cond-mat0605550/fp.tex
1: %
2: %	REVTeX 4
3: %	Global Settings
4: % twocolumn,showpacs,preprint,eqsecnum,aps,eqsecnum,aps,draft,prb
5: %
6: \documentclass[prl,twocolumn,showpacs,superscriptaddress,floatfix]{revtex4}
7: %\documentclass[prl,preprint,showpacs,superscriptaddress,floatfix]{revtex4}
8: 
9: % 
10: %	Packages
11: %
12: \usepackage{amsmath}			% rozsirenie math-prostredia
13: \usepackage{amssymb}			% rozsirenie math-fontov
14: \usepackage{graphicx}			% input obrazkov
15: \usepackage{epic}
16:   \graphicspath{{./Eps/}}
17: %\usepackage{dcolumn}			% zarovnavanie v tabulkach
18: 					% na des. bodku
19: 
20: %
21: %	Bibliography style path
22: %
23: %\bibliographystyle{d:/Pavel/Misc/BibTeX/MyBibStyle}
24: 
25: %
26: %	My macro
27: %
28: \newcommand{\ua}{$\uparrow$}
29: \newcommand{\da}{$\downarrow$}
30: \newcommand{\singlet}{\ua \da}
31: \newcommand{\triplet}{\ua \ua}
32: \newcommand{\tlt}{\tilde t}
33: 
34: \newcommand{\myvec}[3]{ 
35: \left(
36:   \begin{array}{c}
37:    #1 \\
38:    #2 \\
39:    #3
40:   \end{array}
41: \right)
42: }
43: 
44: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
45: %	document begin
46: %
47: \begin{document}
48: 
49: \title{Fermion nodes and nodal cells of noninteracting and interacting fermions
50: }
51: 
52: \author{Lubos \surname{Mitas}}
53: %\email{lmitas@ncsu.edu}
54: \affiliation{Department of Physics, North Carolina State University, Raleigh,
55: NC 27695}
56: 
57: 
58: \date{\today}
59: 
60: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
61: %	Abstract
62: %
63: \begin{abstract}
64: A fermion node is subset of fermionic configurations for which a real wave 
65: function vanishes due to the antisymmetry and the node divides the configurations
66: space into compact nodal cells (domains). We analyze the properties of fermion 
67: nodes of fermionic ground state wave functions for a number of systems. For several
68: models we demonstrate that noninteracting spin-polarized fermions in dimension two 
69: and higher have closed-shell ground state wave functions with the minimal two nodal
70: cells for any system size and we formulate a theorem which sumarizes this result. 
71: The models include periodic fermion gas, fermions on the surface of a sphere, 
72: fermions in a box. We prove the same property for atomic states with up to $3d$ 
73: half-filled shells. Under rather general assumptions we then derive that the same 
74: is true for unpolarized systems with arbitrarily weak interactions using 
75: Bardeen-Cooper-Schrieffer (BCS) variational wave function. We further show that pair 
76: correlations included in the BCS wave function enable singlet pairs of particles 
77: to wind around the periodic box without crossing the node pointing towards the 
78: relationship of nodes to transport and many-body phases such as superconductivity.
79: Finally, we point out that the arguments extend also to fermionic temperature
80: dependent/imaginary-time density matrices. The results reveal fundamental properties 
81: of fermion nodal structures and provide new insights for accurate constructions of 
82: wave functions and density matrices in quantum and path integral Monte Carlo methods.
83: \end{abstract}
84: 
85: \pacs{02.70.Ss, 03.65.Ge}
86: %\pacs{61.46.+w, 73.22.-f, 36.40.Cg}
87: %\keywords{Suggested keywords}
88: 
89: \maketitle
90: 
91: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
92: %
93: %\section{Introduction}
94: \section{\label{sec:level1} Introduction}
95: %The properties of fermionic many-body systems belong to the most
96: %fascinating challenges in condensed matter quantum physics.
97: Let us consider a system of fermions described 
98: by a real wave function $\Psi(R)$ where $R$ denotes 
99: fermions coordinates 
100:  $R=({\bf r}_1, ...,{\bf r}_N)$. Due to the antisymmetry, there
101: exists a subset of fermion configurations for which the 
102: wave function vanishes and
103: this subset is called the fermion node.
104: The fermion node can be  
105: implicitly defined by 
106: $\Psi(R)=0$, assuming that the nodal set does not include
107: configurations for which the wave function vanishes because
108: of other reasons, eg,
109: boundary conditions.
110: In general, for $N$  spin-polarized fermions in  
111:   a $d$-dimensional space,  the fermion node is
112:  a $(dN-1)-$dimensional manifold
113: (hypersurface). 
114: It is a well-known fact that for $d>1$ the
115: antisymmetry alone does not specify 
116: fermion nodes completely. This is not difficult to understand
117: since antisymmetry
118: fixes only lower-dimensional 
119: coincidence hyperplanes with dimensionalities $(dN-d)$.
120: Therefore the fermion nodes and their properties are determined   
121: by interactions and many-body effects.
122: 
123: The fermion nodes play an important role in 
124:  quantum Monte Carlo (QMC) methods which belong to the
125:  most promising and productive
126: approaches for studying quantum many-body
127: systems.
128: Let us briefly describe the basic idea of QMC and its relationship 
129: to fermion nodes. 
130: Consider a  Hamiltonian $H$ and  a
131: trial function $\Psi_T(R)$ which approximates the ground state
132: $\Psi_0(R)$ of $H$ within a given
133: symmetry sector. It is straightforward to show that 
134: $\lim_{\tau\to\infty}\exp(-\tau H)\Psi_T\propto \Psi_0$ 
135: where $\tau$ is a real parameter 
136: (imaginary time). This imaginary time projection can be conveniently
137: carried out by simulating a stochastic process 
138: which maps onto 
139: the imaginary time  Schr\"odinger equation.   
140: The wave function is represented by an
141: ensemble  of $R$-space sampling points which are propagated
142:  according to  $G(R,R', \tau)=
143: \langle R |\exp(-\tau H)|R'\rangle$ and for large $\tau$
144: the ensemble becomes distributed according to $\Psi_0$.
145: Unfortunately, the straightforward application of this idea
146: to fermionic systems encounters a fundamental complication
147: in the fermion sign problem \cite{qmchistory, hammond, jbanderson}.
148: %In the simplest extension of the QMC
149: %approach to fermionic wave functions
150: The fermion sign problem makes QMC studies of large fermionic systems
151: difficult since the statistical errors 
152: of fermionic expectation values grow exponentially in the projection
153: time $\tau$ and also in the number of particles $N$. 
154: %The difficulty stems from the fact that, in general,
155: %probabilistic methods are difficult to apply to
156: %%distributions which are not strictly
157: %nonnegative and finds and one encounters an
158: %exponential loss the fermionic "signal" on the top
159: %of bosonic "noise".
160: 
161:  One possibility for avoiding the fermion sign problem
162: is to employ the fixed-node approximation 
163: which restricts the fermion node
164:  of the solution $\Psi_0(R)$
165: to be identical to the fermion
166: node of an appropriate trial function $\Psi_T(R)$.
167:  The fixed-node approximation introduces an energy bias which
168: scales as the square of the difference between the exact and approximate
169: nodes and therefore the accuracy of 
170: fermion nodes becomes very important. In particular,
171: for the exact node one can obtain the exact energy in computational
172: time proportional to a low-order polynomial in $N$. The fixed-node
173: approximation has been very successful even for rather approximate
174: nodes of 
175: commonly used trial wave functions which are based on
176: Hartree-Fock (HF) determinants or a few-determinant post-HF expansions.
177: %The approximate nodes of these trial functions
178: % allowed to obtain a number of valuable results which
179: %become established benchmarks
180: %and produced predictions which would be difficult to obtain by
181: %other approaches. Let us mention calculations of homogeneous electron gas,
182: %quantum liquids, lattice models and electronic structure of
183: %atoms, molecules, surfaces, solids and other systems
184: % over the past two decades
185: %% \cite{qmchistory,hammond,qmcrev}.
186: %Quantum many-body systems belong to the most intriguing and
187: %exciting challenges in physics. 
188: %Perhaps the most successful approach to investigate the 
189: %properties of quantum systems in recent years has been 
190: %the quantum Monte Carlo method. 
191: The fixed-node QMC electronic structure calculations
192: using HF or post-HF nodes usually recover
193: between 90 and 95 \%
194: of the correlation energy 
195: and energy differences 
196: agree with experiments typically within a few percent.  
197: Such encouraging results have been observed across many 
198: systems such as atoms, molecules, clusters and solids \cite{qmcrev}.
199: It is remarkable that
200: QMC methods have enabled us to "corner" the correlation energy problem
201: into the last few percent of the correlation energy using algorithms which 
202: polynomial scaling.
203: 
204: %Perhaps the most unique aspect of the QMC method is that
205: %it enables us to combine the known analytical properties
206: %of wave functions, such as 
207: %electron-electron cusps, with the
208: %robustness of stochastic approaches and provides thus 
209: %a unique capability to attack challenging quantum many-body
210: %problems. In addition, so far there is no fundamental obstacle
211: %on improving the wave functions and finding better functional
212: %%forms such as Bardeen-Cooper-Schrieffer pairing wave functions, 
213: %pfaffians  and backflow wave functions. One gets not only more accurate
214: %%results which in some cases start to reach the spectroscopic 
215: %accuracy but also revealing insights into the character of many-body effects.
216: 
217: %In this approach one forces the nodes of the solution (ie, places whwer
218: %of configurations for which the wave function vanishes due to the 
219: %antisymmetry) of the solution to be identical to the fermion
220: %node of the trial wave function $\Psi_T$. 
221: 
222: %It is now well
223: %understood that if there are any near-degeneracies present one needs to include 
224: %the corresponding excitations into the determinantal part of the wave function 
225: %post-Hartree-Fock methods such as Configuration Interaction or similar
226: %multi-determinant approaches. Although the analysis of the wave functions 
227: %to find out the important excitations might be tedious once this is done the
228: %%fixed-node QMC is very effective. 
229: 
230: Unfortunately,
231: for many problems even a few percent of the correlation energy can be 
232: significant. Typical
233: examples are  transition metal systems where the fixed-node error
234: can be of the order of several eVs. 
235: It is therefore clear that better understanding of fermion nodes
236: could be an important step forward for the QMC methodology and beyond.
237:  In addition, the fermion nodes are
238: related to other physical quantities and could shed light on other many-body phenomena 
239: which currently are not completely understood.
240: %fermion nodes to other physical quantities and effects which are not completely
241: %understood, some of which will be addressed also in this paper. 
242: 
243: %{\it Exact nodes.}
244: %So far our knowledge about the properties of fermion nodal hypersurfaces is quite limited. 
245: In order to introduce the basic properties of the fermion nodes 
246: and to illustrate the problems 
247: involved we will first mention a few unique cases of interacting systems
248: for which the nodes are known exactly.
249: These examples include a few two- and three-electron atomic states, namely,
250: %It is interesting that
251: %the exact nodes are known for a few special cases such as 
252: triplets of He atom  $^3S(1s2s)$, $^3P(2p^2)$
253: \cite{dariohe}
254: and the exact node of a three-electron atomic state
255:    $^4S(p^3)$ \cite{bajdichnode}. 
256:  The wave functions of these
257:  high symmetry states can be parametrized by appropriate 
258: coordinate  maps in which the node is described
259: by a single variable. For example, in the case of He triplet $^3S(1s2s)$ state the 
260: corresponding "nodal coordinate" is 
261: $\cos\beta={\bf r}^+_{12}\cdot {\bf r}_{12}/( r^+_{12} r_{12})$ where
262: ${\bf r}^+_{ij}={\bf r}_{i}+{\bf r}_{j} $, ${\bf r}_{ij}={\bf r}_{i}-{\bf r}_{j} $.
263: For the triplet $^3P(2p^2)$ the relevant variable is
264: $\cos\omega'={\bf z}_0\cdot ({\bf r}_{1}\times {\bf r}_{2})$ assuming 
265: that the
266: $P$ state is oriented along the $z$-axis which is specified by
267: the unit vector ${\bf z}_0$ \cite{dariohe,bajdichnode}.
268: For the quartet state 
269:  $^4S(p^3)$
270: the node is captured by the variable 
271: $\cos\omega={\bf r}_{1}\cdot ({\bf r}_{2}\times {\bf r}_{3})$.
272: The node in these 
273: systems is encountered whenever the corresponding variable,
274:  $\cos\beta,\cos\omega'$
275: or $\cos\omega$,
276: vanishes.
277: 
278: The nodal surface divides the space of fermion configurations into
279: nodal cells, sometimes also called domains.
280: (The name nodal cell was introduced in the previous paper \cite{davidnode}.
281: In algebraic geometry and topology
282:  nodal cells are usually called nodal domains, see, for example,
283: Ref. \cite{berger}. We will use
284: both of these two expressions interchangeably.)
285:  The few known examples of exact nodes mentioned in the preceding paragraph help to illustrate an 
286: important point, which has been conjectured for the fermionic ground states
287: for some time \cite{davidnode}, namely,
288: that the ground state node divides the configuration space into
289: the minimal number of two nodal cells: a "plus" cell with $\Psi>0$ and
290:  a "minus" cell with $\Psi<0$.  From the node equation 
291: $\cos\beta=0$  of the $^3S(1s2s)$ state we easily find
292: that the "plus" and "minus" domains are given by the 
293: conditions $r_1>r_2$ and $r_1<r_2$, respectively.
294: For the node given by $\cos\omega'=0$ the nodal domains are given
295: by the orientation of ${\bf z}_0, {\bf r}_{1},{\bf r}_{2}$: the three 
296: vectors are either left- or right-handed, corresponding to either "plus" or "minus"
297: cell, respectively. Similarly, for the node $\cos\omega=0$
298: the domains are given by the left- or right-handedness of
299: the vectors ${\bf r}_{1},{\bf r}_{2}, {\bf r}_{3}$.
300: The minimal number of
301: two nodal domains
302: was found  
303: also for $2D$ and $3D$ noninteracting spin-polarized homogeneous
304: electron gas with periodic boundary conditions for up to 200 particles using
305: a numerical proof \cite{davidnode}.
306: 
307: Understanding the fermion nodes and their properties has become a 
308: challenge which might help to advance both practical calculations
309: but also open a deeper insights into the properties of fermionic systems.
310: One can envision two key problems:
311: 
312: a) {\it Topology of the fermion nodes, ie, how many nodal cells
313: are actually present} since this is of high importance for the fixed-node 
314: QMC approaches. Note that if the number 
315: of nodal cells is incorrect (typically {\it higher} than it should be since 
316: mean-fields such as HF 
317: have tendency to divide the configurations space into too many domains) then the QMC sampling
318: around the artificial nodes will be very sparse. This could lead to 
319:  large statistical fluctuations from poor sampling, 
320: and, possibly, to an 
321: effective non-ergodicity due to the finite-time projection time in practical calculations.
322: 
323: b) Once the topology is correct, {\it the accuracy of the manifold 
324: shape becomes important.} This is an area
325: where our insights are particularly limited since the exact nodes for 
326: large interacting systems are virtually unknown except
327: for a few-particle special cases mentioned above. 
328: 
329: In our recent paper \cite{mitasshort},
330: we have made some encouraging steps forward in trying to understand the 
331: topological issues
332: and we have 
333: analytically derived a number of new results regarding the number of nodal domains.
334: In particular, we have shown that spin-polarized 
335: closed-shell ground states of noninteracting harmonic 
336: fermions in $d=2$ and higher 
337: have the minimal number of two nodal domains.
338: % for {\it an arbitrary size of the system}.
339: We have proved the same for
340: %spin-unpolarized harmonic fermions with interactions, 
341: spin-polarized atomic states with several electrons,
342: both for noninteracting and HF wave functions.
343: We have also shown that
344: by imposing additional symmetries one can generate more than two
345:  nodal domains but that
346: interactions  can relax this "nodal degeneracy" to the minimal number of two 
347: domains such as in the case of the $^4S(1s2s3s)$ atomic state.
348: 
349: For noninteracting spin-unpolarized systems, ie, with both spin channels
350: occupied, the number of nodal cells is four since the wave function is
351: a product of spin-up and spin-down Slater determinants ($2\times 2=4$).
352: (Here and later on we assume that the Hamiltonian does not include spin-dependent
353: terms so that the particle spins are conserved. 
354: The wave function is then a product of two determinants which depend 
355: only on the spatial degrees of freedom \cite{qmcrev}.)
356: In the last few years, studies of a few-particle systems have revealed
357: that interactions and many-body effects
358: do affect the nodal topologies
359: and can change the number of nodal cells \cite{dariobe,cmt28}.
360: In particular, for the case of Be atom it has been found that
361: the noninteracting/HF four nodal cells of the singlet ground state
362: change to the minimal number of two due to the electron correlation
363: \cite{dariobe} and qualitatively the same has been observed
364: in  other systems \cite{cmt28}. 
365: In our recent paper \cite{mitasshort} we have found that this is a rather
366: generic property of ground states in interacting systems.
367:  With some conditions, 
368:  we have explicitly demonstrated that interactions lift
369: the "nodal cell degeneracy" in spin-unpolarized systems
370: and smooth out the four noninteracting cells
371: into the minimal two. We have shown this for
372:  $2D$  harmonic fermions in closed-shell states {\it of arbitrary size}
373: using a variational Bardeen-Cooper-Schrieffer wave function.
374: 
375: In this work we further advance these ideas.
376: % and apply it to more types of systems.
377: We analyze the fermion nodes in a number of other paradigmatic fermionic
378: models: homogeneous electron gas with periodic
379: boundary conditions, particles in an infinite well and on the surface of a sphere.
380: For all these systems
381: we prove that in the spin-polarized noninteracting closed-shell ground state the fermion nodes 
382: divide the configuration space into minimal two cells for arbitrary number of
383: particles.
384: We also extend our previous proof for spin-unpolarized systems
385: demonstrating 
386: that interactions smooth out multiple nodal cells of noninteracting/mean-field
387: wave functions into the minimal two for more systems such as homogeneous
388: electron gas and $3D$ harmonic oscillator. 
389: %We show that by introducing the repulsive/attractive interaction
390: %this property does not change.
391: %For spin-unpolarized or partially polarized systems we show that in
392: %the noninteracting case the number of nodals cells is trivially four since
393: %the wave function is a product of determinants in
394: %the spin-up and -down channels. This is also the case of
395: %wave functions for unpolarized systems in Hartree-Fock(HF) theory.
396: % We show that arbitrarily small interaction removes
397: %the spin-up and -down degeneracy and the consequence of this is that the
398: %nodal partitioning changes from four to two.
399: These results
400:  contribute to our understanding
401: of the fermion nodes and their impacts on the accuracy 
402: of wave functions with direct implications for the QMC methods. 
403: 
404: In the last sections we show how in periodic
405: spin-unpolarized interacting fermion gas the pairs of particles 
406: can wind around the box without crossing the node
407: what points towards the connection of nodes 
408: to the transport and to the 
409: existence of other quantum phases such as superconductivity.
410:  Finally, we then
411: generalize the results to the temperature density matrices 
412: with the implications
413: for path integral Monte Carlo methods \cite{davidhe}.
414: 
415: 
416: \section{II. General properties of fermion nodes.}
417: 
418: Let us introduce the basic properties of fermion nodes
419: %The fermion node of a wave function $\Psi(R)$
420: %is implicitly defined as 
421: %the set of configurations for which $\Psi(R)=0$ (here we assume that we have omitted
422: %from the nodal set the configurations for which the wave function vanishes because of 
423: %other reasons such as singular potential or boundary conditions). Assuming that we have 
424: %%$N$ spin-polarized fermions 
425: % in $d$-dimensional space the node is a $(dN-1)-$dimensional manifold 
426: %(hypersurface). 
427: as they were studied
428: by Ceperley \cite{davidnode} some time ago. 
429: 
430: a) Nondegenerate ground states wave functions fulfill the so-called 
431: tiling property.  Let us define a nodal cell
432:  $\Omega(R_0)$ as a subset of configurations which can be reached
433: from the point $R_0$ by a continuous path without crossing the node.
434: %(The name nodal cell was introduced in the previous paper \cite{davidnode}.
435: %In algebraic geometry and topology  
436: % nodal cells are called nodal domains, see, for example,
437: %Ref. \cite{berger}. We will use 
438: %both of these two expressions interchangeably.)
439: The tiling property
440: says that by applying all possible particle permutations to an arbitrary
441: nodal cell
442: one covers the complete configuration
443: space. Note that this {\em does not} specify how many nodal cells are there.
444: 
445: b) If $m$ nodal surfaces cross then the angle of crossing is
446: $\pi/m$.
447:  Furthermore, the symmetry of the node is the same as the
448: symmetry of the state.
449: 
450: 
451: c) 
452: It is possible to show that there are only two nodal cells using the following
453: argument based on triple exchanges. Let us first 
454: introduce the notion of particles connected by triple exchanges. 
455: % (continuously connected) by the following construction. Consider
456: % a spin-polarized system
457: %with the wave function $\Psi(R)$ where $R=({\bf r}_1, ..., {\bf r}_N)$ are particle
458: %coordinates.
459: We will call the three particles $i,j,k$  {\it connected}
460: if there exists a triple exchange path $i\to j, j\to k, k\to i$ which does 
461: not cross the node. If more than three particles are connected then  
462: they can form a connected cluster. An example of 
463: six particles connected into a single cluster 
464: is sketched as follows: 
465: \begin{picture}(12,7.5)(6,4.5)
466: \put(2,5){$\bullet$}
467: \put(6,2){$\bullet$}
468: \put(6,8){$\bullet$}
469: \put(10,5){$\bullet$}
470: \put(14,2){$\bullet$}
471: \put(14,8){$\bullet$}
472: \drawline(3.5,7)(7.7,4)(7.7,10)(3.5,7)
473: \drawline(7.7,10)(15.8,4)
474: \drawline(7.7,4)(15.8,10)(15.8,4)
475: \end{picture}.
476: If there exists a point $R_t$ such that {\it all} the 
477: particles are connected  into a single cluster then $\Psi(R)$ has only two nodal cells.
478: This can be better understood once we realize the following two facts.
479: First, any triple 
480: permutation can be written as two pair permutations. Therefore the connected
481: cluster of triple permutations enables to realize any even parity permutation 
482: without crossing the node. That exhausts all permutations which are available
483: for cell of one sign since the  wave function is invariant to even parity permutations. 
484: Second, the tiling property implies that once the particles are connected for $R_t$
485: the same is true for the entire cell. By symmetry, the same arguments
486: apply 
487: to the complementary "minus" cell which correspond to the odd permutations. 
488: More details on this property can be found in the original Ref. \cite{davidnode}.
489: 
490: % We show that arbitrarily small interaction removes
491: %the spin-up and -down degeneracy and the consequence of this is that the
492: %nodal partitioning changes from four to two.
493: 
494: 
495: \section{\label{sec:sp} III. Noninteracting spin-polarized fermions.}
496: 
497: \subsection{III.a. Homogeneous electron gas. }
498: 
499: %{\it Noninteracting homogeneous electron gas with 
500: %periodic boundary conditions.} 
501: 
502: We consider a
503: system of spin-polarized noninteracting
504: fermions in a periodic box  in $d$ dimensions. The  
505: spatial coordinates are rescaled by the box size
506: so that we can use dimensionless variables and the box
507: becomes
508: $(-\pi,\pi)^d$. 
509: 
510: We first analyze the fermion nodes for $d=1$ since the result will be
511: useful in subsequent derivations. We consider 
512: a system with $N=(2k_F+1)$ particles. In our $1D$ dimensionless
513: units the Fermi momentum becomes an integer,
514: $k_F=1,2 ...$.  The one-particle occupied states are written 
515: as $\phi_n(x)=e^{inx},$ $n=0,\pm 1,...,\pm k_F$ 
516: and the spin-polarized ground state is given by
517: \begin{equation}
518: \Psi_{1D}(1,...,N)={\rm det}\left[ \phi_n(x_j)\right]
519: \end{equation}
520: %\begin{equation}
521: %{\rm det}\{1,e^{\pm ix}, e^{\pm i2x}, ..., e^{\pm ik_Fx}\}
522: %\end{equation}
523:  where
524: $x_j$ is the $j$-th particle coordinate and $j=1,...,N$.
525: We factorize the term 
526: $\exp(-ik_F\sum_jx_j)$ out of the determinant so that the 
527: Slater matrix elements become powers of $z_j=e^{ix_j}$. 
528: The resulting Vandermonde determinant can be 
529: written in a closed form and after some rearrangements we find 
530: \begin{equation}
531: \Psi_{1D}(1,...,N)=
532: e^{-ik_F\sum_jx_j}\prod_{j>k}(z_j-z_k)=\nonumber
533: \end{equation}
534: \begin{equation}
535: =\mu_0\prod_{j>k}
536: \sin(x_{jk}/2)
537: \label{eq:van1d}
538: \end{equation}
539: where 
540: $x_{jk}=x_j-x_k$ and $\mu_0$ is
541: a constant prefactor which is unimportant 
542: for our purposes.  (In the derivations below we will be using
543: the letter $\mu$ for denoting prefactors which are either
544: constants or nonnegative functions in particle coordinates
545: and therefore do not affect the nodes).  
546: %(In order 
547: %to simplify 
548: %the expressions
549: %we will be omitting similar constant prefactors
550: %whenever appropriate.) 
551: %Assuming that in the
552: %expression above the particle positions are all distinct,
553: The derived wave function has the following important properties.
554: 
555: a) The fermion nodes
556: appear at the particle coincidence points. This implies a well-known 
557: result that the ground state wave function in $1D$ have $N!$ nodal cells since any 
558: particle permutation  
559: requires crossing the node at least once. In addition, this also means that 
560: {\it any} fermion configuration which 
561: preserves the particle order is contained within the same nodal domain/cell.
562: %In particular, winding the particle around the box 
563: %without crossing any other particle keeps the configuration in the same
564: %nodal domain. 
565: 
566: % Note that whenever a particle
567: %of some of the terms might need adjustment so that the wave function is continuous.
568: %%For example, a particle with the label $i$ moves in $+x$ direction hits 
569: %$x_i=\pi$ and reenters as $x_i \to x_i-2\pi$
570: %so that the pair terms in the wave function transform to
571: %\begin{equation}
572: %\sin(x_{ij}/2-\pi)=-\sin(x_{ij}/2)=\sin(x_{ji}/2)
573: %\end{equation}
574: %implying exchange of $x_{ij} \to x_{ji}$ in all the
575: % pairs which involve the $i-$th particle.
576: 
577: 
578: b) The wave function is invariant to
579: translations and cyclic exchanges of particles. 
580: Due to the periodic boundary conditions
581: this includes also winding the system
582: around the periodic box, ie, the translation of all particles
583: by $2\pi$. We can formalize 
584: this by introducing an operator $T^{x}_a$ which translates all the particles 
585: along the $x$-axis
586: as $x_j \to x_j+a$ so that the translation invariance can be written as
587: \begin{equation}
588: T_a^x \Psi_{1D}(1,...,N)= \Psi_{1D}(1,...,N)
589: \label{eq:inv1}
590: \end{equation}
591: 
592: Similarly, the wave function  is invariant to 
593: cyclic exchange of all the particles given by
594: $j \to j+1, j=1, ..., N$ and $N+1 \to 1$.
595:  This action is carried out   
596: by $C_{+1}^{x}$ operator where the notation means the 
597: exchange by one site is in the $+x$-direction. 
598: Clearly, the inverse operator
599: $[C_{+1}^x]^{-1}=C_{-1}^x$. Note that here we have assumed that $N$ is odd, in agreement with our 
600: definition.  The invariance holds only for $N$ odd since then
601: the cyclic exchange is equivalent
602: to an even number of pair exchanges. For the sake of completeness, we consider
603: also $N$ even, when the cyclic exchange
604: can be replaced by an odd number of pair exchanges, resulting
605: in the wave function sign flip.
606: In general, for the cyclic exchanges we can therefore write  
607: \begin{equation}
608: C_{\pm 1}^x\Psi_{1D}(1,...,N)=(-1)^{N+1}\Psi_{1D}(1,...,N)
609: \label{eq:inv2}
610: \end{equation}
611: 
612: c) Assuming the particle positions are all distinct, 
613: the cyclic exchange {\it path} can be chosen in such 
614:  a way that it
615: {\it does not cross the node}. 
616: %It is not difficult to imagine 
617: %that the particles move in a concerted manner
618: %and never cross each other along the whole cyclic exchange path. 
619: This is easy to accomplish by maintaining 
620: finite distances between the particles along the path trajectory. 
621: Let us parametrize the exchange path by a parameter $t$ so that the
622: path starts at
623:  $t=0$, the path is completed at $t=1$ and the  
624:  exchange path operator is then denoted as $C^x_{+1}(t)$ with $0\leq t\leq 1$.
625: The fact that the path  
626: does not cross the node (ie, the path is contained within the same 
627: nodal cell) can be then written as
628: \begin{equation}
629: |C_{+1}^x(t)\Psi_{1D}(1,...,N)|>0, \; 0\leq t \leq 1
630: \label{eq:inv3}
631: \end{equation}
632: where, of course, $N$ is assumed to be odd.
633: 
634: Let us now derive the wave functions and generalize the results for 
635: fermions in $2D$. The one-particle states in $2D$ are  
636: $\phi_{nm}(x,y)=e^{i(nx+my)}$.  
637: The states are occupied up to the Fermi momentum $k_F$ so that
638: we have $n^2+m^2\leq k_F^2$, where $k_F$ in $2D$ is not necessarily an integer. 
639: Similarly to our previous paper \cite{mitasshort},
640: we show that the spin-polarized electron gas  
641: for closed-shell
642: ground
643: states have only two nodal cells.
644: %The number of particles is given by
645: %$N=2M(M+1)+1$ where $M$ labels the closed-shell states with the
646: %Fermi surface given by (a rotated) square see Fig. .
647: %$M$ denotes also the highest power of $e^{ix}$ in one-particle states
648: %and we have then $(2M+1)$ lines of particles (vertically or horizontally).
649: 
650: \begin{figure}[ht]
651: \centering
652: \begin{tabular}{ccc} 
653: \includegraphics[width=1.1in,clip]{fig1A.eps} &
654: \includegraphics[width=1.1in,clip]{fig1B.eps} &
655: \includegraphics[width=1.1in,clip]{fig1C.eps} \\
656: (a) & (b) & (c)
657: \end{tabular}
658: \caption{Positions of five fermions in the $2D$
659:  periodic box $(-\pi,\pi)^2$. (a) Particles aligned horizontally,
660: along the lines $y=\zeta_0,\zeta_1,\zeta_2$.
661: (b) Particles aligned vertically, along the lines  $x=\xi_0,\xi_1,\xi_2$.  (c) Particles aligned 
662: in both directions, positioned on a lattice.}
663: %See the text for details. }
664: \label{fig:five2d}
665: \end{figure}
666: 
667: The proof is by induction, therefore 
668: let us first consider $k_F=1$ with five particles
669: occupying $\{1,e^{\pm ix},
670: e^{\pm iy}\}$ one-particle states. 
671: We place the particles as in Fig.\ref{fig:five2d}a so that
672: the coordinates are given by 
673: ${\bf r}_1=(x_1,\zeta_1)$, ${\bf r}_2=(x_2,\zeta_1)$, ${\bf r}_3=(x_3,\zeta_1)$,
674:  ${\bf r}_4=(x_3,\zeta_2)$,  ${\bf r}_5=(x_5,\zeta_0)$.
675: By eliminating the terms with $\zeta_1$ 
676: the wave function can be factorized 
677: as follows  
678: \begin{equation}
679: \Psi_{2D}(1,...,5)=\left| \begin{array}{ccccc}
680:    1  &  1   &  1   &    1  &  1    \\
681:  e^{ix_1}  &  e^{ix_2}   &  e^{ix_3}   &
682:    e^{ix_4}  &  e^{ix_5} \\
683:  e^{-ix_1}  &  e^{-ix_2}   &  e^{-ix_3}   &
684:    e^{-ix_4}  &  e^{-ix_5} \\
685:  e^{i\zeta_1} &  e^{i\zeta_1} &
686: e^{i\zeta_1} &  e^{i\zeta_0} &
687: e^{i\zeta_2} \\
688:  e^{-i\zeta_1} &  e^{-i\zeta_1} &
689: e^{-i\zeta_1} &  e^{-i\zeta_0} &
690: e^{-i\zeta_2} \\
691: \end{array} \right|=\nonumber
692: \end{equation}
693: \begin{equation}
694: \mu_0\left| \begin{array}{ccc}
695:    1  &  1   &  1     \\
696:  e^{ix_1}  &  e^{ix_2}   &  e^{ix_3}  \\
697:  e^{-ix_1}  &  e^{-ix_2}   &  e^{-ix_3}  \\
698: \end{array} \right|
699: \left| \begin{array}{cc}
700:   e^{i\zeta_0}-e^{i\zeta_1} &
701: e^{i\zeta_2} -e^{i\zeta_1}\\
702:  e^{-i\zeta_0} -e^{-i\zeta_1} &
703: e^{-i\zeta_2} -e^{-i\zeta_1} \\
704: \end{array} \right|=\nonumber
705: \end{equation}
706: 
707: \begin{equation}
708: = \mu_0'\Psi_{1D}(1,2,3)\sin(\zeta_{10}/2)\sin(\zeta_{20}/2)\sin(\zeta_{21}/2)
709: \label{eq:five2dzeta}
710: \end{equation}
711: where $\mu_0,\mu_0'$ are irrelevant constant prefactors.
712: We have obtained a product of $1D$ wave function and
713: terms with distances between the
714: {\it lines} $y=\zeta_0,\zeta_1,\zeta_2$.
715: Note that analogous result can be found for the configuration
716: sketched in Fig. 1b so the particles are aligned 
717: in parallel to the 
718: $y$-axis with the wave function given by
719: \begin{equation}
720: \Psi_{2D}(1,...,5)=
721:  \mu_0\Psi_{1D}(5,2,4)\times \nonumber
722: \end{equation}
723: \begin{equation}
724: \times \sin(\xi_{10}/2)\sin(\xi_{20}/2)\sin(\xi_{21}/2)
725: \label{eq:five2dxi}
726: \end{equation} 
727: If the particles are aligned in both directions as in Fig. 1c
728: both expressions apply.
729: % and give the same result. 
730: From the derived wave function it is clear that the node is encountered
731: when particles lying on the same line reorder or 
732: when the lines of particles 
733: cross each other, eg, $\xi_1=\xi_0$.
734: %Note that for all three configurations in Fig. 1 the 
735: %particles occupy the {\it same } nodal domain since the configurations 
736: %can be obtained from each other by repositioning 
737: %the particles without crossing the node.
738: %This is easy to see because
739: %the repositioning can be done without any
740: %reordering of particles along $1D$ directions 
741: %and without any two lines
742: %of particles crossing each other. 
743: The wave functions for the configurations outlined 
744: in Fig.\ref{fig:five2d} possess an  
745: important property. Consider 
746: $\Psi_{1D}(...)$ in Eqs.  \ref{eq:five2dzeta},  \ref{eq:five2dxi} with
747:  groups of particles positioned on the corresponding lines.
748: Assuming the group of particles is allowed to move only along the given line,
749: one can consider this to be an effective $1D$ subspace.
750: The $1D$ wave function for such a subspace  
751: obeys all three conditions derived
752: for $1D$ case, ie, 
753:  Eqs. \ref{eq:inv1}, \ref{eq:inv2} and \ref{eq:inv3}. 
754: For example, for $\Psi_{1D}(1,2,3)$ which appears 
755:  in Eq. \ref{eq:five2dzeta} we have
756:  $|C_{+1}^x(t)\Psi_{1D}(1,2,3)|>0$. This is easy to check also analytically. 
757: We fix the coordinates in 
758:  Fig.1(a) as follows $x_1=-2\pi/3, x_2=0, x_3=2\pi/3$ 
759: and then we can 
760: carry out a "synchronized" cyclic exchange using the translation  
761: by $2\pi/3$ in $x$-direction. The 
762: wave function is constant 
763: along the translation/cyclic exchange path so that we have 
764: \begin{equation}
765: C^x_{+1}(t) \Psi_{1D}(1,2,3)=\Psi_{1D}(1,2,3), \;\; 0\leq t \leq 1
766: \end{equation}
767: because of the translational invariance.
768: 
769: Finally, we are ready to show that there are only two nodal cells in this 
770: five-particle ground state. 
771: Assume that the particles are positioned as in Fig.1c. 
772: Consider the following sequence of 
773: four exchanges, $C^x_{+1}C^y_{+1}C^x_{-1}C^y_{-1}$, where the operators act
774: from the right, ie, $C^y_{-1}$, acts first.  We denote this exchange symbolically as 
775: \begin{equation}
776: C_{\rightarrow\uparrow\leftarrow\downarrow}=C^x_{+1}C^y_{+1}C^x_{-1}C^y_{-1}
777: \end{equation}
778: The exchanges are performed only for  
779: the particles lying on the lines $\zeta_1,\xi_1$ in corresponding directions,
780: eg, $C_{\pm 1}^x$ acts on particles along  $y=\zeta_1$ while  $C_{\pm 1}^y$ 
781: cycles particles along $x=\xi_1$ . 
782: It is easy to check that this results in a triple exchange 
783: $423 \to 342$ while the particles 1 and 5 end up in their original
784: positions. It is therefore clear that particles 2,3,4 are connected by 
785: this triple exchange.
786: It is straightforward to show that 
787: $C^x_{-1} C^y_{-1} C^x_{+1}C^y_{+1}$ exchanges particles 1,2,5 so that 
788: we can conclude that all five particles are connected by triple exchanges    
789: into a single cluster
790: and there are only two nodal domains.
791: 
792: Now we need to extend the arguments and find the wave functions for
793: general cases 
794: with occupied states within the Fermi disk with an arbitrary $k_F$.
795: It is instructive to first derive 
796: the wave functions for $k_F=\sqrt{2}$ with 9 particles
797: occupying
798:  states $\{ 1,e^{\pm ix}, e^{\pm iy}, e^{i(\pm x \pm y)}\}$
799:  and 
800: for $k_F=2$ with 13 particles  with additional states 
801: $\{ e^{\pm i2x}, e^{\pm i2y}\}$.
802: This will enable us to understand how to  
803: write down the wave function
804: for a general case while avoiding rather tedious notations 
805: which would appear in the direct derivations.
806: % for a general case.
807: 
808: \begin{figure}[ht]
809: \centering
810: \begin{tabular}{ccc}
811: \includegraphics[width=1.35in,clip]{fig2A.eps} &
812: \quad &
813: \includegraphics[width=1.35in,clip]{fig2B.eps} \\
814: (a) & \quad & (b)
815: \end{tabular}
816: \caption{Positions of fermions in
817:  the $2D$ periodic box in real space. (a) Positions of the nine-fermion system.
818: (b) Positions of the 13-fermion system.
819: See the text for details. }
820: \label{fig:nine2d}
821: \end{figure}
822: 
823:  Let us position the particles into a pattern which mimics the lattice
824: of the occupied wave vectors in the reciprocal space, see Fig.\ref{fig:nine2d}.
825: Using appropriate algebraic arrangements we subsequently factorize the lines 
826: of particles as given by
827: \begin{equation}
828: \Psi_{2D}(1,...,9)= \nonumber
829: \end{equation}
830: \begin{equation}
831: =\mu_0\Psi_{1D}(1,2,3)\Psi_{2D}(4,...,9) \sin^3(\zeta_{21}/2) \sin^3(\zeta_{10}/2)=\nonumber
832: \end{equation}
833: \begin{equation}
834: =
835: \mu_0'\Psi_{1D}(1,2,3)\Psi_{1D}(4,5,6)\Psi_{1D}(7,8,9)\times \nonumber
836: \end{equation}
837: \begin{equation}
838: \times \sin^3(\zeta_{21}/2) \sin^3(\zeta_{10}/2)\sin^3(\zeta_{20}/2)
839: \end{equation}
840: where $\mu_0,\mu_0'$ are constants.
841: Similarly, for $k_F=2$ we can first factorize
842: the line with the largest number of particles on 
843: $y=\zeta_2$ (see Fig.\ref{fig:nine2d})
844: \begin{equation}
845: \Psi_{2D}(1,...,13)=
846: \mu_0\Psi_{1D}(1,...,5)\sin(\zeta_{02}/2)\sin^3(\zeta_{12}/2)
847: \times
848: \nonumber
849: \end{equation}
850: \begin{equation}
851: \times \sin^3(\zeta_{32}/2)\sin(\zeta_{42}/2)
852: \Psi_{2D}(6,...,13)\nonumber
853: \end{equation}
854: and 
855: then factorize out the particles lying on another line $y=\zeta_3$ to obtain
856: \begin{equation}
857: \Psi_{2D}(1,...,13)=
858: \mu_0\left[\prod_{j\neq 2}\sin^{n_j}(\zeta_{j2}/2)\right]
859: \left[\prod_{j\neq 2,3}\sin^{n_j}(\zeta_{j1}/2)\right]
860: \times
861: \nonumber
862: \end{equation}
863: \begin{equation}
864: \Psi_{1D}(1,...,5)\Psi_{1D}(6,7,8)\Psi_{2D}(9,...,13)
865: \end{equation}
866: where $n_j$ is the number of particles lying on the line $y=\zeta_j$.
867: We have obtained a product of $1D$ wave functions, 
868: terms with distances 
869: between the lines and the $2D$ wave function with lower number of particles
870: positioned in the same type of pattern is in Fig.\ref{fig:five2d}. Obviously, we can
871: further factorize  $\Psi_{2D}(9,...,13)$ using Eq.\ref{eq:five2dzeta} until
872: we end up 
873: with $1D$ factors only.
874: % At this point it has become
875: %clear that we can evaluate 
876: %the determinant by factorizing out the lines of particles
877: %in a recursive manner.
878: Using these insights into the recursive form of the wave function, 
879: for a general case with $M+1$ lines we can write
880: \begin{equation}
881: \Psi_{2D}(1,...,N)=\nonumber
882: \end{equation}
883: %\begin{equation}
884: %=\pm\Psi_{2D}(1,...,N/I_k)\Psi_{1D}(I_k)\prod_{j\ne k} \sin^{n_j}(\zeta_{jk}/2)
885: %\nonumber
886: %\end{equation}
887: \begin{equation}
888: =\mu_0\prod_{k=0}^{M}\left[\Psi_{1D}(I_k)\prod_{j>k } 
889: \sin^{n_j}(\zeta_{jk}/2)\right]
890: \end{equation}
891: where $I_k=i^{(k)}_1, ..., i^{(k)}_{n_k}$
892: denotes the labels of particles
893: lying on the line $y=\zeta_k$.
894: %The sign depends on the number of row exchanges
895: %and on the actual ordering of particles. 
896: In addition, analogous expression can be found if we factorize along the $x=\xi_j$
897: lines, the only difference being replacement $\zeta_{jk}$ by $\xi_{jk}$ 
898: and corresponding replacement of particle sets in $\Psi_{1D}$ wave functions.
899: %Note the translation
900: %invariance of Eq. \ref{eq:recur}.
901: 
902: 
903: \begin{figure}[ht]
904: \centering
905: \begin{tabular}{ccc}
906: \includegraphics[width=1.35in,clip]{fig3A.eps} &
907: \quad &
908: \includegraphics[width=1.35in,clip]{fig3B.eps} \\
909: (a) & \quad & (b)
910: \end{tabular}
911: \caption{Illustrations of the position patterns in $2D$ periodic box for the
912:  size increases from $k_F$ to $k_F+\Delta k_F$.
913: The particle layouts mimic the occupied states in the
914: reciprocal space, however, the particles are positioned in the real space. 
915: The additional particles are in grey/yellow.
916: (a) The additional particles positioned along $x-$ and $y-$axis directions.
917: (b)  The additional particles positioned along diagonals.
918: Lines $x=\xi_k$ and $y=\zeta_l$ are used for the proof that the particles $a,b,c$
919: are connected by a triple exchange.
920: See the text for details. }
921: \label{fig:indu2d}
922: \end{figure}
923: 
924: We are now ready for the induction step. Consider a 
925: spin-polarized closed-shell ground state
926: with a given $k_F$. For this wave function we assume that all the
927: particles are connected by the triple exchanges, ie, there are only two nodal domains.
928: Let us
929: increase $k_F \to k_F+\Delta k_F $ until the Fermi disk
930: includes the next unoccupied star of states $\phi_{nm}(x,y)$ for which
931: $k_F < (n^2+m^2)^{1/2} \leq k_F+\Delta k_F$. The system size increases
932: by the corresponding number of additional particles.
933: Assuming the particles are positioned in the real space 
934: in the same pattern as
935:  the occupied ${\bf k}$-points in reciprocal space (Fig.\ref{fig:indu2d}),
936:  the additional particles will appear at the 
937: borderline of the disk in the real space.
938: % (corresponding to the Fermi disk
939: %in the {\bf k}-space).
940: 
941: The two basic possibilities how the additional particles are positioned 
942: are  given in the Fig.\ref{fig:indu2d}.  
943: %These two are, however, essentially the same in the sense
944: %that the following arguments one how the additional particles 
945: %are connected by exchanges with the original particles will apply to both of them.
946: We need to show that these additional particles are connected 
947: to the original particles
948: by the triple exchanges. 
949:  This can be demonstrated by
950: the sequence of the four cyclic exchanges which we have
951: used above for the five particle case. 
952: It involves particles on
953:   the lines $x=\xi_k$ and $y=\zeta_l$ as schematically drawn in Fig.\ref{fig:indu2d}.
954: Consider the exchange
955: $C_{\rightarrow\uparrow\leftarrow\downarrow}=$
956: $C^x_{+1}C^y_{+1}C^x_{-1}C^y_{-1}$  where the cyclic exchanges in $x$ direction are
957: applied only to particles on the line $x=\xi_k$ and, similarly, 
958: the cyclic exchanges in $y$ direction are applied only to particles along $y=\zeta_l$. 
959:  Note the wave function is invariant to cyclic exchanges
960: since the number of particles along each line 
961: is odd for any closed shell state. It is then straightforward to find 
962: out that $C_{\rightarrow\uparrow\leftarrow\downarrow}$
963:  exchanges particles $a,b,c$ while the rest of the particles remains intact.
964:  Similar exchanges can be carried out for 
965: all additional particles. Finally, this shows
966: that the additional particles are connected 
967: to
968: the particles of the wave function with size $k_F$ and finalizes the 
969: proof.
970: 
971: The proof can be extended into $3D$ and higher dimension by positioning the particles onto
972: an appropriate $3D$ pattern which in real space mimics the occupied states in the Fermi
973: sphere in the reciprocal space. This is possible due to the fact that with proper positioning
974: of particles
975: one can subsequently factorize the Slater
976: determinant along {\it hyperplanes, planes and lines}. Using arguments similar to the $2D$ case 
977: we can perfrom cyclic exchanges without
978: crossing the node in $3D$ and higher dimensions. Therefore the proof for higher dimensions
979: is essentially the same. 
980: 
981: In many cases,
982: the proof can be extended to open shells. If an open-shell state is degenerate
983: it is necessary to fix 
984: the ambiguity in the nodes, for example, by considering
985: wave functions which transform according to an appropriate
986: symmetry subgroup \cite{davidnode, foulkes}. In general, however, depending on symmetries,
987: the number of degenerate
988: states, etc, one cannot rule out possibilities  
989: of states with the number of
990: cells beyond the minimal two  (this will be investigated in the next paper).
991:  
992: This concludes the arguments and the proof that for $d>1$ the 
993: noninteracting spin-polarized
994: fermion gas closed-shell ground states in periodic boundary conditions have only two nodal cells. 
995: 
996: \smallskip
997: \subsection{ III.a. Fermions on the surface of a sphere.}
998: 
999:  Spin-polarized free fermions on the surface of a sphere are 
1000: described by a Slater determinant with 
1001: one-particle states being the spherical harmonics
1002: $Y_{lm}(\vartheta,\varphi)$. The spherical harmonics are polynomials 
1003: in variables $\cos\vartheta$ and $\sin\vartheta e^{\pm i\varphi}$. 
1004:  The Slater matrix elements can be  
1005: rearranged to monomials in these variables and it 
1006: is then straightforward to
1007: factorize the determinant into similar form as obtained
1008: for the homogeneous fermion gas or for the harmonic oscillator
1009: \cite{mitasshort}. Let us assume that the particles are positioned
1010: as sketched in Fig.~\ref{fig:sph}a.  Using familiar expressions for
1011: the first few spherical harmonics we get for the $^5S(sp^3)$ state 
1012: \begin{equation}
1013: \Psi_{sp^3}(1,...,4)=\mu_0(u_1-u_0) v_1^2\prod_{2\leq i<j\leq 4}\sin(\varphi_{ij}/2)
1014: \end{equation}
1015: where we have denoted $u=\cos\vartheta$ and $v=\sin\vartheta$ and $\mu_0$ 
1016: is a constant. Similarly,
1017: for the  $^{10}S(sp^3d^5)$ state with nine fermions we obtain
1018: \begin{equation}
1019: \Psi_{sp^3d^5}(1,...,9)=
1020:  \mu'_0\Psi_{sp^3}(1,...,4) \times \nonumber
1021: \end{equation}
1022: \begin{equation}
1023: \times (u_2-u_0)(u_2-u_1)^3 v_2^6 \Psi_{1D}(5,...,9)
1024: \end{equation}
1025: where we have denoted 
1026: $\Psi_{1D}(1,...,N)=
1027: \prod_{j<k}\sin(\varphi_{jk}/2)$ in agreement with Eq. \ref{eq:van1d}.
1028: 
1029: 
1030: \begin{figure}[ht]
1031: \centering
1032: \begin{tabular}{ccc}
1033: \includegraphics[width=1.35in,clip]{fig4A.eps} &
1034: \quad &
1035: \includegraphics[width=1.35in,clip]{fig4B.eps} \\
1036: (a) & \quad & (b)
1037: \end{tabular}
1038: \caption{
1039: Positions of the particles on the sphere surface for the state $^{10}S(sp^3d^5)$
1040: using coordinate system $u=\cos(\vartheta)$
1041: and $0\leq \varphi \leq 2\pi$.
1042: The particles are positioned along the lines with constant $u=u_j$.
1043: For larger systems the patterns are the same with particles corresponding to increasing
1044: angular momentum $l$ lying on the lines with increasing $u_l$. (a) Particle positions
1045: used for derivation of the wave function factorization. (b) Particles after appropriate
1046: shifts along $\varphi$ and a subsequent rotation.  }
1047: %See the text for details. }
1048: \label{fig:sph}
1049: \end{figure}
1050: 
1051: For an arbitrary size closed-shell $S$ symmetry state the one-particle 
1052: states are occupied up to the angular
1053: momentum $L$ with the corresponding number of particles  
1054: $N=\sum_{l=0}^L(2l+1)=(L+1)^2$. 
1055:  Using the previous two examples 
1056: %the $1D$ wave function for $(2L+1)$ particles 
1057: %lying on the line $u=u_l, -1<u_l<1$  and occupying the states 
1058: %$(u_l)^l,$ $(u_l)^{l-1}(1-u_l^2)^{1/2}e^{\pm i\varphi},$ $..., (1-u_l^2)^{l/2} 
1059: %e^{\pm il\varphi}$
1060: %as
1061: %\begin{equation}
1062: %\Psi_{1D}^{sph}=C(u_l)\prod_{j>i}\sin(\varphi_{ij}/2)
1063: %\end{equation}
1064: %where $C(u_l)$ is a nonnegative factor which depends only on $u_l$.
1065: we can directly write the factorized wave function assuming  
1066: the
1067: particle positions follow the pattern in Fig.~\ref{fig:sph}a. 
1068: % We occupy states up to the angular
1069: %momentum $L$ so that for spin-polarized system we have $N=(L+1)^2$ particles
1070: %in a closed-shell $S$ symmetry state.
1071: We first factorize the line $u=u_L$ and then recursively the rest of the lines
1072: so that we can write
1073: \begin{equation}
1074: \Psi(1,...,N)= \mu_0\Psi_{2D}(1,...,N/I_L)\times
1075: \nonumber
1076: \end{equation}
1077: \begin{equation}
1078: \times \Psi_{1D}(I_L)v_L^{L(L+1)}\prod_{0\leq j<L}(u_L-u_j)^{n_j}=\nonumber
1079: \end{equation}
1080: 
1081: \begin{equation}
1082: =\mu(v_1, ...,v_L)\prod_{k=0}^{L-1}\left[
1083: \Psi_{1D}(I_{L-k}) 
1084: \prod_{j<L-k}(u_{L-k}-u_j)^{n_j} 
1085: \right]
1086: \end{equation}
1087: %v_{L-k}^{(L-k)(L-k+1)} 
1088: where $I_k$ is the subset of particles lying on the line $u=u_k$.
1089:  In the final expression the powers of $v_j$, 
1090: such as $v_L^{L(L+1)}$ in the preceding line, are absorbed
1091: into $\mu(v_1, ..., v_L)$.  This prefactor is nonnegative
1092: and does not affect the nodes or the wave function rotational invariance.
1093: 
1094: %Besides the spherical symmetry the wave function has the property
1095: %that as soon as the particles in $\Psi_{1D}(I_k)$
1096: % do not cross each other along a given line the configuration remains within the
1097: %same nodal domain. 
1098: We use the properties of $1D$ wave functions and 
1099: the rotational invariance to build the proof of the two nodal domains. First,
1100: note that for four particles in $^5S(sp^3)$ state it is easy to show \cite{mitasshort}
1101: that there are only 
1102: two domains. A little bit of algebra shows that
1103:  the node is encountered when all four particles lie on a circle resulting from a
1104: plane cutting the sphere.  Clearly, it is easy to position the four particles 
1105: on the sphere so that the triple exchanges do not cross this node.  For the induction
1106: step assume that for the size $L$ with $N=(L+1)^2$ the particles are connected. We need to show
1107: that for the size with $L \to L+1$ with additional $2(L+1)+1$ particles, the additional 
1108: particles are connected as well. Using the fact that  particles can be shifted along
1109: the $\varphi$ coordinate and rotated, one can reposition the particles as illustrated
1110: on the $^{10}S(sp^3d^5)$ state,  Fig.~\ref{fig:sph}b. 
1111: By applying the factorization to this particle arrangement we see that
1112: the additional particles are connected and the argument applies to an arbitrary
1113: size closed-shell state.
1114:  
1115: %Using these results as well as the insights from the previous paper we can write
1116: %the following theorem.
1117: 
1118: \subsection{\label{sec:boxsp} III.c. Fermions in a box.}
1119: Let us assume a system of fermions in
1120: a box $(0,\pi)^d$ with the condition that the wave function vanishes
1121: at the boundaries. For $d=1$ the one-particle
1122: states are $\phi_n(x)=\sin(nx), n=1,2,... $. Note that this can be written as 
1123: $\phi_n(x)=\sin(x)U_{n-1}(\cos x)$ where $U_n$ is the $n$-th degree Chebyshev polynomial of 
1124: the second kind. 
1125: We map the variables $x_i$ to $p_i=\cos(x_i)$  so that 
1126: $p_i\in (-1,1)$ for $i=1,...,N$. Note that the map $x \to p$ 
1127: is a homeomorphism (ie, it is bijective and continuous with its inverse). 
1128: Homeomorphisms
1129:  preserve topologies, eg, 
1130: the ordering of points, so that $x_a<x_b$ $\Leftrightarrow$ $p_a<p_b$.
1131: Using this map we can
1132: write the wave function for $N$ fermions in the $1D$ box directly as 
1133: \begin{equation}
1134: \Psi_{1D}(1,...,N)={\rm det} [\phi_n(x_i)] 
1135: =\mu_0\prod_k \sin(x_k) \prod_{i>j}(p_i-p_j) 
1136: \end{equation}
1137: where $\mu_0$ is a constant.
1138: 
1139: In the $2D$ box the one-particle states are given by 
1140: $\phi_{nm}(x,y)=\sin(x)\sin(y)U_{n-1}(p)U_{m-1}(q),
1141: \; n,m=1,2, ...$ where we denoted $q=\cos(y)$.
1142: If we absorb $\sin(x)\sin(y)$ into the common prefactor $\mu$,
1143: the Slater matrix
1144:  elements become monomials of the type $p^{n-1}q^{m-1}$. 
1145: The states which are occupied lie within the quarter of 
1146: the Fermi disk $n^2+m^2 \leq k_F^2$ and $ n,m>0$. 
1147: Assume that the particles lie on a lattice in the space of variables $p,q$ 
1148: so that they are positioned 
1149: on $M$  horizontal and vertical lines (see, eg, the upper right quadrant in Fig.3c).
1150: Using the techniques presented above and in the
1151: previous paper \cite{mitasshort} we can write down the wave function as 
1152: \begin{equation}
1153: \Psi_{2D}(1,...,N)= 
1154: \nonumber
1155: \end{equation}
1156: %\pm \Psi_{2D}(1,...,N/I_k)\times
1157: %\begin{equation}
1158: %\times \Psi_{1D}(I_k)\prod_{i<j}(p_i-p_j)^{n_j}=\nonumber
1159: %\end{equation}
1160: \begin{equation}
1161: = \mu(x_1,y_1,...,x_N,y_N)\prod_{k=1}^{M-1}\left[\Psi_{1D}(I_k) \prod_{j>k} (p_k-p_j)^{n_j}\right] 
1162: \label{fig:box2d}
1163: \end{equation}
1164: where the prefactor is a nonnegative function.
1165: Similar expression can be written down by factorizing along the lines $q=q_i$.
1166: The wave function therefore
1167: factorizes in a manner very similar to  the harmonic oscillator
1168: and periodic fermion gas. Therefore the proof of the two nodal domains
1169:  above can be constructed in a similar manner.
1170: First, it is not too difficult to show that
1171: the there are only two domains in the system with three particles.
1172: Next, one assumes that this is the 
1173: case for a system with $M$ lines and then it is straightforward to show
1174:  that the same applies to the system with $M+1$ 
1175: lines.
1176: Note that Eq.\ref{fig:box2d} suggest
1177: that apparently the wave function 
1178: has  both rotation and translation invariance. This is strictly not true,
1179: due to the nonlinearity of the map $(p,q) \to (x,y)$.
1180: What is true, however, is that node is not crossed during 
1181: rotations and translations since  $(x,y)\to (p,q)$ is homeomorphic.
1182: Therefore rotations and translations 
1183: change the wave function values but not the sign since
1184: there is no reordering of either the lines or the particles along the lines.
1185: That enables us
1186: to use rotations and translations to prove the 
1187: connectedness assuming that we take additional care
1188:  to keep the paths contained 
1189: within the box. 
1190: %However, this is not a significant 
1191: %restriction. 
1192: The proof then follows similar line of arguments as presented before.
1193: It is easy to show that the three-particle state has only two nodal cells.
1194: Place the particles on a lattice and position the system into the center of the box.
1195: We assume that
1196: the lattice constant is sufficiently 
1197: small
1198: so that translations by one lattice constant 
1199: would not push the particles out of the box. 
1200: It is then straightforward to show that the four translations 
1201: $T^x_{-a}T^y_{-a}T^x_aT^y_a$
1202: exchange three particles at the edge of the system and 
1203: by similar exchanges one can show connectedness
1204: of all particles for arbitrary size. Therefore the ground state closed-shell wave functions
1205: for particles in a box have the same
1206: generic properties as the fermionic models studied in preceding sections.
1207: 
1208: \section{\label{sec:theorem}
1209:  IV. Minimal number of nodal cells theorem for spin-polarized noninteracting systems.} 
1210: Using the results and proofs derived in this section
1211: and  also in the previous paper \cite{mitasshort} we can write
1212: the following theorem.
1213: 
1214: {\it Theorem.} Consider noninteracting or mean-field spin-polarized 
1215: fermions in $d>1$  with an exact wave function given by a 
1216: Slater determinant of one-particle states times a prefactor which does affect
1217: the fermion nodes.
1218:  Let the one-particle states are  
1219: such that the Slater matrix
1220: elements can be rearranged to monomials in particle coordinates or in coordinates transformed
1221: by a homeomorphic map. Then for an arbitrary size closed-shell ground state 
1222: the corresponding 
1223: wave function has the minimal
1224: number of two nodal domains.
1225: 
1226: The theorem covers a number of paradigmatic models and it is quite suggestive to think about this 
1227: as being 
1228:  related to general mathematical properties of zeros of functions defined through
1229: determinants. In fact, the 
1230: factorizations which enabled us to prove the two nodal cells property is directly
1231: related to the properties of multiple hyperplane configurations and to the multi-variate  
1232: Vandermonde  determinant theorem which can be found in mathematical literature in
1233: various contexts  \cite{varchenko, lagr}.
1234: %alos to interacting systems. 
1235: %For example it automatically covers exactly $delta({\bf r}_i-
1236: %{\bf r}_j)$ interaction which has no effect on the nodes although it changes the 
1237: %wave function in the vicinity of the nodes. Other short-range interaction will
1238: %also leave the nodal structure intact. 
1239: %Clearly, there must be a deeper connection  
1240: %related to the many-dimensional Vandermonde  determinant theorem which can be found
1241: 
1242: Considering that we restricted the proofs to noninteracting systems and we relied
1243: on the fact that the matrix elements are monomials, it is useful to consider
1244: cases
1245: which go beyond such a framework. 
1246: This line of thought leads us to the following interesting questions:
1247: 
1248: i) Is the {\it two nodal cell property} valid for noninteracting cases with Slater 
1249: {\it matrix elements
1250:    not reducible to monomials} ?
1251: 
1252: ii) What is the {\it impact of interactions} ?
1253: 
1254: A tentative answer to the first question is given in the next section where we show that
1255: atomic states in Coulomb potential, which cannot be reduced to monomials due to the
1256: shell structure, exhibit the same property.  We will not investigate the impact of interactions
1257: for spin-polarized systems in this  
1258: paper. However, we will study and prove the two nodal cells for 
1259:  perhaps even more important cases
1260: of interacting spin-unpolarized systems in the 
1261: section VI.
1262: 
1263: \section{\label{sec:spdsp} V. Spin-polarized atomic states.} 
1264: In the previous paper \cite{mitasshort}
1265:  we have proved that the atomic
1266: spin-polarized state $1s2s2p^3$ has two nodal cells for both 
1267: noninteracting and HF wave functions. The proofs for 
1268: atomic states are more 
1269: involved  since for Coulomb potential it is more
1270: difficult to find appropriate factorizations.
1271: The main complication is that orbitals in 
1272: different subshells and 
1273: angular momentum channels have, in general, 
1274: different radial dependences which cannot
1275: be all factorized out of the determinant into a common prefactor. 
1276: Nevertheless, it is possible to demonstrate the two nodal cells for atomic 
1277: states for several spin-polarized (half-filled) main subshells 
1278: (and possibly, for all the states relevant
1279: for the periodic table of elements). We will
1280:  illustrate the idea of the proof on
1281: the spin-polarized $^{15}S(1s2s2p^33s3p^33d^5)$ state with 14 electrons 
1282: and then point out 
1283: how the proof can be extended
1284: to larger systems.
1285: The one-particle orbitals
1286: are $\rho_{1s}(r),$ $\rho_{2s}(r),$ $\rho_{2p}(r)x,$
1287: $\rho_{2p}(r)y,$ $\rho_{2p}(r)z$,...,$\rho_{3d}(r)(2z^2-x^2-y^2)$, etc,
1288: and we use dimensionless coordinates which are rescaled as
1289: ${\bf r}\leftarrow Z{\bf r}/a_0$ with $Z$ being the nuclear
1290: charge and $a_0$ the Bohr radius.
1291: The wave function is given by
1292: \begin{equation}
1293: \Psi_{at}(1,...,14)=\nonumber
1294: \end{equation}
1295: \begin{equation}
1296: {\rm det}\{\rho_{1s}^{*},\rho_{2s}^{*},x,y,z,
1297: \rho_{3s}^{*}, \phi_{3px}^{*},..., \phi_{3dz^2}^{*}, ...\}
1298: \end{equation}
1299: where $\rho_{1s}^{*}(r)=\rho_{1s}(r)/\rho_{2p}(r)$,
1300:  $\rho_{2s}^{*}(r)=\rho_{2s}(r)/\rho_{2p}(r)$ and
1301: $\phi_{3px}^{*}({\bf r})=x\rho_{3p}(r)/\rho_{2p}(r)$, ..., $\phi_{3dz^2}^{*}({\bf r})
1302: =\rho_{3d}(r)(2z^2-x^2-y^2)/\rho_{2p}(r)$, ...,
1303: etc,  where
1304: we factorized the
1305:  nonnegative radial function $\rho_{2p}(r)$
1306: out of the determinant.
1307: 
1308: We will show the connectedness of all the particles in two steps. 
1309: We first demonstrate that the Slater determinant can be factorized 
1310: into subdeterminants corresponding to subshells $1s$,  
1311: $2s2p^3$ and $3s3p^33d^5$. That will enable us to show
1312: that the particles within each subshell
1313: are connected. In the second step we show, that the particles can exchange
1314: between the subshells.
1315: For this purpose we will specify the explicit paths
1316: and carry out a straightforward numerical check that there is no node crossing
1317: by  tracing the wave function along
1318: the paths. 
1319: 
1320: In order to show the factorization into shells
1321: we position the particles as follows: particle 1 
1322: is in the origin, particles 2 to 5 are on the surface
1323: of a sphere with the
1324: radius $\eta_a$ and particles 6 to 14 are 
1325: are on the surface
1326: of a sphere with the radius $\eta_b$. 
1327: The radii $\eta_a$ and $\eta_b$ are given by the radial nodes of 
1328: the orbitals $\rho_{2s}(r)$ and $\rho_{3p}(r)$, 
1329: ie, $\rho_{2s}(\eta_a)=0$ and  $\rho_{3p}(\eta_b)=0$. 
1330: The Slater determinant can be written in the form
1331: \begin{equation}
1332: \Psi_{at}(1,...,14)=
1333: {\rm det}
1334: \left[ \begin{array}{ccc}
1335: {\bf A } & {\bf D } &{\bf  G} \\
1336: {\bf B } & {\bf E' } & {\bf E}\\
1337: {\bf C } & {\bf F } & {\bf  H}\\
1338: \end{array} \right] 
1339: \label{eq:spd}
1340: \end{equation}
1341: where the block matrices ${\bf A}$ to ${\bf H}$ are given as follows
1342: \begin{equation}
1343: {\bf A}=\left[ \begin{array}{ccccc}
1344: \rho_{1s}^{*}(0)   &  a_1       &  a_1   & a_1  & a_1 \\
1345: \rho_{2s}^{*}(0)   &  0       &  0   & 0  & 0 \\
1346: 0 &  x_2      & x_3 & x_4  &  x_5 \\
1347: 0 &  y_2     & y_3  & y_4  &  y_5 \\
1348: 0 &  z_2     & z_3 & z_4 &  z_5 \\
1349: \end{array} \right]  
1350: \end{equation} 
1351: \begin{equation}
1352: {\bf B}=\left[ \begin{array}{ccccc}
1353: \rho_{3s}^{*}(0)   &  a_3       &  a_3   & a_3  & a_3   \\
1354: 0 &  \phi_{3px}^*(2)      &  ...  & 
1355:  &  \phi_{3px}^*(5) \\
1356: 0 &  \phi_{3py}^*(2)      & ...  &  &  \phi_{3py}^*(5)\\
1357: 0 &  \phi_{3pz}^*(2)       &  &  &  \phi_{3pz}^*(5) \\
1358: \end{array} \right]    
1359: \end{equation}   
1360: \begin{equation}
1361: {\bf C}=\left[ \begin{array}{ccccc}
1362: 0 &  \phi_{3dz^2}^*(2)      &  ...  & &  \phi_{3dz^2}^*(5) \\
1363: 0 &  \phi_{3dx^2}^*(2)      & ...  &  &  \phi_{3dx^2}^*(5)\\     
1364: 0 &  \phi_{3dxy}^*(2)       & ...  &  &  \phi_{3dxy}^*(5) \\
1365: 0 &  \phi_{3dyz}^*(2)       & ... &  &  \phi_{3dyz}^*(5) \\
1366: 0 &  \phi_{3dxz}^*(2)       & ... &  &  \phi_{3dxz}^*(5) \\
1367: \end{array} \right]       
1368: \end{equation}   
1369: \begin{equation}      
1370: {\bf D}=\left[ \begin{array}{cccc}
1371:  b_1       &  b_1   & b_1  & b_1   \\
1372:  b_2       &  b_2   & b_2  & b_2   \\
1373: x_6 &  ...        & ...  &  x_9 \\
1374: y_6 &  ...        & ...  &  y_9 \\
1375: z_6 &  ...        & ...  &  z_9 \\
1376: \end{array} \right]
1377: \end{equation}
1378: \begin{equation}
1379: {\bf E}=\left[ \begin{array}{ccccc}
1380:  b_3       &  b_3   & b_3  & b_3 & b_3  \\
1381: 0 &  ...        &  & ... &  0 \\ 
1382: 0 &  ...        &  & ... &  0 \\   
1383: 0 &  ...        &  & ... &  0 \\
1384: \end{array} \right] 
1385: \end{equation} 
1386: \begin{equation}
1387: {\bf F}=\left[ \begin{array}{cccc}
1388:   \phi_{3dz^2}^*(6)      &  ...  & &  \phi_{3dz^2}^*(9) \\
1389:   \phi_{3dx^2}^*(6)      & ...  &  &  \phi_{3dx^2}^*(9)\\
1390:   \phi_{3dxy}^*(6)       & ...  &  &  \phi_{3dxy}^*(9) \\
1391:   \phi_{3dyz}^*(6)       & ... &  &  \phi_{3dyz}^*(9) \\
1392:   \phi_{3dxz}^*(6)       & ... &  &  \phi_{3dxz}^*(9) \\
1393: \end{array} \right]
1394: \end{equation}
1395: \begin{equation}          
1396: {\bf G}=\left[ \begin{array}{ccccc}
1397:  b_1       &  b_1   & b_1  & b_1 & b_1  \\
1398:  b_2       &  b_2   & b_2  & b_2 & b_2  \\
1399: x_{10} &  ...        & ...  &  ... & x_{14}\\
1400: y_{10} &  ...        & ...  &  ... & y_{14}\\
1401: z_{10} &  ...        & ...  &  ... & z_{14}\\
1402: \end{array} \right]
1403: \end{equation}
1404: \begin{equation}
1405: {\bf H}=\left[ \begin{array}{ccccc}
1406:   \phi_{3dz^2}^*(10)      &  ...  & &  &  \phi_{3dz^2}^*(14) \\
1407:   \phi_{3dx^2}^*(10)      & ...  &  & &  \phi_{3dx^2}^*(14)\\
1408:   \phi_{3dxy}^*(10)       & ...  & &  &  \phi_{3dxy}^*(14) \\
1409:   \phi_{3dyz}^*(10)       & ... &  & &  \phi_{3dyz}^*(14) \\
1410:   \phi_{3dxz}^*(10)       & ... &  & &  \phi_{3dxz}^*(14) \\
1411: \end{array} \right]
1412: \end{equation}
1413: and $a_1=\rho_{1s}^{*}(\eta_a)$, $a_3=\rho_{3s}^{*}(\eta_a)$,
1414: $b_1=\rho_{1s}^{*}(\eta_b)$, $b_2=\rho_{2s}^{*}(\eta_b)$, $b_3=\rho_{3s}^{*}(\eta_b)$.
1415: The block matrix ${\bf E'}$ is the same as  ${\bf E}$ except that it has only four columns 
1416: instead of five. 
1417: 
1418: The following three row additions 
1419: in the Slater matrix given by Eq.\ref{eq:spd} 
1420: allow for factorization into subdeterminants.
1421: We first eliminate elements $a_1$ in the first row   
1422: by adding an appropriate multiple of the fifth row. Similarly, 
1423: we eliminate the terms with $b_1$ in the first row  
1424: by adding a multiple of the second row. This
1425: leads to the first row having only one nonzero element
1426: $A'_{11}=\rho_{1s}^{*}(0)$
1427: $-a_1\rho_{3s}^{*}(0)/a_3$
1428: $-(b_1-a_1b_3/a_3)\rho_{2s}^{*}(0)/b_2$ and reduces the Slater matrix
1429: by the first row and the first column.
1430: Finally,
1431:  by adding a multiple of the second row to the fifth row, 
1432: both ${\bf E'}$ and ${\bf E}$ become zero matrices and we can write
1433: \begin{equation}
1434: \Psi_{at}(1,...,14)=A'_{11}\, {\rm det} {\bf B'}\,
1435: {\rm det}
1436: \left[ \begin{array}{cc}
1437:  {\bf D' } &{\bf  G'} \\
1438: {\bf F } & {\bf  H}\\
1439: \end{array} \right] 
1440: \end{equation}
1441: where  ${\bf B'}$ is the matrix ${\bf B}$ without the first column,
1442: while ${\bf D' }, {\bf  G'}$ are the matrices ${\bf D}, {\bf  G}$ without the first row.
1443: The three factors in the expression above correspond to the three main 
1444: subshells namely, $1s$, $2s2p^3$ and $3s3p^33d^5$ with the dependence
1445: on positions of particles 1, 2-5 and 6-14,
1446: respectively.
1447: %Note that so far we have not specified the actual positions in these subshells except that they
1448: %are
1449: %on the same spherical surface: the $1s$ particle is trivially in the origin, 
1450: %$2s2p^3$ on the sphere with
1451: %radius $r=\eta_a$ 
1452: %and  $3s3p^33d^5$ on the sphere with $r=\eta_b$. 
1453: The key point is that each of the subshells represents a system of particles on a sphere in
1454: the $S$-symmetry state and for such cases 
1455: we proved that the particles are connected.
1456: That concludes the argument that the particles within each subshell are connected.
1457: 
1458: \begin{figure}[ht]
1459: \centering
1460: \includegraphics[width=2.0in,clip]{3spdCsphere.eps} 
1461: \caption{Positions of 14 electrons for the proof that triple exchanges connect the particles 
1462: in the $^{15}S(1s2s2p^33s3p^33d^5)$  state.
1463: Particles 2 to 5 lie on the spherical surface with 
1464: the radius equal to the radial node of $\rho_{2s}(r)$ orbital.
1465: Particles 6 to 14 lie on the spherical surface with the radius equal to the radial node of $\rho_{2p}(r)$.}
1466: \label{fig:3spd}
1467: \end{figure}
1468: 
1469: 
1470: More difficult part of the proof is to show that one can exchange the particles between the subshells.
1471: We were able to prove analytically the exchange between $1s$ and $2s2p^3$ in our previous paper. For
1472: larger cases the analytic proof becomes very tedious and it is much more efficient
1473: to evaluate numerically the determinant along the following 
1474: exchange paths.
1475: The particles are positioned as sketched in Fig.\ref{fig:3spd}
1476: with the coordinates given by
1477: ${\bf r}_1=(0,0,0)$, ${\bf r}_2=(\eta_a,0,0)$, ${\bf r}_2=(0,\eta_a,0)$,
1478: ${\bf r}_{4,5}=(0,0,\pm \eta_a)$, ${\bf r}_6=(\eta_b/\sqrt{2},\eta_b/\sqrt{2},0)$,
1479: ${\bf r}_{7,8}=(\pm\eta_b/\sqrt{2},\mp\eta_b/\sqrt{2},0)$, ${\bf r}_{9,10}=(0,0,\pm\eta_b)$,
1480: ${\bf r}_{11,12}=(\eta_b/\sqrt{2},0,\pm\eta_b/\sqrt{2})$ and ${\bf r}_{13,14}=(0,\eta_b/\sqrt{2},\pm\eta_b/\sqrt{2})$.
1481: There exists enough of triple 
1482: exchanges of neighbouring particles, along the sides of corresponding triangles, to connect 
1483: all the particles into a single cluster. 
1484: For example,
1485: the wave function values for exchanges $123\to 231$, $236\to 362$ (and other exchanges for 
1486: illustration) are given in Fig.~\ref{fig:atexch}.
1487: 
1488: \begin{figure}[ht]
1489: \centering
1490: \includegraphics[width=2.4in,clip]{atexch.eps}
1491: \caption{
1492: Wave function of the $^{15}S(1s2s2p^33s3p^33d^5)$ 
1493: state (arb. u.) along triple exchange paths parametrized by $t$. Particles are positioned 
1494: as given in Fig.~\ref{fig:3spd}. The plotted 
1495: exchanges:  1,2,3 (full line); 2,3,4 (dashed line); 2,3,6 (dotted line); 6,7,9 (dashed-dotted line); 
1496: 9,11,13 (double dashed-dotted line). By symmetry, these non-crossing exchanges connect all the particles of this 
1497: state.}  
1498: \label{fig:atexch}
1499: \end{figure}
1500: 
1501: This is enough to show that the three subshells are connected
1502: since many other exchanges are identical due to the symmetries in particle positions 
1503: and the wave function $S$ symmetry. 
1504: %Many other exchanges are possible. 
1505: For the illustration in Fig.~\ref{fig:atexch} we used the noninteracting radial orbitals.
1506:  For HF orbitals one gets the same qualitative picture since
1507: the basic spatial properties of the noninteracting
1508: and HF orbitals are qualitatively the same 
1509: (ie, ordering of radial nodes of $\rho_{nl}(r)$ orbitals,
1510: behaviour at nucleus, tails,
1511:  etc). The proof uses only the fact that some of the radial nodes of the one-particle orbitals
1512: are ordered as in the non-interacting case so it can be extended to HF wave functions as well. 
1513: %In addition, for large $Z$ the
1514: %one-particle states approach the noninteracting ones since the impact of 
1515: %electron-electron interactions is decreasing and orbitals become even more closer to the
1516: %noninteracting ones.
1517: 
1518: One can expand the proof to larger systems, such as for the state with
1519: occupied fourth main subshell  
1520:  $4s4p^34d^54f^7$ and beyond. The factorization is similar: the particles in the
1521: fourth subshell are positioned on the spherical surface with the radius
1522: equal to the radial node of $\rho_{4d}(r)$ orbital. Somewhat long
1523: but straightforward
1524: algebraic rearrangements show
1525: that the Slater determinant of the $30\times 30$ matrix can be 
1526: reduced to product of subdeterminants corresponding to the subshells.
1527: For the purpose of this paper we do not deem necessary to go through the explicit proof.
1528: 
1529: 
1530: \section{\label{sec:hegsu} VI. Interacting spin-unpolarized fermions.}
1531: 
1532: It is straightforward to understand 
1533: the fermion nodes of noninteracting
1534: spin-unpolarized or partially polarized systems with more than one-electron in each
1535: spin-channel.
1536: The wave function is a product of spin-up and -down Slater determinants and the 
1537: number of nodal cells is the product of the number of cells in each subspace.
1538: (As mentioned before, we assume that the Hamiltonian, even with interactions considered
1539: below, does not include terms with spin so that particles can be assigned 
1540: to the spin subspaces.) For the ground
1541: states with two nodal cells in each subspace we get $2\times2=4$ nodal cells.
1542: This is the nodal structure of a single configuration Hartree-Fock wave function
1543: and  as such has been used in many 
1544: fixed-node QMC calculations. However, analyses of small interacting systems
1545: revealed that this is
1546: not correct and interactions do change the nodal topologies
1547: and the number of nodal cells. 
1548: For the first time 
1549: this has been demonstrated 
1550: for the Be atom \cite{dariobe}
1551: and then also for a few other atoms and small molecules \cite{cmt28,bajdichpfaffians}.
1552: 
1553: In the previous paper \cite{mitasshort} we have outlined a proof 
1554: showing the two nodal domains property for $2D$
1555: harmonic fermions in a closed-shell singlet state for arbitrary size. The most interesting feature
1556: of the proof was that the result was very 
1557: robust in the sense that almost any arbitrary weak interaction would induce the change in the 
1558: topology of the nodal surfaces. 
1559: 
1560: We
1561: refresh some of the key notions and arguments here. 
1562: Let us assume a spin-unpolarized system in its
1563: closed-shell singlet ground state of $2N$ particles.
1564: Consider a simultaneous exchange of an odd number of spin-up pair(s)
1565: {\it and} an odd number of spin-down pair(s).
1566: For noninteracting wave functions such simultaneous pair exchanges
1567: imply that the node will be crossed once or multiple times. This must be the case
1568: whenever the spin-up and -down subspaces are independent of each other, such as 
1569: in the mean-field and HF wave functions. However,
1570: if there exists a point $R_f$ such that during the simultaneous
1571: spin-up and -down pair exchanges the inequality
1572:    $|\Psi|>0$ holds along the whole path,
1573:  then the wave function has only two nodal cells.
1574: 
1575: Using this property we can now demonstrate the two nodal cell for several types of systems.
1576: %Consider a singlet ground state of $2N$ interacting particles.
1577: Under rather general conditions, 
1578:  we will show that the correlation included in
1579: the Bardeen-Cooper-Schrieffer (BCS) pairing
1580: wave function \cite{sorella,carlson} given by
1581: \begin{equation}
1582: \Psi_{BCS}(1,...,2N)={\rm det} [\Phi(i,j)] 
1583: \label{eq:bcs}
1584: \end{equation}
1585: smoothes out 
1586:  the noninteracting four nodal cells
1587: into the minimal two. 
1588: Here $\Phi(i,j)=\Phi(j,i)$ is a singlet pair orbital for
1589: $i\negthinspace\negthinspace\uparrow $ and
1590: $j\negthinspace\negthinspace\downarrow$ fermions and we decompose it
1591: into noninteracting and correlated components
1592: $\Phi(i,j)= \Phi_{0}(i,j) + \Phi_{corr}(i,j)$. Using one-particle orbitals
1593: we can write 
1594: $\Phi_{0}(i,j)=\sum_{\alpha}\negthinspace \phi_{\alpha}(i)\phi_{\alpha}(j)$ where
1595: the sum is over HF (or noninteracting) orbitals 
1596: while $ \Phi_{corr}(i,j)=\sum_{\alpha\beta}c_{\alpha\beta}\phi_{\alpha}(i)\phi_{\beta}(j)$
1597: where $\{c_{\alpha\beta}=c_{\beta\alpha}\}$ are variational parameters and 
1598: the sum is over unoccupied or virtual ("correlating") 
1599: orbitals. The BCS wave function was originally introduced for conventional
1600: superconductors, however, it proved to be very successful also in describing 
1601: the correlation effects in electronic structure problems \cite{sorella,bajdichpfaffians}.   
1602: %appropriate expansion coefficients. 
1603: 
1604: \subsection{VI.a. $3D$ harmonic spin-unpolarized fermions with interactions.}
1605: 
1606: For the $3D$ harmonic oscillator one can show that the interactions lead to the
1607: minimal number of nodal cells in a rather simple and elegant way.
1608:   First let us consider
1609: a small system which is easy to analyze. We illustrate
1610: the idea on $2N=8$ particles in the singlet
1611: ground state with the particle positions given in Fig.~\ref{fig:3dho}. 
1612: The corresponding closed-shell
1613: singlet ground state for 
1614: the $3D$ harmonic oscillator is $^1 S(1s^22p^6)$. (Note that for the harmonic potential
1615: the $2p$ state is below the $2s$ state,
1616: unlike for the Coulomb potential.)
1617: For simplicity, we drop the gaussian prefactors since they do not affect
1618: the nodes and 
1619: the pairing 
1620: functions can be then written
1621: as follows
1622: \begin{equation}
1623:  \Phi_{0}(i,j) = 1 + 3{\bf r}_i\cdot {\bf r}_j
1624: \end{equation}
1625: and 
1626: \begin{equation}
1627:  \Phi_{corr}(i,j) = \gamma [3({\bf r}_i\cdot {\bf r}_j)^2 -r_i^2r_j^2]
1628: \end{equation}
1629: where the noninteracting part is constructed from the occupied $1s,2p$
1630:  orbitals while for the correlating 
1631: part the unoccupied $3d$ subshell orbitals were used. 
1632: The constant $\gamma$ is a variational parameter.
1633: Clearly, both the pair orbitals
1634: and the wave function are spherically symmetric.
1635: Assuming the positions as given in the Fig.~\ref{fig:3dho},  which are of high symmetry
1636: and therefore simplify the evaluation of the wave function, we get 
1637: \begin{equation}
1638: \Psi(1,...,8)= 4\gamma\cos\varphi[1-(3+\gamma)\cos^2\varphi]/3
1639: \end{equation}
1640: where $r_1=r_5=0$,  $\; r_i=1, i=2,3,4,6,7,8$
1641: and $\cos\varphi={\bf r}_4\cdot {\bf r_8}$.
1642: 
1643: \begin{figure}[ht]
1644: \centering
1645: \includegraphics[width=1.8in,clip]{3dHO.eps}
1646: \caption{Positions of $2N=8$ fermions of the $3D$ harmonic well system. Fermions 1 to 4 have spins up,
1647: while 5 to 8 have spins down. Fermions 1 and 5 are at the origin while the rest of the particles
1648: lie on the surface of the sphere with $r=1$. Note that rotation of the system
1649: by $\pi$ around the $z-$axis exchanges particles $3,4$ and $7,8$.
1650: See the text for details. }
1651: \label{fig:3dho}
1652: \end{figure}
1653: 
1654: For $\gamma =0$ the wave function vanishes since the particles in both
1655: spin subspaces lie on the noninteracting node (ie, four particles on the same plane). 
1656: The key point is that for
1657: $\gamma\neq 0$ and $\varphi\neq \pi/2$ the wave function
1658: does not vanish.
1659: Since the wave function is spherically symmetric we can rotate the particles 
1660: around the $z-$axis by $\pi$. This transformation 
1661: exchanges particles 3 and 4 in the spin-up 
1662: channel and particles 7 and 8 in the spin-down channel. 
1663:  The wave function is rotationally
1664: invariant and nonvanishing what clearly implies that the BCS wave function has
1665: spin-up and -down subspaces interconnected since simultaneous exchanges in the spin-up  
1666: and spin-down {\it does not } hit the node. 
1667: %This fulfills the condition for 
1668: %the two nodal cells wave functions.  
1669: On the other hand, it is easy to check
1670: that the rotation of the particles in one spin channel only causes a node-crossing
1671: since then $\cos\varphi$ vanishes either at $\pi/2$ or $3\pi/2$. 
1672: %Note that we never specified the interaction and using the BCS variational wave
1673: %function we tacitly assume that it is accurate enough to describe the key 
1674: %effect from the interactions. 
1675: 
1676: It is clear that our argument is correct
1677: whenever the strength of the interaction
1678: is small so that the BCS wave function is accurate enough. 
1679: The variational 
1680: parameter $\gamma$ is related to the interaction strength. This
1681:  implies that the topological 
1682: change from two to four nodal cells takes place for arbitrary small, but 
1683: nonvanishing, interaction strength.
1684: The only assumptions were
1685: that the interaction will induce electron correlation and lead to a nonzero
1686: variational parameter $\gamma$ (or, in general, to a nonvanishing correlation
1687: component) and that the wave function is spherically
1688: symmetric. Our analytic argument therefore shows
1689: that for weak
1690: interactions the nodal cell "degeneracy" is lifted
1691: and the multiple cells are smoothed to the minimal
1692: number of two. 
1693: 
1694: It is important to note that
1695: one can find different pictures in other circumstances. For example,
1696:  for strong or nonlocal interactions, imposed additional symmetries or
1697: large degeneracies, the nodal changes can be different and the resulting
1698: nodal topology might exhibit more than the two nodal cells;
1699: this aspect is further
1700: discussed in the conclusion section and require more investigation. 
1701: 
1702: Coming back to our example of
1703: harmonic fermions with interactions, using similar arguments 
1704: the two nodal cells can be demonstrated for
1705: an arbitrary size closed-shell ground state. 
1706: Consider $2N$ particles in a singlet closed-shell ground state.
1707: The closed-shell states can be labeled by $M$ which represents the "Fermi momentum"
1708: for the harmonic oscillator and $N=(M+1)(M+2)(M+3)/6$.  Let us
1709: decompose $N$ into the maximum 
1710: odd number of 
1711: pairs $N_P$ and the rest. Therefore we write  
1712: $N=2N_P+J$ where $N_P$ is an odd integer while $J$ has one of the values $J=0,1,2,3$, so 
1713: that we can also define $J=N\, {\it mod}\, 4$.
1714: For example, in
1715: the previous example with $2N=8$ particles we have $N_P=1$ and $J=2$. First, we place $J$
1716: particles from each spin channel on the $z$-axis in
1717:   distinct positions.  Next,
1718: we form $N_P$ pairs 
1719:  in each spin subspace
1720: so that, say, the pair $i\negthinspace\uparrow,
1721: (i+1)\negthinspace\uparrow $ is positioned as  
1722: given by $(r_i,\vartheta_i,\varphi_i)=(r_k,\vartheta_k,\varphi_k)$,
1723:  $\,(r_{i+1},\vartheta_{i+1},\varphi_{i+1})=(r_k,\vartheta_k,\varphi_k+\pi)$,
1724: where $r,\vartheta,\varphi$ are the spherical coordinates.
1725: Here $k=1,...,N_P$ labels the pairs
1726: in the spin-up channel; the spin-down particles are 
1727: placed similarly and labeled by the pair index $l=N_p+1,...,2N_P$.
1728: In this configuration the particles lie on the 
1729: noninteracting node since 
1730: \begin{equation}
1731:  {\rm det}\left[\Phi_0(i,j)\right]={\rm det}\left[\sum_{n+m\leq M} \phi_{nm}(i)
1732:  \phi_{nm}(j)\right]=
1733: \nonumber
1734: \end{equation}
1735: \begin{equation}
1736: ={\rm det}\left[\phi_{nm}(i)\right]
1737: {\rm det}\left[\phi_{nm}(j)\right]=\Psi_{HF}^{\uparrow}\Psi_{HF}^{\downarrow}
1738: \end{equation}
1739: The rotation of the system by $\pi$ causes node-crossings 
1740: in both spin channels
1741: so that both spin-up and -down Slater determinants must  vanish due to the rotational
1742:  invariance.
1743: Now, if all the $N_P$ pair distances and angles
1744: $r_k,\varphi_k,\vartheta_k,r_l,\vartheta_l,\varphi_l$ are distinct, 
1745: then each of the matrices $\left[\phi_{nm}(i)\right], \left[\phi_{nm}(j)\right]$
1746: has exactly one linearly dependent row, ie, both have ranks $N_P-1$.
1747: This can be verified directly
1748: for small values of $M$ and then using induction for any $M$.
1749: Consequently, the matrix $\left[\Phi_0(i,j)\right]$ has linear dependence
1750: in one row and one column, ie, it has the rank of $N_P-1$ as well. In general,
1751: adding virtual states through $\Phi_{corr}(i,j)$ provides
1752: independent rows/columns
1753: %(eg, $(M+1)(M+2)/2$ independent rows/columns from the first unoccupied shell)
1754: which eliminate the linear dependency so
1755: that ${\rm det}\left[\Phi_0(i,j)+\Phi_{corr}(i,j)\right]$ is nonzero. Let us now assume 
1756: that the interactions do not break the rotation invariance. Since at this point the
1757: wave function does not vanish and is rotationally invariant,  
1758: it does not vanish for the whole exchange path implying that the correlated 
1759: BCS wave function has only two nodal cells regardless of the size $2N$.
1760: 
1761:  Clearly, the assumption of the rotational invariance for the interacting
1762: case might be too
1763: restrictive since the interactions could possibly break the spherical symmetry.
1764: To demonstrate  the two nodal cells in such a case one would need to show
1765:  that the wave function does not vanish for the complete exchange path.
1766: % small change in a variable makes it nonvanishing.)
1767: 
1768: \subsection{VI.b. $2D$ harmonic spin-unpolarized fermions with interactions for 
1769: $N=2N_P+J$ when $J=2,3$.}
1770: 
1771: In our previous paper we have demonstrated
1772: the two nodal cells for the
1773: ground states with 
1774: the number of particles $2N$ where $N=2N_P+J$ and $J=0,1$. (We use the 
1775: same notation as in $3D$ where $N_P$ is the maximum odd number of pairs and $J=0,1,2$ or 3).
1776: Note that the states with $J=0$ or 1 exist at any size. Here we would like to present 
1777: an alternative
1778: and a little bit longer
1779: proof which covers $J=0,1$ but also remaining cases
1780:  when $J=2,3$, eg, 
1781: for $N=21=2\times9 +3,\;$ $N=28=2\times13 +2$, etc.
1782: In $3D$, the rotation axis can accommodate multiple particles so that one can always
1783: form an odd number of pairs for the exchanges. In $2D$, the
1784: symmetry point (origin) can accommodate only one particle   
1785: from each spin subspace therefore the proof has to be modified.
1786: For this purpose we outline a factorization
1787: which is different from the previous paper \cite{mitasshort}. Let us remind that for $2D$
1788: harmonic oscillator the closed-shell states and the system size are labeled by
1789: $M=1,2, ...$ where $n+m\le M$ with the number of fermions in one spin
1790: channel given by $N=(M+1)(M+2)/2$.  We express the one-particle states 
1791: as polynomials in $r^2$ and $(re^{i\varphi})^m$  where $r, \varphi$ are 
1792: the cylindric coordinates (we again omit the gaussian factors
1793: which are irrelevant for the nodes).  Note that for a given $M$ the 
1794: quantum number $m$ increases or decreases by 2 unlike
1795: in $3D$. Therefore for $M=2$ we have states $re^{\pm i\varphi}$, for $M=3$ the 
1796: additional states are $(r^2-1), r^2e^{i\pm 2\varphi}$, etc. 
1797: We first write down the wave functions for three- and six-particle 
1798: spin-polarized systems with positions given in Fig. \ref{fig:2dho}.
1799: 
1800: \begin{figure}[ht]
1801: \centering
1802: \begin{tabular}{ccc}
1803: \includegraphics[width=1.4in,clip]{figYa.eps} & \quad &
1804: \includegraphics[width=1.4in,clip]{figYb.eps} \\
1805: (a) & \quad & (b) \\
1806:  \quad & \quad & \quad \\
1807:  \quad & \quad & \quad \\
1808: \includegraphics[width=1.4in,clip]{figYc.eps} & \quad &
1809: \includegraphics[width=1.4in,clip]{figYd.eps} \\
1810: (c) & \quad & (d) \\
1811: \end{tabular}
1812: \caption{Alternative positions of three and six spin-polarized fermions for the $2D$ harmonic
1813: potential.
1814: (a) Three particles positioned on the noninteracting node.
1815: (b) Three particles in positions with a nonvanishing wave function.
1816: (c) Six particles positioned on the noninteracting node.
1817: (d) Six particles in positions with a nonvanishing wave function.
1818: See the text for details. }
1819: \label{fig:2dho}
1820: \end{figure}
1821: 
1822: The three particle wave function is given by ${\rm det}\{1,re^{\pm i\varphi}\}$
1823: \begin{equation}
1824: \Psi_{2D}(1,2,3)
1825: =\mu_0r_0^2\prod_{1(2)\leq j <k \leq 3}\sin(\varphi_{jk}/2)
1826: \end{equation}
1827: where $\mu_0$ is a constant.
1828: The product lower bound is 2 if the positions of the particles 
1829: are  as in Fig.~\ref{fig:2dho}a, while it is 1 if the positions  
1830: are as in Fig.~\ref{fig:2dho}b.
1831: Note that in the case of configuration in Fig.~\ref{fig:2dho}a 
1832: rotation by $\pi$
1833: flips the wave function sign so that the wave function vanishes
1834: while for the configuration in  Fig.~\ref{fig:2dho}b it does not.
1835:  For six particles in the positions outlined in  Fig.~\ref{fig:2dho}d the
1836: wave function is clearly nonvanishing since  we can write 
1837: \begin{equation}
1838: \Psi_{2D}(1,...,6)
1839: =\mu_0(r_2^2-r_1^2)r_2^6\prod_{j<k}^6\sin(\varphi_{jk}/2)
1840: \end{equation}
1841: where $\mu_0$ is some constant (in Fig.~\ref{fig:2dho}d we have
1842:  $r_1=0$).
1843: Alternatively,
1844: the six particles can be positioned in pairs as sketched 
1845:  in Fig.~\ref{fig:2dho}c where the particles are positioned 
1846: on the HF node so that the noninteracting wave function vanishes.
1847: 
1848: For the sake of completeness we provide the expression for the wave function for
1849: a general system
1850: with size $M$  
1851: \begin{equation}
1852: \Psi_{2D}(1,...,N) =\nonumber
1853: \end{equation}
1854: \begin{equation}
1855: =\mu(r_1, ..., r_{\tilde M})\prod_{k=1}^{\tilde M}\left[
1856: \Psi_{1D}(I_k)\prod_{1\leq j<k}(r^2_k-r^2_j)^{n_j}\right]
1857: \label{eq:2dho}
1858: \end{equation}
1859: where $\tilde M= {\rm int}[(M+1)/2]$ is the integer part of $(M+1)/2$. The particles with
1860: indices in the subset $I_k$ lie on a circle with the
1861: radius $r_k$ and $n_k=2(M+1-2k)+1$.  The prefactor is again a nonnegative function
1862: of  
1863:  $r_j$ which has 
1864: no impact on the nodes
1865: and is constant for rotations. From the last equation we see that in cases when $M$ is even, eg,
1866: $M=6, N=21$, we end up with factorization of the type $11\times 7 \times 3$ particles 
1867: while for $M$ odd, eg, $M=7,N=28$, we get $13\times 9\times 5 \times 1$. That means that the last
1868: three particles ($M$ even) or the last six particles ($M$ odd) can be always arranged into possibilities 
1869: as sketched in
1870: Fig~\ref{fig:2dho}a-d. 
1871: %The factorization given by Eq.~\ref{eq:2dho} provides an 
1872: %alternative to the expression derived before \cite{mitasshort} where we used cartesian
1873: %coordinates.
1874: It is now clear that we can prove that
1875: the spin-up and -down configuration space are interconnected
1876: for the cases when $J=2,3$ as specified above. 
1877: Using the alternative configurations for three and six particles we can arrange 
1878: the systems into such 
1879: positions that the rotation by $\pi$ exchanges odd number of pairs in both spin channels.
1880: For example, if $N$ is
1881: odd and $(N-1)/2$ is even
1882: %, instead of locating one of the particles of each spin into 
1883: %the origin as we did in the previous paper \cite{mitasshort}, we rather
1884: we position three particles on a triangle as given in Fig.~\ref{fig:2dho}b. 
1885: That takes out one pair from
1886: the $(N-1)/2$ exchanges and the proof then follows using the arguments for $3D$ harmonic fermions.
1887: Similarly, if  $N$ is even and $N/2$ is also even,
1888: we position one particle of each spin at the origin and 
1889: five from each spin channel on a circle. This eliminates three pairs from $N/2$ and again we end 
1890: up with remaining number of pairs being odd so that the rest of the proof follows similarly to the
1891: $3D$ case.
1892: 
1893: \subsection{VI.c. Spin unpolarized interacting $2D$ and $3D$ homogeneous electron gas.}
1894: 
1895: Let us now prove the two nodal cells for the spin-unpolarized closed shells
1896: for $d>1$ homogeneous  gas with interactions. We use the relevant definitions
1897: introduced  previously for the spin-polarized periodic fermion gas (section III.a.).
1898: We assume a $2D$ system of $2N$ particles in the spin singlet ground state
1899: where the one-particle states occupy the Bloch states within the Fermi disk with the
1900: radius 
1901:  $k_F$. Let us first consider $2N=10$ particle 
1902: case and we position the particles 
1903: as in Fig.  \ref{fig:2dhegunpol}. 
1904: The wave function is translationally invariant therefore the action of
1905: $T^{y}_{\pi}$ leaves the wave function unchanged. 
1906:  Assuming first that there is no interaction, we see that the
1907:  translation  exchanges the particles $1,2$ but due to the invariance 
1908: the wave function value does not change: that is possible only if the particles
1909: are sitting on the node.
1910: This also agrees with expressions for the wave functions derived before  
1911: (see Eq. \ref{eq:five2dxi}.) Consider now that we switch-on interactions
1912: and describe the particles by the BCS wave function as given above. 
1913: The pairing orbital for the noninteracting gas is given by
1914: \begin{equation}
1915: \Phi_0(i,j)=\sum_{|{\bf k}|\leq k_F} \phi_{\bf k}(i)\phi_{\bf -k}(j)=1+2[\cos(x_{ij})+\cos(y_{ij})]
1916: \end{equation}
1917: where we used the occupied orbitals $\{1,e^{\pm ix},e^{\pm iy}\}$.
1918: The correlated part is given by
1919: \begin{equation}
1920: \Phi_{corr}(i,j)=
1921: \gamma\sum_{|{\bf k}|> k_F}\phi_{\bf k}(i)\phi_{\bf -k}(j)=
1922: \nonumber
1923: \end{equation}
1924: \begin{equation}
1925: \gamma[\cos(x_{ij})\cos(y_{ij})+\cos(2x_{ij})+\cos(2y_{ij})]
1926: \label{eq:2dcorr}
1927: \end{equation}
1928: where $\Phi_{corr}(i,j)$ includes the next two unoccupied stars of states with $k=\sqrt{2},2$ 
1929: and $\gamma$ is a variational parameter. 
1930: %We position the particles as sketched in Fig. \ref{fig:2dhegunpol}.
1931: %Consider now the translation $T^y_{\pi}$. If the particles 3-5 are positioned in such a way that
1932: %after translation they reach positions related to the original ones by a mirror symmetry around $y$-axis
1933: %% then it is clear that the value of the wave function has to be the same. At the same time the positions
1934: %of particles 4 and 5 are {\it exchanged}. This means that the only way how this would be the case is that
1935: %the wave function is zero. 
1936: 
1937: \begin{figure}[ht]
1938: \centering
1939: \includegraphics[width=1.5in,clip]{figZ2.eps}
1940: \caption{Positions of five spin-up electrons from an interacting ten-fermion system
1941: in the $2D$ periodic box $(-\pi,\pi)^2$.
1942: For clarity, the five down-spin particles  
1943: are not shown. The spin-down particles are positioned in the same pattern
1944: on two lines $x=const$.}
1945: %$\xi_0,\xi_1$. 
1946: %See the text for more details. }
1947: \label{fig:2dhegunpol}
1948: \end{figure}
1949: 
1950: Unlike for the noninteracting wave function (ie, $\gamma=0$) 
1951: the BCS correlated wave function does not vanish and this is indeed the case due to the 
1952: arguments presented above. The Slater 
1953: matrix of uncorrelated wave function has linear dependency while for the correlated case
1954: there always exists a configuration of lines with nonvanishing wave function. This is
1955: due to the elimination of the linear dependency through addition of
1956: virtual orbitals, as explained
1957: for the $3D$ harmonic oscillator.
1958: Note that as soon as the wave function
1959: is translationally invariant the wave function does not vanish
1960: for the whole translation path, implying that the spin-up and -down
1961: nodal domains are interconnected and the wave function has only two nodal cells. 
1962: It is straightforward to extend the proof
1963: to arbitrary size closed-shell system. 
1964: The configuration of particles can be given as follows: an odd number of pairs (eg, one) in each
1965: spin channel
1966: is positioned in a similar way as in
1967: \ref{fig:2dhegunpol} so that the translation by $\pi$ exchanges the particles in the pair. For example,
1968: for a spin-up pair 
1969: $i,i+1$ we specify the coordinates as $y_i=-y_{i+1}=\pi/2$  and $x_i=x_{i+1}=\xi_0$ where $\xi_0$ is
1970: distinct from $x$ coordinates of the rest of particles with the same spin.
1971: The rest of the particles is positioned in such a way so that the $T_{\pi}^x$ brings them into
1972: a symmetric position regarding the reflection around axis $y$. 
1973: The wave function is translationally invariant and therefore uncorrelated wave function
1974: vanishes while for BCS case it is, in general, nonzero. 
1975: The two nodal cells property then 
1976: follows using the same general arguments as in the previous cases.
1977: 
1978: The correlated wave function has another important property  namely that one can wind around the
1979: periodic box a singlet pair of particles (ie, spin-up and spin-down pair) without hitting
1980: the node. Indeed, this is implied by the fact that there are only two nodal cells: 
1981: once this is the case one can then always
1982: find such a path that the pair of particles can wind around the box without
1983: node-crossing. For example, for the ten-particle 
1984: example above we can wind spin-up and -down particle pair around along $y=\zeta_0$ without hitting
1985: the node. We plot the wave function for two types of particle positions: first, the
1986:  spin-down particles are at identical 
1987: positions as the spin-down ones, ie, ${\bf r}_6={\bf r}_1$, ${\bf r}_7={\bf r}_2$, etc 
1988: (Fig.~\ref{fig:wind}a);
1989: second, the particles  in spin-up and -down channel were offset by small displacement
1990: ${\bf r}_{i+5}={\bf r}_i+0.2,$ $i=1,...,5$, see 
1991:  Fig.~\ref{fig:wind}b. 
1992: Consider now the simultaneous translation $T_{2\pi}^y$ of the particles 1 and 6. This 
1993: translation winds
1994: the pair around the box and the wave function values for the paths are  
1995: plotted in Fig.~\ref{fig:wind}a,b. 
1996: The value of the displacement between the particle positions neither the repositioning
1997: of the particles around the box is not crucial and there 
1998: exists a significant size subspace of particle positions for which the winding of the pair
1999: is possible.
2000: We have chosen just two simple examples as qualitative illustrations.
2001: 
2002: Note that the wave function enables to wind also a single particle without the node crossing. 
2003: However, single particle winding is possible both for correlated and uncorrelated wave functions
2004: since, say, the spin-up HF part has two nodal domains and correlated wave function 
2005: only smoothes out the HF nodes. Therefore this property is
2006: not affected by the correlation in any significant manner.
2007: 
2008: 
2009: \begin{figure}[ht]
2010: \centering
2011: \begin{tabular}{cc}
2012: \includegraphics[width=2.6in,clip]{windA.eps} \\
2013: (a)\\
2014:  \quad \\
2015:  \quad \\
2016:  \quad \\
2017: 
2018: \includegraphics[width=2.6in,clip]{windB.eps} \\
2019: (b) \\
2020:  \quad \\
2021: \end{tabular}
2022: \caption{ Winding the particles 1 (spin-up) and 6 (spin-down) around the $2D$ box along 
2023: the line $y=\zeta_0$. The dashed line is the Hartree-Fock wave function (arb.u.) while 
2024: the full line is
2025: the correlated BCS wave function.
2026: a) Particles 1 and 6, 2 and 7, etc, share the same position. The uncorrelated wave function
2027: touches the node quadratically since the spin-up and -down determinants are 
2028: identical. b)  The positions of particles in spin-down channel are offset by 0.2
2029: from the positions in the particle spin-up channel. 
2030: The uncorrelated wave function
2031: crosses the node multiple times since the spin-up and -down determinants are different.
2032: }
2033: \label{fig:wind}
2034: \end{figure}
2035: 
2036: The effect that singlet pair of particles can pass 
2037: through nodal openings between the spin-up and -down subspaces
2038: have been demonstrated for small number of particles before,
2039: for example, in our paper on employing pfaffian pairing functions for electronic structure
2040: problems \cite{bajdichpfaffians}. 
2041: The case here illustrates that this property stems from interconnected
2042: spin-up and -down subspaces and therefore applies to similar two nodal cell wave functions,
2043: in general.
2044: % as soon as the  wave function has 
2045: %only two nodal domains there always exists a path that a pair of particles can wind around the box.
2046: % a pair of spin-up and a pair of spin-down particles 
2047: %can simultaneously exchange
2048: %without crossing the node. Clearly, this effect is generic and does not say much about the phase 
2049: %%in which the electrons are (eg, superconducting state).  
2050: 
2051: It is well known that a Fermi liquid, such as the homogeneous electron gas, becomes unstable to
2052: a weak attractive interaction, develops Cooper pair instability and opens the superconductivity
2053: gap \cite{supercond}. This effect is captured by the BCS wave function which we used in the proof above. 
2054: Cooper pairs can therefore wind around the box without hitting the node although this effect
2055: is not exclusive to Cooper pairs only. 
2056: %Although
2057: % For example, the same effect can be observed in the Fermi
2058: %liquid with repulsive interactions which is typically not a superconductor.
2059:  In fact, the connectedness of  spin-up and -down
2060: subspaces is a rather generic property in the sense 
2061: that it appears also in systems which are not necessarily superconducting, ie, in Fermi liquids
2062: with repulsive interactions. Since superconductivity is characterized by macroscopic phase
2063: coherence and a number of other properties, which might or might not be related to
2064: the nodal topologies, on the basis of the analysis of the BCS wave function above
2065: one expects  that the  nodal opening 
2066: {\it necessarily} appears in the 
2067: superconducting state, 
2068: however,
2069: for the superconductivity to occur this condition is not sufficient.
2070: In addition, here we assume only
2071: the simplest $s-$wave pairing; for the $p-$wave or higher angular momentum pairing the situation
2072: is less clear and needs to be further investigated. 
2073: 
2074: \section{\label{sec:dm} VII. Nodes of fermionic density matrices.}
2075:  
2076: In this part we generalize the ideas presented in preceding 
2077: section to temperature dependent density matrices \cite{davidnode}. Consider
2078: first a system of spin-polarized fermions. The temperature/imaginary time density matrix  
2079: is given by
2080: \begin{equation}
2081: \varrho(R',R,\beta)= \sum_{n} e^{-\beta E_n} \Psi^{*}_n(R')\Psi_n(R)
2082: \label{eq:dm1}
2083: \end{equation}
2084: where $\beta$ is the inverse temperature and  
2085: the sum is over the complete system of eigenstates of a given Hamiltonian $H$. 
2086: It is clear that the density matrix
2087: is antisymmetric in particle exchanges in the same manner as  
2088: the wave function $\Psi(R)$ or $\Psi(R')$. Therefore the notion of fermion
2089: nodes can be generalized also to density matrix 
2090: in $(2dN+1)$ dimensions since there is an explicit dependence on $\beta$ as well. As pointed
2091: out elsewhere \cite{davidnode}, for fixed $R'$ and $\beta$ one can study the nodes and nodal cells 
2092: in the $dN$ dimensional $R-$subspace. Similarly to the wave function, the node then becomes
2093: $(dN-1)$-dimensional manifold with the generalization
2094: that it is dependent on $R'$ and $\beta$. 
2095: For the fixed  $R'$ and $\beta$
2096: the tiling property holds in the same
2097: manner as for the wave functions. The key additional feature of the density matrix is
2098: that once there are only two nodal cells at some initial $\beta_0$ than this property holds for 
2099: any $\beta >\beta_0$ \cite{davidnode}. This is not difficult to understand since the density matrix fulfills
2100: the following linear equation
2101: \begin{equation}
2102: -{\partial \varrho(R,R',\beta)
2103: \over \partial \beta}=H\varrho(R,R',\beta)
2104: \label{eq:dm2}
2105: \end{equation}
2106: with an initial condition
2107: \begin{equation}
2108:  \varrho(R,R',0) = {\cal A} \delta(R-R') ={\det }[\delta({\bf r}_i-{\bf r'}_j)]
2109: \label{eq:dm3}
2110: \end{equation}
2111: where ${\cal A}$ is the antisymmetrizing operator.
2112: 
2113: We now understand that for almost any  Hamiltonian with interactions 
2114: the density matrix will have only
2115: two nodal cells for sufficiently large $\beta$ (ie, at  low temperatures). This is due to the fact
2116: that at sufficiently low temperature  the ground state becomes dominant
2117: (Eq. \ref{eq:dm1}). The key point now is  to show that this
2118: is the case also for {\it high} temperatures.  
2119: The free particle density matrix 
2120: is given by
2121: \begin{equation}
2122:  \varrho (R,R',\beta) = (2\pi\beta)^{-dN/2}{\rm det} \left[\exp(-|{\bf r}_i-{\bf r'}_j|^2/2\beta)\right]
2123: \label{eq:dm4}
2124: \end{equation}
2125:   where we assume atomic units with $\hbar^2/m=1$.
2126:  (For other than free
2127:   boundary conditions, such as for the periodic ones,
2128:  the expression has to be modified accordingly.)
2129: Note that this
2130:  density matrix is universal since   
2131: at sufficiently high
2132: temperatures the interactions become irrelevant. 
2133: 
2134: There are several ways how to prove that 
2135: the high-temperature density matrix has only two nodal cells.
2136: For very small $\beta$ one can use the induction as follows. 
2137: Assume that $N$ particles are described by the density matrix given by
2138: Eq.~\ref{eq:dm4} and the particles are connected by triple exchanges for a fixed $R',\beta$.
2139: We add an additional particle 
2140:  with the label $N+1$ to the system which 
2141: occupies a certain region of the configuration (free) space.
2142: The particle $N+1$ is positioned at the border of the occupied region
2143: and let us assume that the particles with labels 
2144: $N-1$ and $N$ are its closest neighbours. Let us move the three particles $N-1, N, N+1$
2145:  away from the rest without crossing the node (what can be always done
2146: by appropriate positioning). 
2147: Since for small $\beta$ the overlaps of gaussians become small, one can factorize the determinant 
2148: into a product of the three particles $N-1,N,N+1$ determinant and the determinant for the
2149:  rest. 
2150: For the three particles the density matrix has only two nodal cells as one can show easily
2151: \cite{davidnode}
2152: and therefore
2153: the additional particle is connected. This applies to both $R$ and $R'$ subspaces since they must
2154: have identical properties. (In what follows we will show that, in fact, at high temperatures 
2155: the primed and unprimed spaces are connected as well.)  
2156: 
2157: There is also an alternative proof which is interesting also on its own since it provides
2158: a different view on the density matrices. 
2159: Note that the functional form of the free particle density matrix 
2160: has a unique property. Comparing the Eq.~\ref{eq:dm4} with the  
2161:  BCS wave function (Eq.~\ref{eq:bcs}) we see that the
2162: high-temperature density matrix can be identified with a BCS wave function
2163: if we properly define an underlying effective model. Instead of
2164: our original  system of $N$ spin-polarized fermions, let as consider
2165: a model system with $2N$ particles so that
2166:  the configurations $R$ and $R'$ denote positions
2167: of these {\it different } sets of particles, which we will call for simplicity unprimed and 
2168: primed particles. Let us then define a new Hamiltonian ${\tilde H}(R,R')$ 
2169: with
2170: an effective quadratic interaction between the  unprimed and primed particles as
2171: given by
2172: \begin{equation}
2173: {\tilde H}(R,R') =  T + T' + V_0\beta^{-1} \sum_{i,j} |{\bf r}_i - {\bf r'}_j|^2
2174: \label{eq:dm5}
2175: \end{equation} 
2176: where $T$ and $T'$ denote kinetic energy operators for the corresponding sets of particles and 
2177: $V_0$ is a constant with appropriate dimensions. 
2178: Note that particles within the given set, say, unprimed, are antisymmetric but 
2179: otherwise
2180:  do not interact with each other. The interaction appears only between the primed 
2181: and unprimed degrees of freedom as given by the Hamiltonian $\tilde H(R,R')$.
2182: For $\beta \to 0$ the exact wave function for this system is a BCS wave function 
2183: $\Psi(R,R')={\rm det}[\phi({\bf r}_i,{\bf r'}_j)]$ where the indices  
2184: $i$ and $j$ label unprimed and primed particles, respectively. 
2185: The pairing function $\phi({\bf r}_i,{\bf r'}_j)$ is obviously the gaussian given above. 
2186: It is not too difficult to demonstrate that this wave function
2187: (and the density matrix) has only two nodal cells. 
2188: For example, we can expand the gaussian into plane waves
2189: \begin{equation}
2190: \exp(-|{\bf r}_i-{\bf r'}_j|^2/2\beta)=\sum_{\bf k}c_{\bf k}e^{i{\bf k}\cdot({\bf r}_i-{\bf r'}_j)}
2191: \label{eq:dm6}
2192: \end{equation}
2193: where $\{c_{\bf k}\}$ are expansion coefficients.
2194:  The sum is over states within the Fermi sphere of
2195: a periodic box which accommodates $2N$ particles for a given density. Let us specify that
2196: $N$ is large and $\beta$ is such that 
2197: the system can be considered classical so that the actual interaction in the 
2198: original Hamiltonian $H$ is irrelevant.
2199: Then the density matrix corresponds to the Hartree-Fock product
2200: \begin{equation}
2201: {\rm det}[e^{i{\bf k}_i\cdot {\bf r}_j}]{\rm det}[e^{-i{\bf k}_i\cdot {\bf r'}_j}]
2202: \label{eq:dm7}
2203: \end{equation}
2204: However, we have already proved that such system has only two nodal cells. In addition, 
2205: if the sum includes also "excited states"
2206: (ie, beyond the Fermi sphere) due to the primed-unprimed interactions, 
2207:  we find that  the unprimed and primed 
2208: nodal cells are interconnected.  Therefore at classical temperature  $\beta_0$
2209: the density matrix has only two nodal cells and then the same is true for arbitrary
2210: $\beta>\beta_0$. This primed-unprimed
2211: interconnection becomes less pronounced and ceases to exist at $\beta \to \infty$ since
2212: then the density matrix is proportional to the "noninteracting" product
2213: $\Psi_0(R)\Psi_0(R')$ where $\Psi_0$ is the ground state for the original physical system 
2214: of $N$ fermions described by $H$. This proof is therefore based on an interesting duality between the 
2215: $N$ spin-polarized fermions at classical temperatures and the model
2216: system with $2N$ particles with a temperature-dependent,
2217:  harmonic interaction between the unprimed and primed
2218: subspace particles. 
2219: 
2220: Possibly, there might be also a third way how to prove the minimal number of nodal cells for
2221: the high-temperature density matrix  
2222: through diagonalization of the quadratic Hamiltonian $\tilde H$ (we have not 
2223: investigated this  possibility). 
2224: %The derivations above were carried out
2225: %only for  spin-polarized density matrix. However, using basically the same steps
2226: %it is not difficult to show that the same is true also for the spin-unpolarized case
2227: %assuming that the exchange given above was 
2228: The fact that also the density matrices have two nodal cells is important for the path 
2229: integral Monte Carlo methods which for fermions often employ the fixed-node 
2230: approximation adapted for the 
2231: path integrals \cite{davidhe}. 
2232: %The fact that the singlet particle pair can 
2233: %wind around the box without hitting the node indicates that it might be relevant
2234: %for the superconductivity. It is very suggestive that a node-free winding of 
2235: %a singlet pair is 
2236: %% a necessary condition for superconductivity to occur (albeit not sufficient since
2237: %the effect, as we have seen, is rather generic, for example it appears regardless
2238: %whether the interaction is attractive or repulsive). In a transport process the node
2239: %can be considered a scattering potential since the wave function is prescribed to have 
2240: %a particular shape there. In that sense the nodes have probably more deep relationship 
2241: %to transport phenomena and this aspect remains to be investigated.  
2242: 
2243: \section{VIII. Discussion and conclusions.}
2244:  
2245: Inspired by previous
2246: conjectures and numerical studies
2247: \cite{jbanderson,davidnode,lesternode,sorella,foulkes,dariobe,bajdichnode}
2248: the presented analysis and proofs generalize and clarify the properties
2249: of ground states fermion nodes. 
2250: We have employed symmetries, wave function factorizations
2251: and triple exchanges 
2252: to prove that for $d>1$ the closed-shell ground states 
2253: have two nodal cells for arbitrary number of particles
2254: in several paradigmatic models.
2255: In this paper the proofs were carried for closed shells, mainly to avoid 
2256: additional complications from degeneracies and some
2257: of these aspects will be addressed in subsequent papers.
2258:    
2259: It is perhaps more interesting
2260: to discuss viable possibilities when the ground state wave functions might have 
2261: {\it more 
2262: than two} nodal cells. Clearly, it is easy to generate more nodal cells by imposing 
2263: additional symmetry
2264: or boundary conditions. Another possibility comes from very strong interactions,
2265: for example, whenever the effect of interaction
2266: would
2267:  become competitive with the kinetic energy increase from forming additional 
2268: nodes it is possible that more than the minimal two nodal cells can form. 
2269: %Note that the tiling
2270: %property rules out forming of nodal "bubles" and therefore additional nodal cells will
2271: %be all equivalent
2272: %  interaction  electronic systems whenever exchange and correlation
2273: %energy are of the same order of magnitude neither HF nor BCS framework can be considered
2274: %reliable and the nodes in these systems can be different from what we have derived
2275: %above. 
2276: One can also construct nonlocal interactions
2277: which would violate some of the properties mentioned here (eg, the tiling property),
2278: or reorder the states so that, for example, excited states could 
2279: lie below the ground state, etc.
2280: Another candidate for unusual effects in the nodal structure are open-shell systems with
2281: large degeneracies and densities of states at the Fermi level.  Some of these 
2282: interesting issues will be subject of future studies.
2283: 
2284:  
2285: In conclusion,  we have presented a number of new results which reveal the structure
2286: of nodes and nodal cells
2287:  of fermionic wave functions. 
2288: Building upon ideas introduced in previous paper
2289: we were able to demonstrate the minimal number of two nodal cells in several 
2290: spin-polarized models such
2291: as noninteracting fermions in a periodic box, in a box with zero boundary conditions,  fermions
2292: on a spherical surface, etc. This enabled us to formulate a theorem
2293: which states that in $d>1$ the minimal two nodal cells are present for any
2294: Slater determinant with monomial matrix elements for any size which
2295: allows for a closed-shell nondegenerate ground state. We have shown that this 
2296: property extends also to cases which cannot be described by the Slater matrix of monomials
2297: such as the noninteracting and HF atomic states up to the $3d$ shell.
2298: We have studied the effect of interactions on the noninteracting nodal cells
2299: of spin-unpolarized, closed-shell singlets.  Our results show
2300: that, in general, the interactions smooth out
2301: the noninteracting multiple nodal cells into the minimal number of two. 
2302: For interacting homogeneous
2303: electron gas we have demonstrated that the two nodal cells allow 
2304: singlet pairs of particles to wind around 
2305: the periodic box without crossing the node.
2306: Finally, we have 
2307: shown that the temperature/imaginary time
2308:  density matrix has very similar nodal structure and therefore
2309: the minimal two nodal cells property applies also to density matrices with
2310: important implications for path integral Monte Carlo 
2311: simulations. We have demonstrated this by using an appropriate mapping of the density matrix
2312: onto the ground state of a model systems with twice as many particles interacting 
2313: with temeperature-dependent harmonic potentials.
2314:  
2315: 
2316: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2317: %	Bibliography, assumed extension is *.bib
2318: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2319: 
2320: I am grateful to J. Koloren\v c, M. Bajdich and L.K. Wagner for reading the manuscript, 
2321: comments and discussions.  
2322: I would like to acknowledge the support by
2323: %ONR-N00014-01-1-0408, and
2324: NSF DMR-0102668, DMR-0121361 and  EAR-0530110 grants.
2325: 
2326: \begin{thebibliography}{99}
2327: \bibitem{qmchistory} D.M. Ceperley and M.H. Kalos,
2328: in {\sl Monte Carlo Methods
2329: in Statistical Physics}, Ed. K.
2330: Binder, pp.145-194, Springer, Berlin, 1979;
2331: K.E. Schmidt and D. M. Ceperley,
2332:  in {\sl Monte Carlo Methods in Statistical
2333: Physics II}, pp.279-355, Ed.
2334:  K. Binder, Springer, Berlin, 1984.
2335: \bibitem{hammond} B.L. Hammond, W.A. Lester, Jr., and P.J. Reynolds,
2336: {\sl Monte Carlo Methods in ab initio quantum chemistry}, World Scientific,
2337: Singapore, 1994.
2338: \bibitem{qmcrev} M.W.C. Foulkes, L. Mitas, R.J. Needs and G. Rajagopal,
2339:   Rev. Mod. Phys.  {\bf 73}, pp. 33-83 (2001).
2340: \bibitem{jbanderson} J. B. Anderson, J. Chem. Phys. {\bf 65} 4121 (1976);
2341: J. B. Anderson, Phys. Rev. A, {\bf 35}, 3550 (1987).
2342: \bibitem{davidnode} D. M. Ceperley, J. Stat. Phys. {\bf 63}, 1237 (1991).
2343: \bibitem{lesternode} W. A. Glauser, W. R. Brown, W. A. Lester, D. Bressanini, B. L.
2344:   Hammond and M. L. Koszykowski, J. Chem. Phys. {\bf 97}, 9200 (1992).
2345: \bibitem{sorella} M. Casula and S. Sorella, J. Chem. Phys.
2346: {\bf 119}, 6500 (2003)
2347: \bibitem{foulkes} W. M. C. Foulkes, Randolph Q. Hood and 
2348: R. J. Needs, Phys. Rev. B {\bf 60}, 4558 (1999).
2349: \bibitem{dariobe} D. Bressanini, D. M. Ceperley and P. Reynolds, in
2350: {\sl Recent Advances in Quantum Monte Carlo Methods, II}
2351: Ed. W. A. Lester, S. M. Rothstein, and S. Tanaka, World Scientific, Singapore (2002).
2352: \bibitem{dariohe} D. Bressanini,and P. Reynolds, Phys. Rev. Lett. {\bf 95}, 110201 (2005).
2353: \bibitem{bajdichnode} M. Bajdich, L. Mitas, G. Drobny, and L. K.
2354: Wagner,  Phys. Rev. B {\bf 72}, 075131 (2005)
2355: \bibitem{cmt28} L. Mitas, G. Drobny, M. Bajdich, L.K. Wagner,
2356: in {\sl Condensed Matter
2357: Theories, Vol.20}, Eds. J.W. Clark, R.M. Panoff, and H. Li, Nova Science Publishers, New York, 2006;
2358: cond-mat/0409406. 
2359: \bibitem{berger} M. Berger, {\sl A Panoramic View of Riemannian Geometry}, Springer, Berlin, 2000, p.84. 
2360: \bibitem{varchenko} A. N. Varchenko, Math. USSR Izvestiya, N.3, {\bf 35}, 543 (1990)
2361: \bibitem{lagr} C. K. Chui and L. M. Lai, in
2362: {\sl  Nonlinear and Convex Analysis}, 
2363: Eds. B.L. Lin and S.Simons, pp. 23-35, Marcel Dekker, 1987.
2364: %\bibitem{3rdvander} 
2365: \bibitem{bajdichpfaffians} M. Bajdich, G. Drobny, 
2366: L.K. Wagner, and K. E. Schmidt, Phys. Rev. Lett. {\bf 96},
2367: 130201 (2006). 
2368: \bibitem{mitasshort} L. Mitas, submitted, cond-mat/0601485
2369: \bibitem{carlson} J. Carlson, S.-Y. Chang, V. R. Pandharipande, and K. E. Schmidt, Phys. Rev.
2370: Lett. {\bf 91}, 050401 (2003).
2371: \bibitem{supercond} R. Shankar, Rev. Mod. Phys. {\bf 66}, 000129 (1994);
2372: A. J. Leggett, Rev. Mod. Phys. {\bf 47}, 331 (1975);
2373:  H. Bruus and K. Flensberg, {\sl Many-Body Quantum Theory
2374: in Condensed Matter Physics}, Oxford University Press, Oxford, 2004, p. 325.
2375: \bibitem{davidhe} D.M. Ceperley,  Rev. Mod. Phys. {\bf 67}, 279 (1995).
2376: \end{thebibliography}
2377: %XXXXXXX discussion, including abstract, 
2378: %                 End of the document
2379: \end{document}
2380: