1: \documentclass[twocolumn,showpacs,preprintnumbers,amsmath,amssymb,aps,prb,superscriptaddress,floatfix]{revtex4}
2: %\documentclass[showpacs,preprintnumbers,amsmath,amssymb,aps,prb,superscriptaddress]{revtex4}
3: \usepackage{graphicx}
4: \usepackage[latin1]{inputenc}
5:
6: \begin{document}
7: \title{Spin nematic phases in models of correlated electron systems: a numerical study}
8: \author{S.\ Capponi} \affiliation{Laboratoire de
9: Physique Th\'eorique, CNRS UMR 5152, Universit\'e Paul Sabatier, F-31062 Toulouse, France.}
10: \author{F.~F.~Assaad} \affiliation{Institut f\"ur Theoretische Physik und Astrophysik,
11: Universit\"at W\"urzburg, Am Hubland, D-97074 W\"urzburg, Germany}
12:
13: \date{\today}
14: \pacs{{71.27.+a}{71.10.Fd}}
15:
16: \begin{abstract}
17: Strongly interacting systems are known to often
18: spontaneously develop exotic ground states under certain
19: conditions. For instance, spin nematic phases have been discovered in various magnetic models. Such
20: phases, which break spin symmetry but have no net local magnetization, have also been proposed by Nersesyan {\it et al.}~\cite{Nersesyan91} in the
21: context of electronic models. We introduce a $N$-flavor microscopic model that interpolates from the large-$N$ limit, where mean-field is valid and
22: such a nematic phase occurs, to the more
23: realistic $N=1$ case. By using a sign-free quantum Monte-Carlo, we show the existence of a spin nematic phase (analogous to a spin flux phase) for
24: finite $N$; when $N$ decreases, quantum fluctuations increase and this phase ultimately disappears in favor of an s-wave superconducting state. We also show that
25: this nematic phase extends up to a finite critical charge doping.
26:
27: Dynamical studies allow us to clarify the Fermi surface property~: in the nematic phase at half-filling, it consists of four points and the low-energy
28: structure has a Dirac cone-like shape. Under doping, we observe clear signatures of Fermi pockets around these points.
29:
30: This is one of the few examples where numerical simulations show how quantum fluctuations can destroy a large-$N$ phase.
31: \end{abstract}
32:
33: \maketitle
34: %**********************************
35: \section{Introduction}
36: %**********************************
37:
38: Spin nematic phases are characterized by an absence of local magnetic moment and
39: long range orientational ordering of the spin degrees of freedom. Such phases were first introduced by
40: Andreev and Grishchuk~\cite{Andreev84} and then discovered in several magnetic models~\cite{Papanicolaou88,Chandra90}.
41: Recently, spin nematic phases have also
42: attracted attention in the the context of cold atoms where a gas
43: bose condensate of $^{52}Cr$ atoms in magnetic trap has been realized~\cite{Griesmaier05}.
44: Due to the high spin, $J=3$, of the $^{52}Cr$ atoms, it has been suggested that,
45: upon release of the spin degrees of freedom, spin nematic states can be realized~\cite{Pfau06,Diener06}.
46: Mean-field realization of spin-nematic phases in particle-hole symmetric
47: fermionic models on a square lattice have equally been proposed by
48: Nersesyan {\it et al.}~\cite{Nersesyan91}. Those phases are characterized by a checkerboard
49: pattern of {\it alternating spin currents} around elementary plaquettes, and are coined
50: spin flux phases (SFP). They break $SU(2)$ spin symmetry as well as the lattice symmetry,
51: but conserve time reversal symmetry. Such a phase could also possibly occur in an extended Hubbard model~\cite{Schulz89}. The goal of this
52: paper is to provide a \emph{microscopic} realization of SFP.
53:
54: In this article, we concentrate on fermionic models on a square lattice and investigate
55: the stability of the spin-flux phase. To do so, we consider the multi-flavored model
56: Hamiltonian:
57: \begin{eqnarray}
58: \label{model.eq}
59: H & = & - t \sum_{\langle i,j\rangle, \alpha} \left( c^{\dagger}_{i,\alpha} c_{j,\alpha} + h.c. \right) \nonumber \\
60: &- & \frac{g}{2N} \sum_{\langle i,j\rangle }
61: \left( \sum_{\alpha} i c^{\dagger}_{i,\alpha} \vec{\sigma} c_{j,\alpha}
62: -i c^{\dagger}_{j,\alpha} \vec{\sigma} c_{i,\alpha} \right)^2 .
63: \end{eqnarray}
64: Here, the sum runs over nearest neighbors on a two-dimensional square lattice. The spinors
65: $ c^{\dagger}_{i,\alpha} = \left( c^{\dagger}_{i,\uparrow,\alpha},c^{\dagger}_{i,\downarrow,\alpha} \right)$
66: where $\alpha$ ranges from $1$ to $N$.
67: In the limit $N \rightarrow \infty$, the saddle point approximation becomes exact,
68: and we recover the mean-field results of Ref.~\onlinecite{Nersesyan91}.
69: As $N$ is reduced, quantum
70: fluctuations around this saddle point are progressively taken in account and ultimately at
71: $N=1$ we recover a spin-1/2 fermionic Hamiltonian. As we will see, the model,
72: of Eq. (\ref{model.eq}) allows sign free auxiliary field quantum Monte Carlo (QMC) simulations for
73: arbitrary band fillings and values of $ N$. Hence, we can map out the phase diagram as a function of doping for various $N$,
74: at fixed choice of coupling constant $g/t$. In the following, we fix $t=1$ as the energy unit.
75:
76: The article is organized as follows. In the next section, we briefly describe the
77: path integral formulation of the partition function which is the basis of the mean field and
78: quantum Monte Carlo simulations. In section~\ref{Results} we present our numerical results which
79: allow us to map out the phase diagram as a function of $N$ and doping for a fixed coupling strength.
80: Finally, we summarize our results in section~\ref{Conclusions}.
81:
82:
83: %**********************************
84: \section{Mean-field and quantum Monte Carlo simulations. }
85: %**********************************
86:
87: Both the saddle point equations and the QMC simulations are based on a path
88: integral formulation of the partition function. Carrying out a Trotter breakup of the
89: kinetic and
90: interaction terms as well as a Hubbard-Stratonovitch transformation for the interaction term,
91: the partition function reads:
92: \begin{widetext}
93: \begin{eqnarray}
94: & & Z \simeq \int
95: \underbrace{\prod_{\langle i, j \rangle, \tau } {\rm d} \vec{\Phi}_{\langle i, j \rangle}(\tau) }_{ \equiv D \vec{\Phi} }
96: e^{- \sum_{\langle i, j \rangle, \tau } \vec{\Phi}_{\langle i, j \rangle}^2(\tau) /2 }
97: \, {\rm Tr } \left[ \prod_{\tau} e^{ t \Delta \tau \sum_{\langle i,j\rangle, \alpha} \left( c^{\dagger}_{i,\alpha} c_{j,\alpha} + h.c. \right) } \right. \nonumber \\
98: & & \left. e^{\displaystyle \sqrt{ \Delta \tau g/N} \sum_{\langle i,j\rangle } \vec{\Phi}_{\langle i, j \rangle}(\tau)\cdot
99: \left( \sum_{\alpha} i c^{\dagger}_{i,\alpha} \vec{\sigma} c_{j,\alpha}
100: -i c^{\dagger}_{j,\alpha} \vec{\sigma} c_{i,\alpha} \right)
101: } \right] = \nonumber \\
102: & &
103: \int D \vec{\Phi} e^{- \sum_{\langle i, j \rangle, \tau } \vec{\Phi}_{\langle i, j \rangle}^2(\tau) /2 }
104: \, {\rm Tr } \left[ \prod_{\tau} e^{ t \Delta \tau \sum_{\langle i,j\rangle} \left( c^{\dagger}_{i} c_{j} + h.c. \right) } e^{ \sqrt{ \Delta \tau g/N} \sum_{\langle i,j\rangle } \vec{\Phi}_{\langle i, j \rangle}(\tau)\cdot
105: \left( i c^{\dagger}_{i} \vec{\sigma} c_{j}
106: -i c^{\dagger}_{j} \vec{\sigma} c_{i} \right) } \right]^{N}
107: \end{eqnarray}
108: \end{widetext}
109: where
110: $ c^{\dagger}_{i} = \left( c^{\dagger}_{i,\uparrow }, c^{\dagger}_{i,\downarrow} \right) $.
111: Note that in the last equation, the trace runs over a single flavor.
112:
113:
114: As is well known, in the large $N$ limit the mean-field solution is recovered.
115: With the substitution
116: \begin{equation}
117: \vec{\eta}_{\langle i, j \rangle}(\tau) = \frac{1}{\sqrt{N \Delta \tau g} }
118: \vec{\Phi}_{\langle i, j \rangle}(\tau)
119: \end{equation}
120: one obtains
121: \begin{widetext}
122: \begin{eqnarray}
123: \label{HS.eq}
124: & & Z \simeq
125: \int D \vec{\eta} e^{- N S(\eta) } \\
126: & & S(\eta) =
127: \sum_{\langle i, j \rangle, \tau } \Delta \tau g
128: \vec{\eta}^2_{\langle i, j \rangle}(\tau) /2 - \ln
129: {\rm Tr } \prod_{\tau} e^{ t \Delta \tau \sum_{\langle i,j\rangle} \left( c^{\dagger}_{i} c_{j} + h.c. \right) }
130: e^{ \Delta \tau g \sum_{\langle i,j\rangle } \vec{\eta}_{\langle i, j \rangle}(\tau)\cdot
131: \left( i c^{\dagger}_{i} \vec{\sigma} c_{j}
132: -i c^{\dagger}_{j} \vec{\sigma} c_{i} \right) } \nonumber
133: \end{eqnarray}
134: \end{widetext}
135:
136: In this large $N$ limit, the integral over the the fields $\eta$ is dominated by the saddle point configuration and fluctuations around the saddle point are negligible. Moreover,
137: by neglecting the $\tau$ dependence of the Hubbard-Stratonovitch fields, one obtains~:
138: \begin{eqnarray}
139: & & S(\eta) = \beta g \sum_{\langle i,j\rangle}
140: \vec{\eta}_{\langle i, j \rangle}^2 /2 - \ln
141: {\rm Tr } e^{-\beta H_{MF}(\eta) } \nonumber \\
142: & & H_{MF}(\eta) = - t \sum_{\langle i,j\rangle} \left( c^{\dagger}_{i} c_{j} + h.c. \right)\nonumber \\
143: & & - g \sum_{\langle i,j\rangle } \vec{\eta}_{\langle i, j \rangle}\cdot
144: \left( i c^{\dagger}_{i} \vec{\sigma} c_{j}
145: -i c^{\dagger}_{j} \vec{\sigma} c_{i} \right)
146: \end{eqnarray}
147: and the saddle-point equations read~:
148: \begin{equation}
149: \frac{\partial S(\eta) } { \partial \vec{\eta}_{\langle i,j \rangle}} = 0
150: \end{equation}
151: which correspond precisely to the mean-field solution~\cite{Nersesyan91}
152: \begin{equation}
153: \vec{\eta}_{\langle i,j \rangle} = \langle i c^{\dagger}_{i} \, \vec{\sigma} \, c_{j}
154: -i c^{\dagger}_{j} \, \vec{\sigma} \, c_{i} \rangle_{H_{MF}}
155: \end{equation}
156:
157:
158: \subsection{Staggered spin flux mean field solution}
159: We consider the restricted order parameter:
160: \begin{equation}
161: \vec{\eta}_{\langle i,j \rangle} = \eta (-1)^{i_x + i_y}
162: ( \delta_{j-i,a_x} - \delta_{j-i,a_y} )\, \vec{e}_z.
163: \end{equation}
164: This choice is certainly valid at half-band filling where perfect nesting pins the
165: dominant instabilities to the wave vector $\vec{Q} = (\pi,\pi) $. Away from half-filling,
166: order parameters at incommensurate wave vectors may be favorable. This is not taken into
167: account in the present Ansatz.
168: With the above form, the mean-field Hamiltonian reads:
169: \begin{widetext}
170: \begin{equation}
171: H_{MF} = \sum_{\vec{k} \in MBZ, \sigma}
172: \left( c^{\dagger}_{\vec{k},\sigma}, c^{\dagger}_{\vec{k} + \vec{Q},\sigma} \right)
173: \left(
174: \begin{array}{cc}
175: \varepsilon(\vec{k}) - \mu & -\sigma g \eta \overline{z(\vec{k})} \\
176: -\sigma g \eta z(\vec{k}) & \varepsilon(\vec{k} + \vec{Q}) - \mu
177: \end{array}
178: \right)
179: \left(
180: \begin{array}{c}
181: c_{\vec{k},\sigma} \\ c_{\vec{k} + \vec{Q},\sigma}
182: \end{array}
183: \right)
184: \end{equation}
185: \end{widetext}
186: where the sum over $\vec{k}$ is restricted to the magnetic Brillouin zone (MBZ) and
187: with
188: \begin{equation}
189: z({\vec{k}}) = -2i \left( \sin ( k_x + Q_x/2 ) -
190: \sin ( k_y + Q_y/2 ) \right)
191: \end{equation}
192: Thus, we obtain the saddle point equation
193: \begin{eqnarray}
194: \eta = & & \frac{1}{2 L^2} \sum_{\vec{k} \in MBZ, \sigma}
195: \langle
196: \left( c^{\dagger}_{\vec{k},\sigma}, c^{\dagger}_{\vec{k} + \vec{Q},\sigma} \right)
197: \nonumber \\
198: & & \left(
199: \begin{array}{cc}
200: 0 & \sigma \overline{z(\vec{k})} \\
201: \sigma z(\vec{k}) & 0
202: \end{array}
203: \right)
204: \left(
205: \begin{array}{c}
206: c_{\vec{k},\sigma} \\ c_{\vec{k} + \vec{Q},\sigma}
207: \end{array}
208: \right)
209: \rangle
210: \end{eqnarray}
211: which we solve self-consistently.
212:
213: \begin{figure}[!ht]
214: \includegraphics[width=7cm]{MF}
215: \caption{(Color online) Mean Field order parameter as a function of band-filling for various coupling strengths $g$. }
216: \label{MF.fig}
217: \end{figure}
218:
219: Figure~\ref{MF.fig} shows the order parameter as a function of doping. The staggered spin flux
220: phase survives up to a finite critical doping, where a first order transition to the paramagnetic
221: phase occurs. This transition is signaled by a jump
222: in the order parameter. At half-band filling, the single
223: particle dispersion relation is given by,
224: \begin{equation}
225: E(\vec{k}) = \pm \sqrt{ \varepsilon^2(\vec{k}) + \Delta^2(\vec{k}) }
226: \end{equation}
227: with $\Delta(\vec{k}) = 2 g \eta \left( \cos(k_x) - \cos(k_y) \right)$. Hence, it exhibits
228: Dirac cones around the $( \pm \pi/2, \pm \pi/2)$ $k$-points and the Fermi {\it surface}
229: is given by those four points. In the doping range where the order parameter
230: does not vanish the Fermi surface consists of hole-pockets centered around the
231: above mentioned $\vec{k}$-points (See Fig.~\ref{NkMF.fig}). Note that
232: due to the d-wave symmetry of the order parameter, the Fermi surface is not invariant
233: under reflections across the $(1,1)$ axis
234: (i.e. $k_x \rightarrow k_y $ and $k_y \rightarrow k_x $).
235: As appropriate for a first order transition, the Fermi surface changes abruptly from hole
236: pockets around $(\pi/2,\pi/2)$ in the spin flux phase to a
237: large Fermi surface centered around $ (0,0) $ in the paramagnetic phase.
238:
239: \begin{figure}[!ht]
240: \includegraphics[width=8cm]{Nk_074}
241:
242: \includegraphics[width=8cm]{Nk_075}
243:
244: \includegraphics[width=8cm]{Nk_09}
245:
246: \includegraphics[width=8cm]{Nk_098}
247: \caption{(Color online) Mean-field solution of $n(\vec{k})$ over the Brillouin zone as a function of band filling for a fixed $g=1$.}
248: \label{NkMF.fig}
249: \end{figure}
250:
251: %**********************************
252: \subsection{Quantum Monte-Carlo}
253: %*********************************
254:
255: The Hamiltonian occurring in~(\ref{HS.eq}) is quadratic in the fermionic variables so that
256: the trace can be carried out analytically to obtain:
257: \begin{equation}
258: Z \simeq \int D \vec{\Phi}
259: e^{- \sum_{\langle i, j \rangle, \tau } \vec{\Phi}_{\langle i, j \rangle}^2(\tau) /2 }
260: \left[ {\rm det} M(\Phi)\right]^N
261: \end{equation}
262: Since the spin current is even under time reversal symmetry, it has been shown that,
263: when $g\geq 0$, the
264: fermionic determinant is positive for each Hubbard-Stratonovitch configuration~\cite{Wu04}.
265: This absence of sign problem is valid for any lattice topology and any filling.
266: Hence each configuration of Hubbard Stratonovitch fields
267: can be sampled according to its weight with Monte Carlo techniques.
268:
269:
270: In the following, we study the ground-state properties of the Hamiltonian~(\ref{model.eq}) with
271: the projector
272: auxiliary field QMC algorithm on a two-dimensional square lattice. The details on this algorithm can be found for example in
273: Ref.~\onlinecite{Capponi00},
274: where the authors consider a very similar model from the technical point of view. Dynamical
275: information is obtained by using a recent implementation of the stochastic analytical
276: continuation~\cite{Beach04a,Sandvik98}.
277:
278:
279:
280: \section{Numerical results}\label{Results}
281: \subsection{Phase diagram}
282: For $N=1$, the interaction term of our model can be rewritten, up to a constant as:
283: \begin{eqnarray}
284: \label{Interaction_N1}
285: &-& \frac{g}{2} \sum_{\langle i,j\rangle }
286: \left( \sum_{\alpha} i c^{\dagger}_{i,\alpha} \vec{\sigma} c_{j,\alpha}
287: -i c^{\dagger}_{j,\alpha} \vec{\sigma} c_{i,\alpha} \right)^2 \nonumber \\
288: =
289: &-&3g\sum_{\langle i,j\rangle } (c^{\dagger}_{i,\uparrow}c^{\dagger}_{i,\downarrow}c^{\phantom{\dagger}}_{j,\downarrow} c^{\phantom{\dagger}}_{j,\uparrow}
290: + h.c.) \nonumber \\
291: &+&\frac{3}{2}g \sum_{\langle i,j\rangle } (n_i-1)(n_j-1) \nonumber \\
292: &-&2g\sum_{\langle i,j\rangle } \vec{S}_i \cdot \vec{S}_j
293: \end{eqnarray}
294: As apparent, there is a pair-hopping term so that we can expect an on-site superconducting
295: instability which we pick up by measuring
296: the equal time pair correlation functions:
297: \begin{equation}
298: SC(\vec{R})=\langle c^\dagger_{\vec{R}\uparrow} c^\dagger_{\vec{R}\downarrow} c^{\phantom{\dagger}}_{0\downarrow} c^{\phantom{\dagger}}_{0\uparrow} \rangle.
299: \end{equation}
300: In the presence of long range off-diagonal order, its Fourier transform at $\vec{Q}=(0,0)$
301: \begin{equation}\label{SC_Q}
302: SC(\vec{Q}=(0,0))=\frac{1}{L^2}\sum_{\vec{R}} SC(\vec{R})
303: \end{equation}
304: should converge to a finite value in the thermodynamic limit, whereas it vanishes as $1/L^2$ in the absence of long-range order.
305: \begin{figure}[h]
306: \includegraphics[width=8cm,clip]{supra_d0v2}
307: \caption{(Color online) Finite-size scaling of the superconducting correlations vs the inverse number of sites $1/L^2$ of the system
308: ($L$ is the linear size) for the
309: half-filled case and different $N$. The $\vec{Q}=(0,0)$ Fourier correlations are plotted (open symbols) and, when long range order is present,
310: we also show the largest distance real-space correlations (filled symbols).}
311: \label{supra_d0.fig}
312: \end{figure}
313:
314: The order parameter of the spin flux phase expected in the limit of $N \rightarrow \infty $,
315: reads:
316: \begin{eqnarray}\label{sfp.eq}
317: \langle \left[ \vec{J}^s(\vec{R},\vec{R}+\vec{x})-\vec{J}^s(\vec{R},\vec{R}+\vec{y}) \right]
318: \nonumber \\
319: \cdot \left[ \vec{J}^s(\vec{0},\vec{0}+\vec{x})-\vec{J}^s(\vec{0},\vec{0}+\vec{y}) \right]
320: \rangle
321: \end{eqnarray}
322: where
323: \begin{equation}
324: \vec{J}^s(\vec{R},\vec{R}+\vec{x})= i \left[
325: c^\dagger_{\vec{R},\mu} \vec{\sigma}_{\mu \nu} c^{\phantom{\dagger}}_{\vec{R}+\vec{x},\nu}
326: -
327: c^\dagger_{\vec{R}+\vec{x},\mu} \vec{\sigma}_{\mu \nu} c^{\phantom{\dagger}}_{\vec{R},\nu}
328: \right]
329: \end{equation}
330: is the bond spin current operator along $x$ (summation over repeated indices is assumed). A similar definition holds along the $y$ direction.
331: As expected from the mean-field solution, we have observed that the strongest signal occurs in
332: the d-wave symmetry channel and the Fourier transform of these correlations is maximum at
333: $\vec{Q}=(\pi,\pi)$.
334: \begin{figure}
335: \includegraphics[width=8cm]{jspind2_d0}
336: \caption{(Color online) Similar as the previous plot for $d$-wave spin current correlations.}
337: \label{jspind_d0.fig}
338: \end{figure}
339:
340: \begin{figure}
341: \includegraphics[width=0.4\textwidth,]{eta_n1}
342: \caption{Order parameter $\eta$ as a function of $N$ at half-filling. The data point at $N= \infty$ corresponds to
343: the mean-field value. }
344: \label{eta_N}
345: \end{figure}
346:
347:
348: \begin{figure}
349: \includegraphics[width=8cm]{sz2}
350: \caption{(Color online) Finite-size scaling of the spin-spin correlations vs the inverse
351: volume, $1/L^2$, of the system for the
352: half-filled case and different $N$. For all cases, the $\vec{Q}=(\pi,\pi)$ Fourier correlations (which is the largest one) vanish in the thermodynamic limit, indicating the absence
353: of magnetic order.}
354: \label{sz.fig}
355: \end{figure}
356:
357:
358: {\it Technical details:} we are
359: able to simulate up to $20\times 20$ square lattices with
360: a projection parameter $\theta t =10$. We fix $\Delta \tau=0.1$ for the
361: Trotter decomposition. For simplicity, in the following, we fix the coupling constant $g=1$.
362:
363: {\it Half-filling case~:}
364: On Fig.~\ref{supra_d0.fig}, we plot the scaling of the superconducting correlations for different values of $N$. According to our definition~(\ref{SC_Q}),
365: the $\vec{Q}=(0,0)$
366: correlations are divided by the number of sites so that a finite value in the thermodynamic limit indicates long range order.
367: Moreover, when this is the case, we also plot the largest distance real-space correlations that should converge to the same
368: value in the thermodynamic limit. In particular, we have a clear signal of
369: s-wave superconducting phase for $N=1$ and $N=2$ but this phase disappears for larger $N$.
370:
371: At $N=1$ and under the canonical transformation,
372: \begin{equation}
373: c^{\dagger}_{i,\downarrow} \rightarrow (-1)^i c_{i,\downarrow} \; \; \; \;
374: c^{\dagger}_{i,\uparrow } \rightarrow (-1)^i c^{\dagger}_{i,\uparrow}
375: \end{equation}
376: the interaction term transforms as,
377: \begin{eqnarray}
378: - \frac{g}{2}
379: \left( i c^{\dagger}_{i} \vec{\sigma} c_{j}
380: -i c^{\dagger}_{j} \vec{\sigma} c_{i} \right)^2 \rightarrow
381: 6g \vec{S}_i \cdot \vec{S}_j & & \nonumber \\
382: - \frac{g}{2} (n_i -1) (n_j -1) +g ( c^{\dagger}_{i,\uparrow} c^{\dagger}_{i,\downarrow} c^{\phantom{\dagger}}_{j,\downarrow}
383: c^{\phantom{\dagger}}_{j,\uparrow} + h.c.),
384: \end{eqnarray}
385: and the kinetic energy remains invariant. The observed long-ranged
386: pairing correlations map onto long ranged tranverse spin-spin correlations. Since the
387: transformed model has an $SU(2)$ spin symmetry, the transverse spin-spin correlations on
388: any finite lattice take the same value as the longitudinal spin-spin correlations.
389: Transforming back to the original model maps the longitudinal spin-spin correlations to
390: a charge density wave order. Hence at half-filling and $N=1$ pairing correlations and
391: charge density wave correlations are locked in by symmetry.
392: Clearly, doping breaks this symmetry~\footnote{Precisely the same transformation has
393: been used to
394: investigate the half-filled attractive Hubbard model~\cite{Scalettar89}.}.
395:
396:
397: Similarly, we investigate the occurrence of SFP phase by computing the
398: corresponding correlations (see Eq.~(\ref{sfp.eq})). Moreover, for a finite system, the $SU(2)$
399: symmetry of these correlations cannot be broken so that we average
400: over the 3 components to reduce the error bars. For example, we plot on Fig.~\ref{jspind_doping.fig}~(b) all three components of the
401: SFP correlations showing that indeed $SU(2)$ symmetry is restored (within the error bars) despite being
402: explicitely broken for a given Hubbard-Stratonovitch configuration.
403: On Fig.~\ref{jspind_d0.fig}, we plot the scaling of these correlations
404: showing that SFP is present in the thermodynamic limit for $N\ge 3$ but is not stable for smaller
405: values of $N$. Again, when the $\vec{Q}=(\pi,\pi)$ Fourier component
406: indicates long range order (LRO), we also plot the scaling of the largest distance
407: correlations that should have the same value in the thermodynamic
408: limit for an ordered phase. In particular, we recover the existence
409: of SFP phase for large $N$ as found at the mean-field level. A non
410: trivial result is that this phase survives for \emph{finite} $N\ge 3$ since
411: our QMC simulations, which are free of the sign problem, include
412: quantum fluctuations.
413: We observe rather strong finite size effects and in particular, the signal is weaker when
414: the lattice contains
415: $(\pi/2,\pi/2)$ k-points in its Brillouin zone. Indeed, as we will discuss in the
416: following, this point correspond to the low-energy excitations.
417: Since the staggered spin current correlation function of Fig.~\ref{jspind_d0.fig}
418: converges to the square of the
419: order parameter $\eta$, we can extract this quantity as a function of $N$. Our results
420: are plotted in Fig.~\ref{eta_N}. As apparent, the Monte Carlo results at finite values of $N$
421: smoothly scale to the mean field result valid at $N = \infty$.
422:
423:
424: In the SFP at $N \geq 3$, we have computed the spin-spin correlation functions
425: $\langle \vec{S}_{ \vec{i}} \cdot \vec{S}_{ \vec{j}} \rangle$.
426: As shown on Fig.~\ref{sz.fig}, our results show no sign of long-range spin ordering and hence the absence of a magnetic moment.
427: This confirms the point of
428: view that the spin-flux phase that we observe for $ N \geq 3 $ indeed corresponds to a
429: \emph{spin nematic phase}.
430:
431: \begin{figure}
432: \includegraphics[width=8cm]{jspind_N10L12}
433: \includegraphics[width=8cm]{jspind2_N3L10}
434: \caption{(Color online) (a) Spin current correlations, averaged over the 3 components,
435: vs ${\vec{k}}$ along the $(\Gamma X M)$ path in the Brillouin zone.
436: Parameters are $L=12$, $N=10$ and $g=1$. As a function of doping $\delta$ (increasing from top to bottom), the maximum at ($\pi,\pi$) is rapidly suppressed.
437: (b) Same for $L=10$ and $N=3$. Here, we plot all 3 components showing that $SU(2)$ symmetry is recovered in our data.
438: As a function of doping (increasing from top to bottom),
439: the maximum at ($\pi,\pi$) is rapidly suppressed.}
440: \label{jspind_doping.fig}
441: \end{figure}
442:
443: \begin{figure}
444: \includegraphics[width=8cm]{phasediag}
445: \caption{(Color online) Phase diagram as a function of doping and $1/N$ obtained with QMC simulations. At large
446: $N$, we recover the existence of a spin flux phase (SFP) over a finite doping region. For larger doping or
447: smaller $N$, the system develops s-wave superconducting correlations (SC) or stays in a metallic phase (see text).}
448: \label{phasediag.fig}
449: \end{figure}
450:
451:
452: {\it Doping~:}
453: Since the SFP instability occurs in the particle-hole channel, it is favored at half-filling
454: where the Fermi surface exhibits perfect nesting. But the question
455: about its existence under doping remains open.
456: Our QMC simulation being free of the sign-problem for any filling,
457: we have computed the phase diagram as a function of doping for different values of $N$.
458: As expected from the mean-field results, we show on Fig.~\ref{phasediag.fig} that the SFP
459: phase has a finite extension up to a critical doping. In particular, when $N=10$, we find
460: a critical doping around 0.15.
461:
462: On Fig.~\ref{jspind_doping.fig}, we plot the spin current correlations vs ${\vec{k}}$ along the $(\Gamma X M)$ path in the Brillouin zone (shown in
463: Fig.~\ref{brillouin.fig}).
464: For doping $\delta$ smaller than a critical value, we see a large maximum at ($\pi,\pi$), whereas for larger doping, the correlations become featureless,
465: indicating the disappearance of SFP order. We also observe that the maximum of the spin-current correlations move away from $(\pi,\pi)$:
466: for example, when $L=12$, $N=10$ and with doping $\sim 0.17$, the maximum is at $(5\pi/6,\pi)$ and equivalent k-points.
467:
468: By performing a finite-size scaling analysis of our data, we are able to draw conclusion
469: about the existence or not of a SFP in the thermodynamic limit
470: as a function of doping and $N$. These results are summarized in the phase diagram
471: shown in Fig.~\ref{phasediag.fig}. Several comments are in order~: \\
472: {\bf i)} Under doping the SFP only survives for large values of $N$, hence showing that it
473: is very sensitive to fluctuations. At $N=10$ the
474: critical doping, $\delta_c \simeq 0.15 $ at which the SFP vanishes is
475: substantially smaller than the $N \rightarrow \infty $ restricted mean-field result
476: $\delta_c \simeq 0.25$. In the vicinity of the phase transition from the SFP to the
477: paramagnetic phase transition at $N \rightarrow \infty$, fluctuations will play an important
478: role. Hence at large but finite values of $N$, substantial differences can be expected.
479: Furthermore, our restricted mean-field solution does not allow for incommensurate
480: ordering as suggested by the Monte Carlo results. \\
481: {\bf ii)} The rest of the phase diagram is dominated by an s-wave SC phase. At
482: $N=1$, the s-wave SC persists of course away from half-filling since it
483: is a particle-particle instability that does not require any nesting property.
484: As $N$ grows, the s-wave SC order parameter is reduced and ultimately vanishes to
485: produce the paramagnetic phase in the $N \rightarrow \infty $ limit.
486: At $N=10$ and $\delta > \delta_c \simeq 0.15$ we observe no SFP correlations and no
487: SC correlations either. Therefore, it seems that we might have a metallic phase with
488: a Fermi surface. However, due to the presence of pair hopping in the model, we know that
489: this metallic phase will ultimately be unstable at very low temperature toward s-wave
490: superconductivity. This instability probably occurs at too low temperature to be observed
491: in our simulations. To confirm this point of view, let us write the interaction term as:
492: \begin{eqnarray}
493: -\frac{g}{2N} \sum_{b} \left( \sum_{\alpha=1}^{N} \vec{J}^{s}_{b,\alpha}\right)^2 =
494: -\frac{g}{2} \sum_{b} \sum_{\beta=1}^{N} \nonumber \\
495: \left[ \left( \frac{1}{N}\sum_{\alpha\neq \beta} \vec{J}^{s}_{b,\alpha} \right)
496: \vec{J}^{s}_{b,\beta } +
497: \frac{1}{N} \vec{J}^{s}_{b,\beta} \cdot \vec{J}^{s}_{b,\beta}\right]
498: \end{eqnarray}
499: where $b=\langle \vec{i},\vec{j}\rangle $ denotes a nearest neighbor bond, and
500: $ \vec{J}^{s}_{b,\alpha} $ is the spin current on bond $b$ for flavor index $\alpha$.
501: This forms shows that in the large-N limit, the flavor index $\beta$ is embeded in the
502: {\it mean field } spin current of the other flavors,
503: $ \frac{1}{N}\sum_{\alpha\neq \beta} \vec{J}^{s}_{b,\alpha} $.
504: The second term, $ \frac{1}{N} \vec{J}^{s}_{b,\beta} \cdot \vec{J}^{s}_{b,\beta}$ is a
505: factor $1/N$ smaller than the {\it mean-field } term and provides the pair hopping term as
506: explicitly shown in Eq.~(\ref{Interaction_N1}). At $N = \infty $, only the mean-field
507: term survives. For large but finite values of $N$, a small pairing term is present and
508: will in the paramagnetic phase trigger a superconducting instability at low temperatures. \\
509: {\bf iii)} The coexistence a spin flux phase and superconductivity at finite doping
510: cannot be excluded. As we will see below, doping the SFP leads to a Fermi
511: surface consisting of hole pockets centered around the
512: $(\pm \pi/2,\pm \pi/2)$ points in the Brillouin zone. Due to the presence of pair
513: hopping, this Fermi surface should be unstable towards a superconducting state. As a result,
514: for intermediate values of $N$ such that we observe SFP at zero doping and such that the superconducting
515: signal at finite doping is not strongly reduced due to a $1/N$ factor, we have found numerical evidence that supports coexistence of both phases
516: for small doping (for instance, $N=4$ and $\delta\sim 0.06$ in the phase diagram).
517:
518: \subsection{Dynamical properties}
519:
520: To study the charge degrees of freedom, we compute the single particle spectral function
521: $A(\vec{k},\omega)$ which is related to the imaginary time Green function via:
522: \begin{equation}
523: \langle c^{\dagger}_{\vec{k},\alpha}(\tau) c_{\vec{k},\alpha} \rangle
524: = \frac{1}{\pi}
525: \int_{0}^{\infty} {\rm d} \omega e^{-\tau \omega} A(\vec{k}, -\omega).
526: \label{AKOM}
527: \end{equation}
528:
529: In order to plot this single particle spectral function, we follow a path in the Brillouin
530: zone shown on Fig.~\ref{brillouin.fig}.
531: \begin{figure}
532: \includegraphics[width=4cm]{brillouin}
533: \caption{In the following spectral function, from bottom to top, the $k$ value follows the path $\Gamma$, $X$, $M$ and then, along the $(\pi,0)$ to $(0,\pi)$ line.}
534: \label{brillouin.fig}
535: \end{figure}
536: In Fig.~\ref{Akw_SFP_d0.fig} we first plot two examples of
537: $A(\vec{k},\omega)$ in the SFP at half-band filling. As expected from the mean-field
538: calculation we observe a clear sign of the Dirac cones around the $ (\pi/2,\pi/2)$
539: ${\vec{k}}$-points in the Brillouin zone. The overall dispersion relation compares favorably
540: to the mean-field result
541: \begin{equation}\label{meanfield_dispersion.eq}
542: E(\vec{k}) = \pm \sqrt{ \varepsilon^2(\vec{k}) + \Delta^2(\vec{k}) }
543: \end{equation}
544: with $\Delta(\vec{k}) = 2 g \eta \left( \cos(k_x) - \cos(k_y) \right)$.
545:
546: \begin{figure}
547: \includegraphics[width=8cm]{Akw_N10_L12_d0}
548:
549: \includegraphics[width=8cm]{Akw_N4_L16_d0}
550:
551: \caption{(Color online) Spectral functions vs frequency $\omega$ for different $\vec{k}$ points at half-filling. Up: $N=10$ and $L=12$;
552: bottom: $N=4$ and $L=16$. Both set of parameters correspond to SFP and Dirac cones
553: around $(\pi/2,\pi/2)$ are clearly seen. }
554: \label{Akw_SFP_d0.fig}
555:
556: \end{figure}
557:
558: In order to be more quantitative, we plot the dispersion relation of the main branch (by looking at the maximum of
559: the spectral function) as a function of the distance to the nodal point. On Fig.~\ref{dirac.fig}, we clearly see the Dirac cone
560: structure with different velocities parallel and perpendicular to the Fermi surface. By fitting the slopes, we obtain for this set
561: of parameters, $v_\perp=3.11$ and $v_\parallel=0.578$. At the mean-field level, according to Eq.~(\ref{meanfield_dispersion.eq}),
562: these two velocities are equal respectively to $2\sqrt{2}$ and $2\sqrt{2}g\eta$ so that their ratio gives directly access to the SFP order parameter $\eta$ (when
563: $g=1$). Of course, the QMC data include some renormalization of these values and
564: the ratio of $v_\parallel / v_\perp =0.186$ is inconsistent with the SFP
565: order parameter ( $ \eta \simeq 0.45$) as extracted from the spin current
566: correlations (see Fig.~\ref{jspind_d0.fig}).
567:
568: \begin{figure}
569: \includegraphics[width=8cm]{dirac}
570:
571: \caption{(Color online) Dispersion relation obtained with $N=4$ and $L=16$ at half-filling. The energy is taken at the maximum of the spectral function and data
572: are plotted vs distance to $(\pi/2,\pi/2)$. The small asymmetry is due to the uncertainty in the spectral functions obtained with
573: a Maximum Entropy technique. }
574: \label{dirac.fig}
575: \end{figure}
576:
577:
578: In the mean-field approach, doping the SFP leads to hole pockets centered around the
579: $(\pi/2,\pi/2)$ $\vec{k}$-points. Numerical QMC simulations confirms this as shown on Fig.~\ref{Akw_SFP_d0_12.fig}. At $N=10$ and
580: $\delta = 0.12$ we observe a clear signature of the Dirac cone, however the quasiparticle
581: at $\vec{k} = (\pi/2,\pi/2) $ lies above the Fermi energy.
582:
583: \begin{figure}
584: \includegraphics[width=8cm]{Akw_N10_L12_d0_12}
585: \caption{(Color online) Spectral functions vs $\omega-\mu$ for different $k$ points. Parameters are: $N=10$, $L=12$
586: and doping $0.12$. It corresponds to a doped SFP phase and Dirac cones around $(\pi/2,\pi/2)$ are clearly seen. The
587: chemical potential is $\mu=-0.48$.}
588: \label{Akw_SFP_d0_12.fig}
589: \end{figure}
590:
591: \begin{figure}
592: \includegraphics[width=8cm]{Akw_N3_L12_d0_12}
593:
594: \includegraphics[width=8cm]{Akw_N2_L16_d0}
595:
596: \caption{(Color online) Spectral functions vs frequency $\omega$ for different $k$ points.
597: Top: $N=3$ and $L=12$ for doping $0.12$ ($\mu=-0.7$).
598: Bottom $N=2$ and $L=16$ at half-filling ($\mu=0$).
599: Both set of parameters correspond to s-wave superconducting phase. }
600: \label{Akw_SC.fig}
601: \end{figure}
602:
603: Deep in the superconducting phase, the single particle dispersion relation follows the
604: the mean-field form:
605: \begin{equation}
606: E(\vec{k}) = \pm \sqrt{ ( \epsilon(k) - \mu)^2 + \Delta_{sc}^2}
607: \end{equation}
608: with $k$-independent superconducting gap. Such a dispersion relation is consistent with
609: the Monte Carlo data of Fig. \ref{Akw_SC.fig} at $ N=3 $ and $\delta = 0.12$. In the
610: proximity of the SFP at $N=2$ and half-band filling (see Fig. \ref{Akw_SC.fig})
611: the dispersion is not captured by the above form. In particular, a modulation of the
612: gap along the $(\pi,0)$ to $(0,\pi)$ line is observed. This modulation is reminiscent of the
613: Dirac cone structure observed in the spin-flux phase, and hence allows the interpretation
614: that short range spin currents survive in the superconducting state in the proximity of the
615: SFP.
616:
617: %**********************************
618: \section{Conclusion}
619: %*********************************
620: \label{Conclusions}
621: We have shown the existence of a spin nematic phase in a two-dimensional electronic model. This
622: phase had been proposed at the mean-field level and it is the aim of this article to
623: investigate its stability against
624: quantum fluctuations. In order to do so, we have performed unbiased
625: QMC simulations (without sign problem) for a variety
626: of models with a flavor parameter $N$. With this trick, we can interpolate from the large $N$
627: limit where the mean-field is valid to finite $N$ regime where quantum fluctuations around the
628: saddle point are progressively taken into account.
629: We find that the spin-flux phase is stable for a range of finite $N$ but is ultimately destroyed
630: for the more realistic $N=1$ case. This is one of the few examples
631: where QMC simulations are able to show how quantum fluctuations can destroy a large-$N$ phase~\cite{Assaad04}.
632: Since our model includes pair hopping processes,
633: when the spin-flux phase is found to be unstable, it is replaced by an on-site $s$-wave superconductivity.
634:
635: We have also investigated the phase diagram as a function of doping since our QMC simulations
636: are free of the sign problem for any filling. Because the spin-flux
637: phase is due to the nesting property of the Fermi surface, it is expected to disappear with doping.
638: This is indeed what we found but still, this phase can
639: persist over a finite range of doping, similarly to what is found at the mean-field level.
640:
641: Using a recently developed analytical continuation technique, we have computed dynamical
642: spectral functions. In the SFP phase, we clearly see Dirac cones forming
643: around $(\pi/2,\pi/2)$ and equivalent $k$-points, where the dispersion becomes linear.
644: Up to a critical doping, these structures are stable so that the Fermi surface
645: becomes pocket-like.
646:
647:
648: \begin{acknowledgments}
649: We would like to thank C. Wu and S.-C. Zhang for motivating this work and useful discussion.
650: Financial support from the
651: Bayerisch-Franz\"osisches Hochschulzentrum / Centre de Coop\'eration Universitaire Franco-Bavarois
652: is acknowledge. S.~C. thanks IDRIS
653: (Orsay, France) for use of supercomputer facilities.
654: \end{acknowledgments}
655:
656: \begin{thebibliography}{10}
657:
658: \bibitem{Nersesyan91}
659: A.~A. Nersesyan, G.~I. Japaridze, and I.~G. Kimeridze, J. Phys.: Cond.
660: Matt. {\bf 3}, 3353 (1991).
661:
662: \bibitem{Andreev84}
663: A.~F. Andreev and I.~A. Grishchuk, Sov. Phys. JETP {\bf 60}, 267 (1984).
664:
665: \bibitem{Papanicolaou88}
666: N. Papanicolaou, Nucl. Phys. B {\bf 305}, 367 (1988).
667:
668: \bibitem{Chandra90}
669: P. Chandra, P. Coleman, and A.~I. Larkin, J. Phys.: Cond. Matt. {\bf 2},
670: 7933 (1990).
671:
672: \bibitem{Griesmaier05}
673: A. Griesmaier, J. Werner, S. Hensler, J. Stuhler, and T. Pfau, Phys. Rev. Lett.
674: {\bf 94}, 160401 (2005).
675:
676: \bibitem{Pfau06}
677: L. Santos and T. Pfau, Phys. Rev. Lett. {\bf 96}, 190404 (2006).
678:
679: \bibitem{Diener06}
680: R.~B. Diener and T.-L. Ho, Phys. Rev. Lett. {\bf 96}, 190405 (2006).
681:
682: \bibitem{Schulz89}
683: H.~J. Schulz, Phys. Rev. B {\bf 39}, 2940 (1989).
684:
685: \bibitem{Wu04}
686: C. Wu and S.-C. Zhang, Phys. Rev. B {\bf 71}, 155115 (2005).
687:
688: \bibitem{Capponi00}
689: S. Capponi and F.~F. Assaad, Phys. Rev. B {\bf 63}, 155114 (2001).
690:
691: \bibitem{Beach04a}
692: K.~S.~D. Beach, cond-mat/0403055 (2004).
693:
694: \bibitem{Sandvik98}
695: A. Sandvik, Phys. Rev. B {\bf 57}, 10287 (1998).
696:
697: \bibitem{Assaad04}
698: F.~F. Assaad, Phys. Rev. B {\bf 71}, 075103 (2005).
699:
700: \bibitem{Scalettar89}
701: R.~T. Scalettar, E.~Y. Loh, J.~E. Gubernatis, A. Moreo, S.~R. White, D.~J.
702: Scalapino, and R.~L. Sugar, Phys. Rev. Lett. {\bf 62}, 1407 (1989).
703:
704: \end{thebibliography}
705:
706: \end{document}
707: