_{x}\cdot P)\Big{]}^{\dagger} (40)

we finally arrived at the following Chern number formula

C=1πIm Tr[P(KxP)(KyP)]\displaystyle C=-\frac{1}{\pi}\textrm{Im Tr}\Big{[}P(K_{x}\cdot P)(K_{y}\cdot P)\Big{]} (41)

which is the claimed result of Eq.(24).

References

  • Hasan and Kane (2010) M. Z. Hasan and C. L. Kane, Rev. Mod. Phys. 82, 3045 (2010).
  • Qi and Zhang (2011) X.-L. Qi and S.-C. Zhang, Rev. Mod. Phys. 83, 1057 (2011).
  • Arovas et al. (2022) D. P. Arovas, E. Berg, S. A. Kivelson, and S. Raghu, Annual review of condensed matter physics 13, 239 (2022).
  • Hirsch (1985) J. E. Hirsch, Physical Review B 31, 4403 (1985).
  • Schulz (1990) H. Schulz, Physical review letters 64, 2831 (1990).
  • Fradkin (2013) E. Fradkin, Field Theories of Condensed Matter Physics, 2nd edition (Cambridge University Press, Cambradge, UK, 2013).
  • Kitaev (2006) A. Kitaev, Ann. Phys. 321, 2 (2006).
  • Chen et al. (2018) Z. Chen, X. Li, and T. K. Ng, Phys. Rev. Lett. 120, 046401 (2018).
  • Li et al. (2019) X.-H. Li, Z. Chen, and T. K. Ng, Phys. Rev. B 100, 094519 (2019).
  • Miao et al. (2017) J.-J. Miao, H.-K. Jin, F.-C. Zhang, and Y. Zhou, Phys. Rev. Lett. 118, 267701 (2017), URL https://link.aps.org/doi/10.1103/PhysRevLett.118.267701.
  • Miao et al. (2019) J.-J. Miao, D.-H. Xu, L. Zhang, and F.-C. Zhang, Physical Review B 99, 245154 (2019).
  • Ezawa (2018) M. Ezawa, Phys. Rev. B 97, 241113 (2018), URL https://link.aps.org/doi/10.1103/PhysRevB.97.241113.
  • Deng and He (2021) Y. Deng and Y. He, Physics Letters A 397, 127260 (2021).
  • Abrahams et al. (1979) E. Abrahams, P. W. Anderson, D. C. Licciardello, and T. V. Ramakrishnan, Phys. Rev. Lett. 42, 673 (1979), URL https://link.aps.org/doi/10.1103/PhysRevLett.42.673.
  • Li et al. (2009) J. Li, R.-L. Chu, J. K. Jain, and S.-Q. Shen, Physical Review Letters 102, 136806 (2009).
  • Shen (2012) S.-Q. Shen, Topological insulators (Springer, Berlin, 2012).
  • Chen et al. (2017a) R. Chen, D.-H. Xu, and B. Zhou, Physical Review B 95, 245305 (2017a).
  • Chen et al. (2017b) R. Chen, D.-H. Xu, and B. Zhou, Physical Review B 96, 205304 (2017b).
  • Klier et al. (2019) J. Klier, I. Gornyi, and A. Mirlin, Physical Review B 100, 125160 (2019).
  • Guo et al. (2010) H.-M. Guo, G. Rosenberg, G. Refael, and M. Franz, Phys. Rev. Lett. 105, 216601 (2010).
  • Guo (2010) H.-M. Guo, Phys. Rev. B 82, 115122 (2010).
  • Shapourian and Hughes (2016) H. Shapourian and T. L. Hughes, Phys. Rev. B 93, 075108 (2016).
  • Chen et al. (2015) C.-Z. Chen, J.-T. Song, Q.-F. S. H. Jiang, Z.Wang, and X. Xie, Phys. Rev. Lett. 115, 246603 (2015).
  • Louvet et al. (2016) T. Louvet, D. Carpentier, and A. A. Fedorenko, Physical Review B 94, 220201 (2016).
  • Ryu and Nomura (2012) S. Ryu and K. Nomura, Physical Review B 85, 155138 (2012).
  • Qi et al. (2006) X.-L. Qi, Y.-S. Wu, and S.-C. Zhang, Phys. Rev. B 74, 085308 (2006).
  • Prodan et al. (2010) E. Prodan, T. L. Hughes, and B. A. Bernevig, Phys. Rev. Lett. 105, 115501 (2010), URL https://link.aps.org/doi/10.1103/PhysRevLett.105.115501.
  • Prodan (2011) E. Prodan, Journal of Physics A: Mathematical and Theoretical 44, 113001 (2011), URL https://dx.doi.org/10.1088/1751-8113/44/11/113001.
  • Groth et al. (2009) C. W. Groth, M. Wimmer, A. R. Akhmerov, J. Tworzydło, and C. W. J. Beenakker, Phys. Rev. Lett. 103, 196805 (2009), URL https://link.aps.org/doi/10.1103/PhysRevLett.103.196805.
  • Hua et al. (2019) C.-B. Hua, R. Chen, D.-H. Xu, and B. Zhou, Phys. Rev. B 100, 205302 (2019), URL https://link.aps.org/doi/10.1103/PhysRevB.100.205302.