0710.0030/tmi.tex
1: \documentclass[prl,aps,amssymb,showpacs,twocolumn]{revtex4}
2: 
3: \usepackage{dcolumn}
4: \usepackage{graphicx}
5: \usepackage{bm}
6: 
7: 
8: \usepackage{amsmath}
9: \usepackage{amssymb}
10: \usepackage{amsthm}
11: \usepackage{amsfonts}
12: \usepackage{enumerate}
13: \usepackage{latexsym}
14: 
15: \input{epsf}
16: 
17: 
18: \begin{document}
19: \title{Topological Mott Insulators}
20: \author{S. Raghu$^{1}$, Xiao-Liang Qi$^1$, C. Honerkamp$^2$, and Shou-Cheng Zhang$^1$}
21: \affiliation{$^1$Department of Physics, McCullough Building,
22: Stanford University, Stanford, CA 94305-4045}
23: \affiliation{$^2$Theoretical Physics, Universit\"{a}t W\"{u}rzburg, D-97074 W\"{u}rzburg, Germany}
24: \date{\today}
25: \begin{abstract}
26: We consider extended Hubbard models with repulsive interactions on
27: a Honeycomb lattice and the transitions from the semi-metal phase
28: at half-filling to Mott insulating phases.  In particular, due to
29: the frustrating nature of the second-neighbor repulsive
30: interactions, topological Mott phases displaying the quantum Hall
31: and the quantum spin Hall effects are found for spinless and
32: spinful fermion models, respectively.  We present the mean-field
33: phase diagram and consider the effects of fluctuations within the
34: random phase approximation (RPA). Functional renormalization group
35: analysis also show that these states can be favored over the
36: topologically trivial Mott insulating states. 
37: \end{abstract}
38: 
39: \pacs{71.10.-w, 71.10.Fd, 71.27.+a, 71.30.+h, 73.22.Gk, 73.43.-f}
40: 
41: \maketitle \emph{Introduction - } Partly motivated by the
42: discovery of the high $T_c$ superconductivity, Mott insulators
43: have attracted great attention in recent years. Defined in a
44: general sense, interactions drive a quantum phase transition from
45: a metallic ground state to an insulating ground state in these systems. Most Mott
46: insulators found in nature also have conventional order
47: parameters, describing, for example, the charge-density-wave (CDW)
48: or the spin-density-wave (SDW) orders. However, Mott insulators
49: with exotic ground states, such as the current carrying ground
50: states have also been proposed theoretically
51: \cite{Affleck1988,Wen1996,Varma1999,Chakravarty2001}.
52: In parallel with the study of strongly correlated systems, there
53: has recently been a growing interest in realizing topologically
54: non-trivial states of matter in band insulators. In the quantum
55: anomalous Hall (QAH) insulators\cite{Haldane1988,Qi2006}, the
56: ground state breaks time reversal symmetry but does not break the
57: lattice translational symmetry. The ground state has a bulk
58: insulating gap, but has chiral edge states. In the quantum spin
59: Hall (QSH) insulators\cite{kane2005,bernevig2006a,bernevig2006d},
60: the ground state does not break time reversal symmetry, has a
61: bulk insulating gap, but has helical edge states, where electrons
62: with the opposite spins counter-propagate. The QSH state has
63: recently been observed experimentally in $HgTe$ quantum
64: wells\cite{bernevig2006d,koenig2007}.
65: 
66: Given the tremendous interest in finding Mott insulators with
67: exotic ground states, and the recent discovery of the
68: topologically non-trivial band insulators, it is natural to ask
69: whether one can find examples of topological Mott insulators,
70: which we define as states with bulk insulating gaps driven by the
71: interaction, and inside which lie topologically protected edge
72: states. Furthermore, electronic states in the Mott insulator
73: phases are characterized by topological invariants, namely, the
74: $U(1)$ Chern number\cite{Thouless1982} in the case of the QAH
75: state, and the $Z_2$ invariant\cite{kane2005A} in the case of the
76: QSH state.  In this letter, we report on the first example of such a 
77: case by systematically studying Hubbard models with repulsive
78: interactions on a two dimensional honeycomb lattice.  In
79: particular, we demonstrate how, due to the frustrated nature of
80: second-neighbor repulsion on this lattice, topological Mott phases
81: displaying the QAH and the QSH effects are generated dynamically.
82: We present the mean-field phase diagram, proceed to consider
83: effects beyond mean-field theory via a functional renormalization
84: group (fRG) treatment, and consider the effects arising from
85: fluctuations.  The two dimensional graphene sheet, which was
86: recently realized experimentally  \cite{Novoselov2005,Zhang2005}, 
87: contains the basic degrees of freedom of
88: our model. However, presently we do not know how to tune the
89: interaction experimentally so that our proposed state can be
90: realized.
91: 
92: \emph{Spinless Fermions and the QAH state - }
93: \begin{figure}
94: \includegraphics[width=3.0in]{geom2.eps}
95: \caption{Interactions considered in our model Hamiltonian
96: (left-most hexagon), Eq. \ref{hamiltonian}. Various order
97: parameters are shown for the A-sublattice (open circles) in the
98: middle hexagon and for the B-sublattice (filled circles) in the
99: right-most hexagon.  The CDW order parameter $\rho$ is a real
100: scalar, SDW order parameter $\bm M$ a real vector and the QAH/QSH
101: order parameters $\chi_A, \chi_B$ are complex 4-vectors.  In the
102: case of spinless fermions, $\chi_A, \chi_B$ are complex scalars.  $\chi_A,\chi_B$ are both defined
103: on the directed second neighbor links defined by $\bm b_i$.    }
104: \label{geometry}
105: \end{figure}
106: The model Hamiltonian for spinless fermions with nearest-neighbor
107: and next-nearest neighbor interactions is  written
108: as\begin{eqnarray} \label{hamiltonian} H=-\sum_{\langle ij \rangle}t
109: \left(c_i^\dagger c_j+h.c. \right)+V_1 \sum_{\langle i,j \rangle }
110: (n_{i}-1)( n_{j}-1) \nonumber  \\ + V_2\sum_{\langle \langle i,j
111: \rangle \rangle}  (n_{i}-1) (n_{j}-1) -\mu\left( \sum_in_i-N \right)
112: \end{eqnarray}
113: where $V_1 $ and $V_2$ are nearest-neighbor and
114: next-nearest-neighbor interaction strengths, respectively.  Since
115: the honeycomb lattice is bipartite, consisting of two triangular
116: sublattices (referred to here as A and B), nearest-neighbor
117: repulsion will favor a charge density wave (CDW) phase with an 
118: order parameter $\rho = \frac{1}{2} \left(\langle c^{\dagger}_{i
119: A} c_{i  A} \rangle - \langle c^{\dagger}_{i  B} c_{i  B} \rangle
120: \right)$ that is consistent with overall charge conservation and
121: describes a phase with a broken discrete (inversion) symmetry.
122: However, since the second neighbor interactions within a
123: sublattice are frustrated, CDW order will be suppressed; instead
124: we consider the possibility of orbital ordering by defining the
125: following order parameter for $i,j$ next nearest neighbors:
126: $\chi_{ij}=\chi_{ji}^* = \langle{c_i^\dagger c_j} \rangle$.  Let
127: $\bm a_1, \bm a_2, \bm a_3$ be the nearest-neighbor displacements
128: from a B-site to an A-site such that $\bm z \cdot \bm a_1 \times
129: \bm a_2$ is positive.   We also define the displacements $\bm b_1
130: = \bm a_2 - \bm a_3$, $ \bm b_2 = \bm a_3 - \bm a_1$, etc, which
131: connect two neighboring sites on the same sublattice (Fig.
132: \ref{geometry}).  A translational and rotational invariant ansatz
133: of $\chi_{ij}$ is chosen as
134: \begin{eqnarray} \chi_{i,i+{\bf
135: b}_s}=\left\{\begin{array}{c c}\chi_A = \vert \chi \vert
136: e^{i\phi_A},&i\in A\\ \chi_B = \vert \chi \vert e^{i\phi_B},&i\in
137: B\end{array}\right.
138: \end{eqnarray}
139: which are \emph{complex} scalars that live along the directed second
140: neighbor links.
141: 
142: Owing to translational
143: symmetry, the mean-field free energy at $T=0$ is readily obtained: 
144: \begin{eqnarray}
145: F \left( \rho, \chi, \bar{\phi}, \phi \right) &=& -  \sum_{\bm k}
146: \sqrt{  \vert  t(\bm k) \vert ^2 + \left( V_1 \rho +2 V_2 \vert \chi
147: \vert S_{\bm k  + \bar{\phi} }   S_{\phi} \right)^2 } \nonumber
148: \\
149: &&+ 3L^2 \left( V_1  \rho^2 + 2V_2 \vert \chi \vert ^2 \right)
150: \end{eqnarray}
151: We have defined the following quantities: $t(\bm k) = \sum_{n=1}^3
152: \exp \left( i \bm k \cdot a_n \right)$, $ \bar{\phi} = \left( \phi_A
153: + \phi_B \right)/2, \phi = \left(\phi_A - \phi_B \right)/2,S_{\bm k+
154: \bar{\phi}}=\sum_{n=1}^3 \sin {\left(\bm k\cdot b_n + \bar{\phi}
155: \right)}, S_{\phi} = \sin{\phi} $.
156: Thus, the next-neighbor hopping
157: amplitudes are purely real only when both $\phi = 0$ and $\bar{\phi}
158: = 0$.
159: 
160: When both $\rho$ and $\chi= 0$, and at half-filling, the system
161: remains a semi-metal consisting of two Fermi points $\bm K_{\pm}$
162: which obey $\bm K_{\pm} \cdot \bm b_i = \pm 2 \pi/3$ and the density
163: of states vanishes linearly; the dispersion in the vicinity of these
164: so called Dirac points is governed by a 2D massless Dirac
165: Hamiltonian in $\bm k$-space . The CDW phase corresponds to an
166: ordinary insulator with a gap at the Fermi energy. As for the order
167: parameter $\chi$, which describes the second-neighbor hopping, its
168: \emph{phase} relative to the nearest neighbor hopping amplitude
169: plays an important role in determining its properties: while a
170: non-zero $Re(\chi)$ merely shifts the energy of the Dirac points, a
171: non-zero imaginary part $Im(\chi)$ \emph{opens a gap} at the Fermi
172: points.  Thus, when the system remains at half-filling, it is more
173: favorable to develop  purely imaginary next-neighbor hopping
174: amplitudes; such a configuration corresponds to a phase with
175: \emph{spontaneously broken time-reversal symmetry}.
176: 
177: To see whether such a phase can be favored, we minimize the
178: free-energy and arrive at the following self-consitent
179: equations:\begin{equation} \rho = \frac{1}{2 L^2} \sum_{\bm k}
180: \frac{V_1 \rho + 2V_2\chi S_{\bm k + \bar \phi} S_{\phi}}{\sqrt{
181: \vert t(\bm k) \vert^2
182:  +\left(V_1 \rho + 2V_2\chi S_{\bm k + \bar{\phi} } S_{\phi}\right)^2 } }
183: \end{equation}
184: \begin{equation}
185: \chi = \frac{S_{\phi}}{6L^2} \sum_{\bm k} \frac{S_{\bm k +
186: \bar{\phi} }   \left(V_1 \rho + 2V_2 \chi S_{\bm k + \bar \phi }
187: S_{\phi} \right)}{\sqrt{ \vert t(\bm k) \vert^2 +\left(V_1 \rho +
188: 2V_2 \chi S_{\bm k  + \bar{\phi} } S_{\phi}\right)^2 }
189: }
190: \end{equation} When $\chi = 0$, it is easy to see from the first
191: equation above that CDW order develops continuously at a critical
192: value $V_{1c}$ given by
193: \begin{equation}
194: \frac{1}{V_{1c}} = \frac{1}{2L^2} \sum_{\bm k} \frac{1}{\vert t(\bm k) \vert},
195: \end{equation}
196: Due to the vanishing density of states (DOS) near the Fermi points,
197: there is no instability towards CDW formation with infinitesimal
198: interactions.  Interestingly, the self-consitent equation for
199: $\chi$ shows that a non-trivial self-consistent solution can only
200: occur when $\phi \ne 0$; a detailed investigation of these
201: equations \cite{Raghu2007} show that when $V_1 = 0$, beyond a
202: critical value of $V_{2c}>0$, which satisfies\begin{equation}
203: \label{gapeqQAHE} \frac{1}{V_{2c}} = \frac{1}{3L^2} \sum_{\bm
204: k}\frac{S^2_{\bm k + \bar{\phi}} }{\vert t_{\bm k} \vert},
205: \end{equation}
206: a phase in which $\vert \chi \vert > 0$ ,$\bar{\phi} = 0$, and
207: $\phi = \pm \pi/2$ is favored.  Such a phase is also stable at
208: finite $V_1$ and is thus does not require fine-tuning (see Fig.
209: \ref{pd2}). In this phase, the system acquires purely imaginary
210: second-neighbor hoppings with a chirality which is determined by
211: the sign of $\phi$. There is a discrete symmetry breaking
212: corresponding to choosing $\phi = \pm \pi/2$, each of which breaks
213: time-reversal symmetry. The band insulator version of the CDW
214: state was considered in Ref. \cite{Semenoff1984}, while the QH
215: state on a honeycomb lattice was considered in Ref.
216: \cite{Haldane1988}. As discussed in Ref. \cite{Haldane1988} the
217: QAH phase is a topological phase in which the filled states form a
218: band which has a non-zero topological Chern number
219: \cite{Thouless1982} and is an integer quantum Hall effect phase
220: that is realized \emph{without} Landau levels. Similar states can
221: be constructed from magnetic semiconductors\cite{Qi2006}. We shall
222: generally refer to QH states without Landau levels, and with full
223: lattice translation symmetry as the \emph{Quantum Anomalous Hall}
224: (QAH) states. In our case, the topologically non-trivial gap for
225: the QAH state arises from many-body interactions rather than
226: single particle physics, and we shall refer to such states as
227: topological Mott insulators.
228: \begin{figure}
229: \includegraphics[width=2.5in]{PD2.eps}
230: \caption{Phase diagram for spinless fermions ($t=1$).  The
231: semi-metallic (SM) state that occurs at weak-coupling is separated
232: from the CDW and the topological QAH states via a continuous
233: transition (blue curve). The line separating the QAH and CDW marks a
234: first-order transition (shown in red), which terminates at a bi-critical point.}
235: \label{pd2}
236: \end{figure}
237: 
238: We have obtained the complete  phase diagram in the $V_1-V_2$
239: plane; within mean-field theory, there is a continuous transition
240: from the semi-metal to either the CDW or the QAH phase and there
241: is also a first-order transition from the CDW to the QAH phase. By
242: integrating out the fermionic fields, it is possible to construct
243: a Landau-Ginzburg (LG) theory expansion near the nodal region
244: where the order parameters are vanishingly small.  Due to the
245: linear dispersion of the Fermi points, the LG free-energy contains
246: anomalous terms of the form  $\vert \rho \vert^3$ and $\vert
247: Im(\chi) \vert^3$ arising from linear dispersion in the vicinity
248: of the Fermi points \cite{Raghu2007}.  The significance of such
249: terms is that even within mean-field theory, the CDW order
250: parameter, for instance, grows as $(V_1 - V_{1c})$ rather than the
251: usual $\left( V_1 - V_{1c} \right)^{1/2}$ \cite{Sorella1992}.
252: Furthermore, the Landau-Ginzburg theory describing the competition
253: between the CDW and QAH phases confirms the existence of the first
254: order line between these two phases that terminates in the
255: bicritical point leading into the semimetal phase
256: \cite{Raghu2007}.
257: 
258: \emph{Spinful fermions and the QSH state -} Next, we take into
259: account the spin degrees of freedom and include an onsite Hubbard
260: repulstion in our model Hamiltonian ($\mu = 0$):
261: \begin{eqnarray}
262: H=-\sum_{\langle ij \rangle  \sigma}t\left(c_{i \sigma}^\dagger c_{j
263: \sigma}+h.c. \right)+U\sum_i n_{i \uparrow} n_{i \downarrow}
264: \nonumber \\ + V_1\sum_{\langle i,j \rangle }   (n_{i}-1)( n_{j}-1)
265: + V_2\sum_{\langle \langle i,j \rangle \rangle}  (n_{i}-1)(
266: n_{j}-1)\label{hamilton}
267: \end{eqnarray}
268: where $n_i = n_{i \uparrow} + n_{i \downarrow}$.  Since the
269: Honeycomb lattice is bipartite, onsite repulsion gives rise to a
270: spin density wave phase (SDW); a standard decomposition of the Hubbard term 
271: introduces the order parameter $\bm M$ describing an
272: antiferromagnetic SDW: $\bm M = \frac{1}{2} \left( \langle \bm
273: S_{iA} \rangle-\langle \bm S_{iB} \rangle \right)$ .  As in the
274: spinless case, nearest-neighbor repulsion favors a CDW.  However,
275: there are several possible phases due to second-neighbor
276: repulsion.   Again, since the second-neighbor repulsion is
277: frustrated, we are again led to the possibility of a topological
278: phase similar to the QAH.  However, the spin degrees of freedom
279: introduce two possibilities (translation invariance along with
280: spin conservation eliminate other possibilities): 1) \emph{two
281: copies} of QAH states - i.e. the chirality of the
282: second-neighbor hopping is the \emph{same} for each spin projection, 2)
283: the QSH state, where the chiralities are \emph{opposite} for each
284: spin projection.   The latter possibility breaks a
285: \emph{continuous} global $SU(2)$ symmetry associated with choosing
286: the spin projection axis;  however, \emph{time-reversal symmetry
287: is preserved}. The QSH state on the honeycomb lattice was
288: considered in Ref. \cite{kane2005}, where the insulating gap
289: arises from the microscopic spin-orbit coupling. It was shown
290: latter that the magnitude of of the spin-orbit gap is negligibly small
291: in graphene \cite{Min2006,Yao2007}. In our case, the insulating gap is generated
292: dynamically from the many-body interaction. In this sense, our
293: effect can be viewed as an example of dynamic generation of
294: spin-orbit interaction\cite{Wu2004}.
295:   Introducing the Hubbard-Strataovich fields (sum over repeated indices implied)
296: $\chi^{\mu}_{ij} = c^{\dagger}_{i \alpha} \sigma^{\mu}_{\alpha
297: \beta} c_{j \beta}$, $\mu = 0 \ldots 3$, where $\sigma^{\mu} =
298: \left( 1, \bm \sigma \right)$, the next-neighbor interactions can
299: be recast using the identity $ (n_{i } -1)(n_{j}-1) =1  -
300: \frac{1}{2} \left( \chi^{\mu }_{ij} \right)^{\dagger}
301: \chi^{\mu}_{ij} $. Physically, if $\langle \chi^0 \rangle \neq 0$,
302: then we are in the QAH phase.  If, on the other hand, one of the
303: vector components $\langle \chi^{i} \rangle \neq 0$, then we are
304: in the QSH phase.  A translationally invariant decomposition of
305: the next-neighbor interactions via $ \langle \chi^{\mu}_{i,i+\bm
306: b_s} \rangle = \chi^{\mu}e^{i \phi^{\mu}_A} , i \in A $ (and
307: similarly for the other sublattice) gives rise to 
308: a $4 \times 4$ Hamiltonian is readily diagonalized in
309: a tensor product basis $\bm \sigma \otimes \bm \tau$, where $\bm
310: \sigma$ and $\bm \tau$ are Pauli matrices in spin and sublattice
311: space, respectively. This way, each phase corresponds to a
312: particular non-zero expectation value of a fermion bilinear
313: $\sum_{\vec{k}}\Psi_{\vec{k}}^\dagger \hat{d}(\vec{k})
314: \Psi_{\vec{k}}$, where $\hat{d}(\vec{k}) \propto \tau^3$ for the
315: CDW and QAH, $\hat{d}(\vec{k}) \propto \sigma^3\tau^3$ for SDW and
316: QSH.  A detailed study of the free-energy at $T=0$ and its saddle
317: point solutions \cite{Raghu2007} produces the phase diagram shown
318: in Fig. \ref{pd3}.  In addition to the ordinary CDW and SDW
319: insulating phases, there is a phase for $V_2 > V_{2c} \approx 1.2
320: t$ in which the 4-vector is purely imaginary (as in the spinless
321: case), collinear, and staggered from one sublattice to the next:
322: $\langle \chi^{\mu}_{i i+\bm b_n, A} \rangle = - \langle
323: \chi^{\mu}_{i i+\bm b_n, B} \rangle$, and both QAH and QSH are
324: equally favorable ground states, having identical free energies
325: within mean-field theory.  Additionally, there is never a
326: coexistence of both QAH and QSH phases; indeed, a Landau-Ginzburg
327: treatment in this region explicitly shows the absence $SO(4)$
328: symmetry of the vector $\chi^{\mu}$.  This occurs due to the
329: difference of the manner in which $\chi^0$ and $\vec{\chi}$ are
330: coupled to the fermionic fields - which favors either a phase with
331: broken $Z_2$ symmetry (QAH) or with broken $SU(2)$ symmetry, but
332: never both simultaneously \cite{Raghu2007}.
333: 
334: Quantum fluctuations, however, lift the mean-field degeneracy
335: between the QAH and QSH phases.  To quadratic order in quantum
336: fluctuations (RPA) about the QSH phase , we obtain an effective
337: action $ S_{eff} = \sum_{\vec{k}} \delta \chi^{\mu}(\vec{k},
338: \Omega) K_{\mu \nu} (\vec{k}, \Omega ) \delta \chi^{\nu}
339: (-\vec{k}, -\Omega)$ which shows the presence of six modes (2
340: longitudinal and 4 transverse modes), and 2 of the transverse
341: modes correspond to degenerate Goldstone modes whose velocity is
342: proportionality to the Fermi velocity $v \approx v_f = 3t/2\vert
343: \bm a \vert $.  Thus, the zero-point motion associated with these
344: gapless modes, lowers the free energy of the QSH state relative to
345: the QAH state.
346: 
347: \begin{figure}
348: \includegraphics[width=3.0in]{PD3.eps}
349: \caption{Complete mean-field phase diagram for the spinful model.  The transitions from the semimetal (SM) to the insulating phases are continuous, whereas transitions between any two insulating phases (red lines) are first-order.  
350: } \label{pd3}
351: \end{figure}
352: 
353: 
354: \emph{Renormalization Group Analysis - } Mean field
355: theory generally starts with a given, in a sense, biased Ansatz,
356: and investigate the self-consistency of the mean field solution.
357: Therefore, it is important to investigate the topological Mott
358: states with a method without any a priori bias. Next we go beyond
359: mean-field theory and RPA using the temperature($T$)-flow
360: functional renormalization group
361: (fRG)\cite{Honerkamp2001}\cite{Honerkamp2001A}.  
362: In this scheme, we discretize the $\vec{k}$-dependence of the
363: interaction \cite{Zanchi2000}
364: and consider all possible scattering processes between a set of initial
365: and final momenta that occur between
366: points on rings around the Dirac points (inset of Fig. \ref{rgfig}).
367: Starting with $T_0 \sim 2t$, the
368: temperature $T$ is lowered, and a flowing (renormalized) interaction
369: $V_T$ is obtained by the
370: coupled summation of the $T$-derivatives of {\em all} one-loop channels.
371: Due to this, the method is
372: unbiased and goes beyond the mean-field-level.  
373: Applying the scheme to the Hamiltonian, Eq.
374: \ref{hamilton}, we search for {\em
375:   flows to strong coupling}, where for a low temperature $T_c$ certain
376: components of $V_T $ become large. Then the approximations break
377: down, and the flow is stopped. Information on the low-$T$ state is
378: obtained from analyzing which coupling functions grow most
379: strongly and from susceptibilities for static external fields
380: coupling to the various order parameters.  In this scheme, a
381: tendency towards ordering at a finite vector $\bm Q$ can be
382: detected as a growth of the associated vertex $V_T$.  However, we
383: have found that largest couplings occur at $\bm Q = 0$, which
384: strongly supports the mean-field results presented above.
385: 
386: For onsite and nearest-neighbor repulsions $U>U_c \approx 3.8t$
387: and $V_1> V_{1c} \approx 1.2t$, the flow to strong coupling is
388: either an SDW instability for dominant $U$ or a CDW instability
389: for dominant $V_1$, in good agreement with a
390: 1/N-study\cite{Herbut2006} and
391: Quantum-Monte-Carlo\cite{Sorella1992}. For more details, see Ref.
392: \cite{Honerkamp2007}. If we include a sufficiently strong
393: second-nearest-neighbor repulsion $V_2>1.6t$, the flows change
394: qualitatively; there is a leading growth of the QSH
395: susceptibility. In Fig. \ref{rgfig} a) and b) we compare the
396: $T$-flows of various susceptibilities for $V_1>V_2$ and for
397: $V_2>V_1$. For the latter case, the QSH susceptibility grows most
398: strongly toward low $T$, followed by the QAH susceptibility, which
399: is consistent with the RPA treatment of the Goldstone modes in the
400: QSH. The QSH phase remains stable even when a moderate onsite
401: interaction of $U=t$ or $U=2t$ is introduced. Hence the global
402: structure of the mean-field phase diagram is confirmed by the fRG
403: results. Note however that the slope of the lines of critical
404: $V_1$ versus $V_2$ differs. We interpret this a competition effect
405: captured by the fRG, where $V_2$ decreases the CDW tendencies
406: induced by $V_1$.
407: 
408: 
409: \begin{figure}
410: \includegraphics[width=3.5in]{rgfig.eps}
411: \caption{{\em a)} Data for $U$=0, $V_1$=1.4$t$, $V_2$=0.  Susceptibilities of each phase vs. $T$ are shown: CDW (black);  SDW (green); QAH(red) and QSH (blue).{\em b)} Same for $U$=0, $V_1$=0, $V_2$=1.8$t$ (QSH instability).  The QSH phase has a larger susceptibility than QAH.   {\it Inset}: fRG phase diagram at $U$=0, indicating SM (blue) and insulating (red) regions (CDW dominates at large $V_1$, QSH at large $V_2$.)   The colorbar correspond to $T_c$ below which the insulating phases develop in fRG.}
412: \label{rgfig}
413: \end{figure}
414: 
415: \emph{Discussion - } We have shown that topological phases
416: displaying the QAH and QSH effects can be generated from strong
417: interactions - thus, we refer to these phases as \emph{topological
418: Mott insulators}. Both phases have associated with them
419: conventional order parameters which develop continuously at the
420: quantum critical phase transition from the semi-metallic state.
421: However, these states are also described by topological quantum
422: numbers which jump discontinuously at the transition. Although
423: the interaction strengths needed to produce these phases are
424: strong, we expect that our mean-field, RPA and fRG treatment to
425: provide strong evidence for the existence of a topological Mott
426: insulator: the Stoner criterion, which states that perturbation theory 
427: breaks down when interaction strengths are comparable to the 
428: inverse of the effective DOS, suggests that perturbation theory 
429: is more robust in our system due to the vanishing of the 
430: DOS at the Fermi points.  
431: An open issue remains to find other possible realizations
432: of the phases described here. Furthermore, the nature of the low energy
433: effective coupling of the gapless bulk Goldstone modes with the gapless
434: \emph{edge} degrees of freedom, and the possibility of
435: fractionalized excitations in these phases are interesting
436: open issues that we leave for future work.
437: 
438: \section{Acknowledgments}
439: We are grateful to D. P. Arovas and S. A. Kivelson for insightful discussions.   
440: This work is supported by the NSF under grant numbers DMR-0342832, 
441: the US Department of Energy, Office of Basic Energy Sciences
442: under contract DE-AC03-76SF00515, BaCaTec (C.H.),
443: and the Stanford Institute for Theoretical Physics (S.R.).    
444: 
445: 
446: 
447: \bibliography{tmi}
448: 
449: \end{document}
450: