1: %\documentclass[preprint,aps]{revtex4}
2: %\documentclass[aps,prl,preprint,groupedaddress]{revtex4}
3: \documentclass[twocolumn,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
4: %\documentclass[prl,showpacs, twocolumn,groupedaddress,floatfix,letterpaper]{revtex4}
5:
6:
7: %\newcommand{\mwcd}[1]{\marginpar{\protect\scriptsize \protect\setlength
8: % {\baselineskip}{8pt} to editor- MS revision: #1}}
9: %\setlength {\marginparwidth}{7.0cm}
10:
11: \usepackage[dvips]{graphicx}% Include figure files
12:
13: \begin{document}
14:
15: \title
16: {Ground-state energy and Wigner crystallization in thick 2D-electron
17: systems}
18:
19: %
20: \author
21: {D. Jost
22: }
23: \email[Email:\ ]{daniel.jost@ens-lyon.fr}
24: \affiliation{
25: Ecole Normale Sup\'erieure de Lyon, 46 All\'ee d'Italie 69364 Lyon, France}
26: %
27: %
28: \author
29: {M. W. C. Dharma-wardana}
30: \email[Email for correspondence:\ ]{chandre.dharma-wardana@nrc.ca}
31: \affiliation{
32: Institute of Microstructural Sciences,
33: National Research Council of Canada, Ottawa, Canada. K1A 0R6
34: }
35: %\date{07-Nov-2002 mosfet paper}
36: %\date{20-Nov-2004 quasi paper}
37: %\date{30-July-2005 wigner}
38: \date{\today}
39: %
40: \begin{abstract}
41: The ground state energy of the 2-D Wigner crystal is
42: determined as a function of the thickness of the electron
43: layer and the crystal structure.
44: The method of evaluating the exchange-correlation energy is
45: tested using known results for the infinitely-thin 2D system.
46: Two methods, one based on the local-density approximation
47: (LDA), and another based on the constant-density approximation
48: (CDA) are established by comparing with quantum Monte-Carlo (QMC)
49: results. The LDA and CDA estimates for the Wigner transition
50: of the perfect 2D fluid are at $r_s=38$ and 32 respectively,
51: compared with $r_s=35\pm5$ from QMC. For thick-2D layers as
52: found in
53: Hetero-junction-insulated-gate field-effect transistors,
54: the LDA and CDA predictions of the Wigner transition are at
55: $r_s=20.5$ and 15.5 respectively.
56: Impurity effects are
57: not considered here.
58: \end{abstract}
59: \pacs{PACS Numbers: 05.30.Fk, 71.10.-w, 71.45.Gm, 71.15.Mb}
60: %\vspace{0.5in}
61: %
62: \maketitle
63: %
64: \section{Introduction}
65: %{\it Introduction.}---
66: Two-dimensional (2D) electron layers
67: exist, for example, at the interface between GaAs and
68: Ga$_{1-x}$Al$_x$ As, or at the interface of a metal oxide and a
69: semiconductor. Such interfaces are important in field-effect
70: transistors and other devices.
71: The state of this 2D electrons system can be fluid or,
72: at sufficiently low density, the electrons condense and
73: crystallize\cite{ando,krav}.
74: We define
75: the 2D-density parameter $r_s$ given by $A/N=\pi{r_s}^2$ where $A$
76: is the area
77: occupied by the $N$ electrons. The Wigner solid appears for $r_s\geq
78: 35 a_0$ \cite{attac} where $a_0=\hbar^2/(m^*e^{*2})$ is the Bohr radius.
79: Here $m^*,e^*$ are effective parameters for the mass and the charge
80: of the electron, and absorb the dielectric constant, the band
81: mass and other material properties of the system. Thus
82: in GaAs, the effective atomic unit of energy is reduced from 27.21 eV
83: to the milli-Volt range. These (reduced) atomic units,
84: such that $m^*=e^*=\hbar$=1 will be
85: assumed through out this paper.
86: There has been many studies of 2D-electron liquids or Wigner crystals
87: \cite{attac, rapisarda,tanatceper,ceper,bonsall,iyakutti}, especially
88: using quantum Monte Carlo (QMC)
89: simulations and other methods\cite{bonsall,prl2}
90: assuming that the 2-D layers are infinitely thin.
91: However, although the 2D electrons reside
92: in the $(x,y)$-plane, they have a transverse
93: density $\eta(z)$
94: confined to the lowest subband of the hetero-structure \cite{ando}.
95: While the quasi-2D electron liquid has recently seen much
96: attention, both experimental\cite{tan, stormer} and
97: theoretical\cite{ssc,asgari,zhang,morsen}, the Wigner crystal in
98: thick 2D layers has not been followed up since the work of
99: Fujiki and Geldart
100: \cite{fuji2}. Fujiki et al., have
101: determined the effect of the 2D-layer thickness on the
102: electrostatic energy and found that the hexagonal lattice
103: is the most-stable crystal structure, as with the
104: $\delta$-thin 2D layer ($\delta$2D).
105: However, they did not consider the effect of
106: exchange and correlation which is usually addressed
107: via Quantum Monte Carlo methods, or via a detailed analysis of
108: the correlated phonons in the
109: electron crystal\cite{bonsall}.
110: We note that recent Hartree-Fock (HF) calculations of $\delta$2D
111: Wigner crystals
112: using large plane-wave basis sets,
113: e.g, those of Trail et al\cite{trail}.,
114: seem to recover a HF energy nearly identical to the single-
115: Gaussian {\it harmonic} approximation for localized electrons.
116: Various aspects of such a model have been considered in
117: a brief but insightful paper by Nagy\cite{nagy,polini}.
118: In this study we show that the single-Gaussian approximation,
119: and the local-density approximation (LDA),
120: can recover the QMC total
121: energy with surprising accuracy. We also show that a
122: method based on constructing a constant-density approximation
123: (CDA) to the inhomogeneous density\cite{ggsavin,ssc}
124: can be profitably used for calculating the
125: electrostatic potentials and the exchange-correlation
126: energies of these systems.
127:
128: The plan of the paper is as follows. In section~\ref{coul-sec} we
129: introduce the Hamiltonian, the effective Coulomb interaction
130: in quasi-2D layers, and calculate the electrostatic energy
131: of the lattice for several 2D-crystal structures. Here we use the CDA
132: to replace Fang-Howard type densities in the $z$-direction\cite{ando,ssc},
133: thus simplifying the analytical work.
134: The details of lattice-sum evaluations
135: are relegated to an appendix.
136: In section~\ref{perfect} we consider the
137: $\delta$-thin 2D layer and present results for the
138: gaussian-localized model.
139: we also present the
140: exchange-correlation energy $E_{xc}$ calculation using the CDA and the
141: LDA.
142: The resulting total energy is very
143: close to that of QMC and recovers a liquid$\to$solid
144: Wigner transition (WT) at
145: $r_s\sim$32 to 38, while the current QMC estimate is $r_s=35\pm5$.
146: In section~\ref{gau-thick} we consider Gaussian-localized
147: 2D systems with finite
148: thickness, for 2D layers found in HIGFETS.
149: That is, in systems where the layer thickness is
150: also defined by the sheet density, as in Tan et al\cite{tan,stormer}.
151: Here we have no QMC results for comparison. The total
152: energy of the quasi-2D Wigner crystal is compared with the
153: total energy of the quasi-2D liquid\cite{ssc}.
154: Here the WT is found to occur at $r_s\sim$15 to 21
155: in quasi-2D layers realized in clean HIGFETS,
156: %
157: \section{The Coulomb energies of 2D lattices}
158: \label{coul-sec}
159: The Hamiltonian of our system is, in atomic units,
160: \begin{equation}
161: H=H_{ke}+H_{ee}+H_{eb}+H_{bb}
162: \end{equation}
163: where the first term is the kinetic energy of the electrons.
164: The three remaining terms are the electron-electron interaction and
165: the interactions involving the uniform, static neutralizing background,
166: indicated by the subscript $b$.
167: This neutralizing background arises from a homogeneous
168: layer of donor ions which have acquired a positive charge after
169: donating their valence electrons in forming the 2D-electron layer.
170: We assume that the electron layer is confined
171: near the plane $z=0$ and extends into
172: the region of $z>0$ due to the finite width of the envelope function.
173: The donor ions are modeled by a homogeneous layer of
174: positive charge of areal density $\rho_d=N/A$, situated at $z=-b_d$,
175: where $b_d=|b_d|$ is a positive quantity.
176: The $z$-direction density is $\eta(z)$, and
177: in the plane, an areal density $\rho_e(\mathbf{r})$ with $\mathbf{r}=(x,y)$.
178: The subband distribution $\eta(z)$ is usually modeled by a
179: Fang-Howard distribution $\eta_{fh}(z)=(1/2 b^3)z^2 e^{-\frac{z}{b}}$
180: (n.b., our $b=1/b$ used in Ref.~\cite{ando}),
181: or various other forms, e.g, that of a quantum well.
182: The form of the density is obtained by fitting to a self-consistent
183: calculation of the Schrodinger equation for the electron motion in the
184: $z$-direction. In our work, we do not repeat this calculation, but simply
185: take the value of the parameter $b$, or other parameters
186: needed to define the self-consistent solution for the subband.
187: Moreover, as discussed below, such inhomogeneous
188: densities can be replaced by a constant-density slab
189: having an {\it equivalent electrostatic potential},
190: using the CDA discussed by Dharma-wardana\cite{ssc}.
191: The CDA method \cite{ssc} involves
192: replacing an inhomogeneous density $\eta(z)$ by
193: a slab of constant-density
194: $\bar{\eta}$ of width $a$ linked to $\eta(z)$ by
195: \begin{equation}
196: \label{cda-eqn}
197: \bar{\eta}=\frac{1}{a}=\int {\eta(z)}^2 \,dz
198: \end{equation}
199: This equation has also been proposed by Gori-Giorgi
200: et al\cite{ggsavin}., in a method for
201: calculating system-adapted correlation
202: energies.
203: Using Eq.~\ref{cda-eqn} a Fang-Howard (FH)
204: density of length scale $b$ can
205: be replace by a homogeneous
206: density of width $a=(16b)/3$.
207: %
208: \begin{figure}[t]
209: \includegraphics*[width=8cm,height=6cm]{density.eps}
210: \caption{(Color online) Profiles of the Fang-Howard density (solid curve)
211: for $b=4$ and its
212: equivalent constant density approximation(CDA, dashed curve).
213: Inset: the bare Coulomb potential $1/r$ and the Coulomb
214: potential $F(r)/r$ modifed by the Fang-Howard profile. The
215: triangles are calculated using the CDA. The CDA width
216: a=21.33 for $b$=4 corresponds to a HIGFET at $r_s=32.496$.
217: }\label{fang}
218: \end{figure}
219: %
220:
221: Consider two electrons in a quasi-2D
222: layer separated by a distance $r$ in the 2D plane,
223: and located at $z_1$ and $z_2$,
224: with a FH distribution $\eta(z)$ in the $z$ direction.
225: Then the Coulomb interaction is of the form
226: \begin{equation}
227: W(r)=\int d z_1\, d z_2
228: \frac{\eta(z_1)\eta(z_2)}{|r^2+(z_1-z_2)^2|^{1/2}}
229: \end{equation}
230:
231: $W(r)$ may be written as $F(r)/r$ where
232: $F(r)$ is the form factor. No analytic form exists if
233: $\eta(z)$ is the FH form, while the $q$-space form,
234: $F(q)2\pi/q$
235: is analytically available.
236: For GaAs/AlAs based HIGFET-like systems, it takes the form,
237: \begin{equation}F(q)=[1+\frac{9q}{8b'}+
238: \frac{3q^2}{8b'^2}][1+\frac{q}{b'}]^{-2}
239: \end{equation}
240: where $b'=1/b$ and follows the definition in
241: Ref.~\cite{ando}.
242: However, if the FH distribution is
243: replaced by the CDA, then $F(r)$ and $F(q)$ are given by
244: \begin{eqnarray}
245: \label{cdmeqn}
246: %W(r)=\frac{2}{w}\left[\ln\frac{r}{-w+\surd{(r^2+w^2)}}
247: %\,\,-(r+\surd{(r^2+w^2)})\right]
248: W(r)&=&V(r)F(s),\;\; s=r/a, \;\; t=\surd{(1+s^2)} \\
249: F(s)&=&2s\left[\log\frac{1-t}{s}
250: \,\,+1-t\right]
251: \end{eqnarray}
252: and
253: \begin{eqnarray}
254: W(q)&=&V(q)F(p),\,\,\, p=qa, \,V(k)=2\pi/q\\
255: F(p)&=&(2/p)\{(e^{-p}-1)/p+1\}
256: \end{eqnarray}
257: The form factors $F(r)$, $F(q)$ are a
258: measure of the reduction of the strength of
259: the 2D interaction due to the thickness effect.
260: These results provide equivalent analytical
261: formulae for the FH-density,
262: and tend to the ideal 2D Coulomb
263: potential when the width $a$ tends to zero.
264: Also, for HIGFETS, it is known that the
265: FH-parameter $b$ is linked to the 2D density parameter
266: $r_s$. Hence it can be shown\cite{ssc} that
267: \begin{eqnarray}
268: \label{fh2w}
269: b&=&(2r_s^2/33)^{1/3}\\ %using the b=b_{DJ}=1/b_{FH}
270: a&=&16b/3\\ %16b_{dj}/3 instead of 16/(3b_{fh}
271: \beta&=&b/r_s=[2/(33r_s)]^{1/3}
272: \end{eqnarray}
273: Hence $\beta$, the FH parameter $b$
274: in units of $r_s$, and also the ratio $a/r_s$,
275: i.e., ($z$-width)/(2D-disk radius)
276: decrease as $r_s^{1/3}$ with increasing $r_s$.
277: %
278: %
279: %
280: %
281: \subsection{Coulomb energy}
282: %
283: In the following we do not at first specify the form
284: of the transverse density $\eta(z)$.
285: In calculating the Coulomb energy $E_{Cou}$, i.e, the
286: electrostatic energy, we
287: isolate the long-range contributions which cancel in the $q=0$
288: limit, since we are dealing with a homogeneous, neutralizing, static
289: background. The total Coulomb energy
290: is the sum,
291: \begin{equation}\label{ecoul}
292: E_{Cou}=\lim_{q \rightarrow 0} [E_{dd}(q)+E_{ee}(q)+2E_{ed}(q)]
293: \end{equation}
294: where
295: \begin{equation}
296: E_{dd}(q) =\frac{1}{2}\int d^2r \int
297: d^2r'{\rho_d}^2\,\frac{e^{i\mathbf{q}\cdot(\mathbf{r-r'})}}{|\mathbf{r-r'}|}
298: \end{equation}
299:
300: \begin{widetext}
301: {\setlength\arraycolsep{2pt}
302: \begin{eqnarray}
303: E_{ee}(q) & = & \frac{1}{2} \int d^2r \int
304: d^2r'\rho_e(\mathbf{r})\rho_e(\mathbf{r'})e^{i\mathbf{q}
305: \cdot(\mathbf{r-r'})}\int_{0}^{\infty} dz \int_{0}^{\infty}
306: dz' \frac{\eta(z)\eta(z')}{[(\mathbf{r-r'})^2+(z-z')^2]^{1/2}}\\
307: E_{ed}(q) & = & -\frac{1}{2}\int d^2r \int d^2r'\rho_d
308: \rho_e(\mathbf{r'})e^{i\mathbf{q}\cdot(\mathbf{r-r'})}
309: \int_{0}^{\infty} dz'\frac{\eta(z')}{[(\mathbf{r-r'})^2+(b_d+z')^2]^{1/2}}
310: \end{eqnarray}}
311: \end{widetext}
312: $E_{dd}$ is the interaction energy of the ions, $E_{ee}$ is
313: the interaction of the electron layer and $E_{ed}$ is the
314: energy due to interaction between the ions and the electrons.
315: To calculate these terms, we proceed as in Fujiki
316: and Geldart \cite{fuji2,fuji1}.
317:
318: \begin{equation}\label{edd}
319: E_{dd}(q)=\frac{\pi A{\rho_d}^2}{q}=\frac{N}{q{r_s}^2}
320: \end{equation}
321: We introduce the integral transformation
322: \begin{equation}\label{trans}
323: \frac{1}{|\mathbf{v}|}=\int_{0}^{\infty} \frac{dy}{\sqrt{\pi}}
324: y^{-1/2}e^{-y|\mathbf{v}|^2}
325: \end{equation}
326: and note that
327: %\begin{widetext}
328: {\setlength\arraycolsep{2pt}
329: \begin{eqnarray}\label{eed}
330: E_{ed}(q) & = & -\frac{1}{2}\int d^2r\int d^2r'\rho_d
331: \rho_e(\mathbf{r'})e^{i\mathbf{q}\cdot(\mathbf{r-r'})}
332: \int_{0}^{\infty} dz' \eta(z')\nonumber\\
333: & & \int_{0}^{\infty}\frac{dy}{\sqrt{\pi}}
334: y^{-1/2}e^{-y|\mathbf{r-r'}|^2-y|b_d+z'|^2} \nonumber\\
335: & = & -\frac{N}{2\sqrt{\pi}{r_s}^2}\int_{0}^{\infty}dy\, y^{-3/2}
336: e^{-\frac{q^2}{4y}}\nonumber\\
337: & & \int_{0}^{\infty} dz'\eta(z')e^{-y|z'+b_d|^2}
338: \end{eqnarray}}
339: %\end{widetext}
340: For $E_{ee}$, we use a lattice sum technique based on the
341: $\theta$ Jacobi function, Eq.(\ref{theta1}),
342: and its imaginary transform, Eq.(\ref{theta2}), given below:
343: \begin{equation}\label{theta1}
344: \theta(z,X)\equiv \sum_{l=-\infty}^{\infty} e^{2\pi lz}e^{-\pi l^2 X}
345: \end{equation}
346: \begin{equation}\label{theta2}
347: \theta(z,X)=\frac{e^{\pi z^2/X}}{\sqrt{X}}\theta(\frac{z}{iX},\frac{1}{X})
348: \end{equation}
349: We decompose the lattice into rectangular sublattices
350: indicated with sublattice vectors $\mathbf{\rho_j}$.
351: So, the position vectors of the electrons on
352: nodes $I$ and $J$ are given by
353: \begin{center}
354: $\mathbf{r_I}=m a_1 \hat{x} + n a_2 \hat{y},
355: \quad \mathbf{r_J}=(m'a_1+\rho_{j}^{x})
356: \hat{x}+(n'a_2+\rho_{j}^{y})\hat{y}$
357: \end{center}
358: where $m,m',n,n'$ are integers,
359: $a_1,a_2$ are lattice constants of sublattices.
360: For example, in a square lattice $a_1=a_2$ and
361: %
362: $\{ \mathbf{\rho_j} \}=\{(0;0)\}$,
363: in a hexagonal lattice $a_2=\sqrt{3}a_1$ and $ \{ \mathbf{\rho_j}
364: \}=\left\{(0;0),\left ( \frac{a_1}{2};\frac{a_1\sqrt{3}}{2} \right)\right\}$.
365: %
366: To proceed further, we need to specify the form of the density.
367: If the electrons are assumed to be
368: exactly localized on the nodes of the crystal, then
369: %
370: \begin{equation}\label{rhoefg}\rho_{e\delta}(\mathbf{r})=\sum_{I}
371: \delta(\mathbf{r-r_I}).\end{equation}
372: %
373: Such exact localization of the electrons provides the model for
374: the classical electrostatic energy, i.e, {\it the
375: Madelung energy}.
376: In the quantum calculation
377: we suppose that each electron is localized around a node $I$ of the
378: lattice and the wavefunction is
379: taken to be a gaussian normalized over the 2D plane,
380: \begin{equation}\label{phi} \phi_I(\mathbf{r})=\sqrt{\frac{2\alpha}{\pi}}e^{-\alpha
381: (\mathbf{r-r_I})^2}\end{equation}
382: The parameter $\alpha$ is chosen to
383: minimize the total energy. Hence the localized density is
384: \begin{equation}\label{rhoegauss}
385: \rho_{eG}(\mathbf{r})=\frac{2\alpha}{\pi}\sum_{I} e^{-2\alpha
386: (\mathbf{r-r_I})^2}
387: \end{equation}
388: The gaussian-width parameter $\alpha$ is of the form
389: $a/r_s^{3/2}$, with $a$ taking a lower-bound value of 0.5 (see
390: Ref~\cite{nagy}).
391: These two forms of the density will be studied below, and the Gaussian
392: approximation will be justified by comparison with
393: results from detailed plane-wave calculations.
394: %
395: \subsection{Calculation with the $\delta$-distribution}
396: \label{gau-thin}
397: Using Eq.(\ref{trans}) and (\ref{rhoefg}) we have
398: %\begin{displaymath}
399: \begin{eqnarray}
400: \label{delta-ee}
401: E_{ee}(q)&=&\int_{0}^{\infty} \frac{dy}{2\sqrt{\pi y}}f(y)
402: \sum_{I\ne J}e^{i\mathbf{q\cdot(r_I-r_J)}}e^{-y|\mathbf{r_I-r_J}|^2} \\
403: %\end{displaymath}
404: %with
405: %\begin{displaymath}
406: f(y)&=&\int_{0}^{\infty} dz \int_{0}^{\infty} dz'\eta(z)\eta(z')e^{-y(z-z')^2}
407: \nonumber
408: %\end{displaymath}
409: \end{eqnarray}
410: The details of the evaluation are given in the Appendix.
411:
412: %\begin{table*}
413: \begin{table}
414: \caption{The
415: Madelung energy, $E_{Cou}$ per electron are given for
416: different values of the Fang-Howard
417: parameter $\beta=b/r_s$ for hexagonal(hex), square(sq),
418: rectangular(rec), centered rectangular(cr) lattices defined by
419: their unit vectors $a_1$:$a_2$. The $r_s$ parameter
420: in the corresponding HIGFET, Eq.~\ref{fh2w},
421: is also given.
422: The energies are in
423: units of $1/r_s$. Thus the Madelung energy in Hartrees
424: for a $\delta$-thin
425: hexagonal lattice is $-1.106103/r_s$}
426: \label{tabecoul}
427: \begin{ruledtabular}
428: \begin{tabular}{lccccc}
429: HIGFET $r_s$ &$\infty$ & 60606 & 60.606 & 0.06060\\
430: $\;\;\;(a1:a2)$ $\beta\to$\rule{0pt}{2.6ex} & $0$ & $10^{-2}$ & $10^{-1}$ & $1$ \\
431:
432: \hline
433: hex($\sqrt{3}:1$) \rule{0pt}{2.6ex} & $-1.106103$ & $-1.052959$ & $-0.591433$
434: & $3.144793$ \\
435:
436: cr($\sqrt{2}:1$) \rule{0pt}{2.6ex} & $-1.104080$ & $-1.050937$ & $-0.589507$ &
437: $3.145401$ \\
438:
439: sq(1:1) \rule{0pt}{2.6ex} & $ -1.100244$ & $-1.047103$ & $-0.585854$ & $3.146555$ \\
440:
441: rec($\sqrt{2}$:1) \rule{0pt}{2.6ex} & $-1.078201$ & $-1.025072$ & $-0.564890$ &
442: $3.153217$\\
443:
444: rec($\sqrt{3}:1$) \rule{0pt}{2.6ex} & $-1.042843$ & $-0.989733$ & $-0.531301$ &
445: $3.163948$ \\
446:
447: \end{tabular}
448: \end{ruledtabular}
449: \end{table}
450:
451:
452: We have evaluated $E_{Cou}$, Eq.~\ref{ecoul} for
453: different lattices:
454: square, rectangular, hexagonal and
455: centered-rectangular. The Coulomb energy depends
456: only on $\beta=(b/r_s)=(3a)/(16 r_s)$, $r=(a_2/a_1)$
457: and $\{\mathbf{\rho_j}\}$.
458: Our numerical calculations of $E_{Cou}$ are summarized in
459: Table \ref{tabecoul}. Results for $\beta=10^{-2}$ are at
460: unrealistically low HIGFET densities, but are of formal
461: interest. Results for even smaller values
462: of $\beta$ may be found in Fujiki et al\cite{fuji1,fuji2}.
463: A comparison with the results of Ref.~\cite{fuji2}
464: shows that our results are in
465: agreement when a geometrical term arising from the slight
466: difference in the models is taken into account.
467: (As seen from the details given in the appendix,
468: we have an additive term
469: $N(2a/3)/{r_s}^2=N(32b/9)/{r_s}^2$ in our calculation while Fujiki
470: and Geldart have $N(33b/8)/{r_s}^2$. Agreement is obtained if
471: we replace our term by theirs).
472:
473:
474: It is seen that the total Coulomb energy increases as
475: $\beta=b/r_s$ increases for all cases studied. The hexagonal lattice
476: has the lowest energy for all $\beta$. Moreover, there is
477: no crossing between the different energy curves for any of the
478: lattice structures.
479:
480:
481: The dependence of the total Coulomb energy of the
482: centered-rectangular lattice
483: and rectangular lattice as a function of the
484: ratio $r=a_1/a_2$ for the quasi-2D system
485: remains similar to the $\delta$-thin case. Two equivalent minima
486: at $r=\sqrt{3}$ and $1/\sqrt{3}$ correspond to
487: the hexagonal structure. For the
488: rectangular lattice, the minimum corresponds to $r=1$, i.e., to the
489: square structure. We choose the range $\beta$=0.05 to 0.5, which
490: corresponds to $r_s\sim$ 0.5 to $\sim$500
491: and fit the Madelung energy of the stable hexagonal
492: lattice (see Table \ref{tabecoul}) to the analytic form
493: \begin{equation}
494: \label{thick-madelung}
495: E_{Cou}(r_s,\beta)=\sum_{i=0}^{i=4}c_i\beta^i/r_s
496: \end{equation}
497: where $c_0=-1.106103$, $c_1$=5.34722, $c_2=-2.15257$, $c_3$=1.48663, and
498: $c_4=-0.430473$.
499:
500: In Figure \ref{bfh}, we have plotted the Coulomb
501: energy as a function of $r_s$
502: using Eq.(\ref{fh2w}) to relate the thickness to the $r_s$ value.
503: We observe that the thickness of the system
504: has a significant effect on the energy.
505: %.
506: \begin{figure}
507: \includegraphics*[width=8cm,height=6cm]{bfh.eps}
508: \caption{(Color online) (Coulomb energy per electron in atomic units for a perfectly 2D
509: system and for 2D layers in a HIGFET, using
510: Eq.(\ref{fh2w}) to define the thickness.}\label{bfh}
511: \end{figure}
512: %
513: \subsection{Classical calculation with the gaussian distribution.}
514: If the electron distribution at each site were
515: a gaussian, the classical {\it electrostatic}
516: energy, $E_{ee}$, can be calculated using the same
517: techniques as before (Appendix).
518: %\begin{widetext}
519: {\setlength\arraycolsep{2pt}
520: \begin{eqnarray}
521: E_{ee}(q,\alpha) & = & \frac{1}{2\sqrt{\pi}}\int_{0}^{\infty}
522: dy\,y^{-1/2}\left(\frac{\alpha}{y+\alpha}\right)e^{-\frac{q^2}{4(y+\alpha)}}\nonumber\\
523: & & \sum_{I\ne
524: J}
525: e^{-\frac{y\alpha}{y+\alpha}(\mathbf{r_I-r_J})^2}
526: e^{-i\frac{\alpha}{y+\alpha}\mathbf{q\cdot(r_I-r_J)}}
527: \end{eqnarray}}
528: %\end{widetext}
529: We use the same integral separation with $E_{ee}^{<}$ and $E_{ee}^{>}$,
530: the Jacobi function $\theta$ and its transformation.
531: We may verify that when $\alpha$ tends to zero, that
532: is to say the gaussian distribution tends to the $\delta$-distribution,
533: $E_{ee}$ reduces to the Madelung energy of the
534: previous section. Also, if there is no effective
535: overlap among the gaussian distributions,
536: the distributions can be replaced by equivalent
537: point charges at the lattice sites and the Coulomb
538: energy should reduce to the Madelung
539: energy. However, as already remarked by
540: previous authors\cite{nagy,polini}, the charge
541: is not perfectly contained within the Wigner-Seitz
542: disk in the 2D problem.
543: The variations of the thickness and of the lattice type give results
544: similar to the $\delta$-thin case.
545: We consider the variation of $\alpha$ to minimize the
546: total energy within a quantum calculation, and hence do not develop this
547: classical calculation any further.
548: %
549: %
550: \section{Perfectly two-dimensional systems}\label{perfect}
551: In this section we consider a perfect,
552: i.e., $\delta$-thin 2D layer within a
553: Kohn-Sham density-functional approach\cite{kohn}
554: to the quantum mechanics of the problem. Since the
555: $\delta$-thin 2D system has been studied extensively, we use it
556: as a reference system to examine the LDA and the CDA as
557: useful tools for calculations of $E_{xc}$ of Wigner
558: crystals. The Hohenberg-Kohn theorem
559: asserts that the total energy is a functional
560: of the one-electron density,
561: and that it is a minimum for the true density distribution. We model the
562: one-electron density as a sum of gaussians centered on each lattice
563: site, and hence the variational problem reduces to a determination
564: of the width parameter $\alpha$ of the Gaussian that minimize the
565: total energy.
566: The total energy of the system at a given $r_s$ can be written as:
567: \begin{equation}
568: \label{dft-etot}
569: E_T= E_{HF}(\alpha,r_s) +E_{xc}(\alpha,r_s)
570: \end{equation}
571: where $E_{HF}(\alpha,r_s)$ is the Hartree-Fock energy of an electron.
572: It will be seen that this is effectively
573: the energy of an electron on a single site,
574: and moving in
575: the potential well created by the gaussian
576: distributions on other sites. If the gaussians were
577: perfectly localized, the Coulomb energy would not
578: depend on $\alpha$. The effect of the overlap
579: can be easily included in the variational problem,
580: with the energy given by $<\psi|H|\psi>/<\psi|\psi>$, and
581: this has an effect for small $r_s$. Here $\psi$ is
582: a Slater determinant of gaussians. For the hexagonal lattice,
583: the
584: overlap contribution from two nearest-neighbour gaussians is
585: $$s_{ij}(r_s)=\exp[-(\alpha/2)(1.09r_s)^2]$$
586: where $1.09r_s$ is the nearest-neighbour distance.
587: Unless the contrary is stated, the results reported here
588: will include the overlap correction.
589: The $\alpha$ which minimizes
590: the Hartree-Fock problem is not the same as that which minimized
591: the total energy inclusive of $E_{xc}$. In the next section we
592: look at the problem without $E_{xc}$
593:
594: \subsection{The Hartree-Fock energy $E_{HF}$}
595: \label{hf-sec}
596: The Hartree-Fock energy is composed of the classical Madelung energy
597: which defines a constant energy term, plus the quantum mechanical
598: energy associated with the motion of the electron in the field of the
599: other electrons. Since the electrons are strongly
600: localized, especially for large $r_s$,
601: a Slater determinant made up of one gaussian function
602: at each lattice site is commonly assumed.
603: The total energy consists of a
604: kinetic energy term and a potential energy term. These two
605: terms are equal by the virial theorem and hence we only
606: need to evaluate the kinetic energy.
607: Usually, Hartree-Fock energies contain a sizable exchange
608: contribution. However, the localized-gaussian exchange energy is
609: easily shown to be negligible,
610: and we called it the Wigner-exchange
611: energy, $E_{Xwc}$.
612:
613: In Table~\ref{hf-tab} we compare our localized-gaussian
614: (harmonic) calculation with the results of the extensive
615: plane-wave HF calculation by Trail et al\cite{thank-trail}.
616: %
617: \begin{table}
618: \caption{Comparison of the plane-wave calculation\cite{trail}
619: of the HF energies $E_{HF}$ of the $\delta$-thin 2D hexagonal Wigner
620: lattice with the single-Gaussian harmonic lattice energies.
621: $E^*_{har}$ and $E_{har}$ are energies without and
622: with the overlap corrections.
623: }
624: \begin{ruledtabular}
625: \begin{tabular}{ccccccc}
626: $r_s\to$ & 20& 30&40 & 60 \\
627: \hline
628: -$E_{HF}\times10 $ &0.447270 & 0.311642 &0.239528 &0.164036\\
629: -$E^*_{har}\times10$ & 0.447155 & 0.311786 &0.239822 & 0.164530\\
630: -$E_{har}\times10$ & 0.437058 & 0.308344 &0.238326 & 0.164113\\
631: \end{tabular}
632: \end{ruledtabular}
633: \label{hf-tab}
634: \end{table}
635: %
636: The results shown in Table~\ref{hf-tab}
637: show that the localized single-gaussian model is adequate to
638: describe the Hartree-Fock approximation
639: for this system\cite{trail2}.
640:
641: Note that our calculation is effectively an ``Einstein model'' of
642: oscillators, and the kinetic energy is
643: given by,
644: \begin{equation}
645: E_K(\alpha)= -\frac{N}{2}<\phi_I|\nabla_I^2|\phi_I>=N\alpha
646: \end{equation}
647: The
648: gaussian width which minimizes the energy
649: may be fitted by the form $\alpha=0.6263/r_s^{1.57}$.
650:
651:
652: Since the exchange of electrons actually involves a delocalization
653: process, we believe that the exchange integral evaluated with fixed
654: gaussians does not lead to a true evaluation of the exchange in
655: these systems.
656: The Wigner-exchange energy between
657: two electrons of spin $s_i,\,s_j$, is by definition
658: {\setlength\arraycolsep{2pt}
659: \begin{eqnarray}\label{excdef}
660: E^{ij}_{Xwc}& = & -\int d^2 r_i\,d^2 r_j
661: \phi_I(\mathbf{r_i})\phi_J(\mathbf{r_j})\frac{1}{r_{ij}}
662: \phi_I(\mathbf{r_j})\phi_J(\mathbf{r_i})\nonumber\\
663: & = & -\sqrt{\alpha\pi}e^{-\alpha(\mathbf{r_I-r_J})^2} \delta_{s_i,s_j}
664: \end{eqnarray}}
665: We can define a polarization parameter
666: $\zeta=(N_{\uparrow}-N_{\downarrow})/N$, therefore
667: \begin{equation}
668: \label{wigner-X}
669: E_{Xwc}(\alpha,\zeta)=-\frac{1}{2}\sum_{i\ne j} E^{ij}_{Xwc}
670: \end{equation}
671: This
672: $E_{Xwc}$, may be safely neglected for the values of $\alpha$
673: occurring in this problem.
674: %
675: %
676: %
677: \subsection{The CDA exchange and correlation energies}
678: %
679: \label{correlperf}
680: The correlation energy is the most difficult
681: object to calculate, and QMC has been the
682: preferred approach, even though this requires a
683: a major numerical effort.
684: However, for an uniform density profile,
685: the
686: correlation energy is well known \cite{attac}. Hence, as in
687: Eq.i~\ref{cda-eqn}, we map the inhomogeneous density in the
688: $(x,y)$ plane, $\rho(\mathbf{r})$, to a
689: homogeneous form via the $<\rho(\mathbf{r})^2>$ average of the CDA method.
690: Given a gaussian distribution,
691: {\setlength\arraycolsep{2pt}
692: \begin{eqnarray}
693: \bar{\rho} & = & \frac{1}{\pi {r_s}^2}\int d^2 r\, |\phi(\mathbf{r})|^2
694: |\phi(\mathbf{r})|^2 \nonumber \\
695: & = & \frac{\alpha}{\pi^2 {r_s}^2}
696: \end{eqnarray}}
697: We define the effective $r_s$ parameter $\bar{r_s}$
698: corresponding to the CDA density by $\bar{\rho}=1/(\pi{\bar{r}_s}^2)$,
699: \begin{equation}
700: \label{rsCDA}
701: \bar{r}_s=r_s\sqrt{\frac{\pi}{\alpha}}
702: \end{equation}
703: The correlation energy in the CDA, $E_{c}^{cda}$ for the
704: inhomogeneous distribution, inclusive of spin-polarization effects,
705: is now evaluated using $\bar{r}_s$ in any of the
706: well-known 2D-functionals\cite{attac}.
707: Note that for typical values of $\alpha$ at $r_s=20$, the
708: CDA density parameter is $\sim 400$, while at $r_s$=100,
709: it becomes $\sim$7000. Thus
710: we see that the CDA replaces the inhomogeneous fluid with
711: sharp Gaussian peaks by a uniform, {\it ultra-low-density}
712: 2D fluid.
713: In calculating $E_c(\bar{r}_s)$ using, say, the formula
714: due to Attaccalite et al., a difficulty arises since it is
715: fitted to a maximum $r_s$ of 40, together with
716: asymptotic forms, while
717: the CDA calls for $r_s$ values which are one or two orders
718: bigger. Nevertheless, we find
719: surprisingly good results (see below).
720:
721: At this point we ask if the exchange energy, evaluated
722: for this ultra-low density fluid, should also be
723: included. We believe that this is indeed the case. The fixed-gaussian
724: Wigner-exchange, Eq.~\ref{wigner-X} simply does not allow
725: any exchange, and ignores the possibilities of tunneling,
726: ring-exchange etc., that exist in the system. We consider that the
727: estimate of exchange obtained from the ultra-low density
728: fluid of the CDA accounts for such exchange effects. This
729: point of view is justified {\it post-facto} by the good
730: agreement of our total energies with the QMC total energies.
731:
732: \subsection{The LDA exchange and correlation energies}
733: Another approach to replacing the inhomogeneous electron density
734: by a homogeneous fluid-density is the
735: local-density approximation (LDA) \cite{kohn}.
736: Here a uniform density corresponding to each
737: local density $n(r)$ is invoked.
738: Thus a local-density parameter
739: $\bar{r}_s(r)$ is defined by
740: \begin{equation}
741: \label{rsLDA}
742: \frac{1}{\pi\bar{r_s}^2}=
743: \frac{\rho(r)}{\pi{r_s}^2}\quad \Longrightarrow \quad
744: \bar{r}_s(r)=r_s e^{\alpha r^2} \sqrt{\frac{\pi}{2\alpha}}.
745: \end{equation}
746: Hence, knowing the exchange-correlation
747: energy density $e_{xc}$ for a homogeneous
748: system, the exchange-correlation energy
749: of the inhomogeneous system is
750: given by
751: \begin{equation}\label{xclda}
752: E_{xc}^{LDA}=\int d^2r \, e_{xc}[\bar{r}_s(r)]\,\rho(r)
753: \end{equation}
754: Just as in the CDA, the LDA demands the evaluation
755: of $E_{c}$ at densities which are beyond
756: the range of the standard fits. Thus LDA needs
757: $r_s(r)\sim 300$ to 5000 at $r_s=20$, i.e., a
758: little less extreme than the CDA.
759: Hence, some of the shortcomings of the LDA may
760: also be due to poorly known correlation energies at
761: the exceptionally high $r_s$ values that are
762: required.
763: The LDA can be further improved by including
764: gradient corrections. However, we have not
765: included them in this study.
766: %
767: \subsection{Minimization of the total energy $E_T$}
768: \label{perfmin}
769: %
770: We have now all the energy contributions
771: needed to calculate the ground-state energy of a
772: perfect two-dimensional
773: (i.e., $\delta$-thin) Wigner crystal at a given value of
774: the density parameter
775: $r_s$.
776: The energy minimum with respect to $\alpha$
777: is found to be insensitive to the polarization of
778: the lattice. This in agreement with previous studies
779: \cite{attac,rapisarda,tanatceper,zhu}.
780: \begin{table}
781: \caption{Results of energy minimization for a hexagonal
782: lattice. The Madelung energy $E_M=-1.106103/r_s$
783: has been subtracted out from the total energy.
784: The CDA and LDA results are compared with the GFMC
785: calculations of Tanatar
786: \& Ceperley, T-C \cite{tanatceper}
787: % and Rapisarda
788: %\& Senatore $E_{rs}$ \cite{rapisarda}).
789: The energies are in $10^{-2}$ atomic
790: units. }
791: \label{tabpol}
792: \begin{ruledtabular}
793: \begin{tabular}{ccccccc}
794: $\;\;\;r_s\to$ & 20 & 30 & 40 & 50 & 60 \\
795: \hline
796: CDA &0.9247 & 0.4824 &0.3059 & 0.2156 &0.1625 \\
797:
798: LDA &0.9404 & 0.4899 &0.3102 & 0.2185 &0.1645 \\
799:
800: T-C &0.9167 & 0.4983 &0.3234 & 0.2313 &0.1758 \\
801:
802: %$E_{rs}$ \rule{0pt}{2.6ex} & $-4.617(1)$ & $-3.191(3)$ & $-2.4430(1)$ & $-1.9817(1)$ &
803: %$-1.44097(1)$ & $-1.0245(1)$ \\
804: \end{tabular}
805: \end{ruledtabular}
806: \end{table}
807: In Table \ref{tabpol}, we give the energy correction to the
808: Madelung energy
809: obtained by the minimization of $E_T$, using the CDA or the LDA
810: for evaluating the exchange-correlation effects,
811: together with the results of previous work \cite{tanatceper}.
812: QMC results by Rapisarda and Senatore\cite{rapisarda} are
813: very similar to those of Tanatar et al., and the agreement
814: is similar.
815: The optimal $\alpha$ which minimizes the energy
816: is found to be given by
817: $\alpha=a/r_s^{3/2}$
818: with a=0.639 for both CDA and LDA approaches. Here
819: we note that Nagy's method\cite{nagy} predicts
820: a value $\alpha_{ng}=0.5/r_s^{3/2}$ as a lower bound.
821: A crucial test of the accuracy of the CDA and LDA would lie
822: in their ability to predict the liquid$\to$solid phase transition.
823: This is addressed in section~\ref{wpt-sec}.
824: The total energy can be represented by:
825: %
826: %
827: %
828: \begin{equation}
829: \label{etot-fit}
830: E_T(r_s)=\frac{a_1}{r_s}+\frac{a_2}{{r_s}^{3/2}}+\frac{a_3}{{r_s}^2}+O({r_s}^{-5/2})\quad
831: r_s\gg 1
832: \end{equation}
833: where $a_1=-1.106103$ is the Madelung constant and $a_2$ is the
834: zero-point energy of the lattice.
835: We determined the coefficient $a_3$ by a least-square fit.
836: The results are summarized in Table~\ref{tab-123},
837: together with previous results.
838: %
839: \begin{table}
840: \caption{Coefficients $a_1-a_3$ in Eq.~\ref{etot-fit}
841: fitting the CDA and LDA
842: total energy (for the range $r_s$=20 to 100) are compared
843: with previous work.
844: }
845: \label{tab-123}
846: \begin{ruledtabular}
847: \begin{tabular}{ccccccc}
848: &CDA & LDA &BM\cite{bonsall} & RS\cite{rapisarda} & TC\cite{tanatceper} \\
849: \hline
850: $-a_1$ & 1.1061 & 1.10610 & 1.1060 & 1.104715 &1.10610 \\
851: $a_2$ & 0.8142 & 0.8142 & 0.8142 & 0.7947 &0.8142 \\
852: $a_3$ & 0.2456 & 0.1194 & ... & 0.07338 &0.0254 \\
853: \end{tabular}
854: \end{ruledtabular}
855: \end{table}
856: %
857: These results justify
858: our use of the CDA and the LDA for evaluating the
859: total energy of quasi-2D Wigner crystal phases for which there
860: are no QMC calculations as yet.
861: %
862:
863: \section{Influence of the thickness}
864: \label{gau-thick}
865: We consider a
866: quasi-2D electron crystal where each electron is localized
867: at each lattice site with a gaussian distribution centered on each site
868: in the $(x,y)$-plane,
869: while the $z$-extension may typically have the form of a Fang-Howard density.
870: As before, such $z$-distributions can be replaced by a constant-density
871: form for ease of calculations. Also, we assume that the 2D layers
872: are in HIGFETS, and as such the FH-parameter $b$ is automatically
873: specified (via Eq.~\ref{fh2w} )when the $r_s$ parameter defining
874: the 2D-layer density is specified.
875:
876: The kinetic energy and the harmonic energy of the quasi-2D
877: system are still given by $E_K(\alpha)=N\alpha$ since
878: this is a result of the assumed gaussian form of the wavefunction.
879: However, the simple Coulomb potential $1/r$ has changed to
880: $F(r)/r$ where $F(r)$ is the form factor arising from the
881: subband distribution. The Wigner-exchange energy, i.e.,
882: the exchange between two localized electrons is now even
883: weaker than in Eq.~\ref{excdef}. Hence this type of
884: exchange is totally negligible.
885:
886: \subsection{The evaluation of $E_{xc}$ for thick-2D layers using
887: CDA and LDA.}
888: As described in Eq.~\ref{fh2w}, the z-distribution
889: is mapped onto a uniform slab of width $a$; in HIGFETS
890: this is
891: directly related to the $r_s$ parameter in the 2-D plane.
892: The inhomogeneous 2-D distribution in the plane is also
893: mapped onto a homogeneous distribution via the CDA
894: as in Eqs.~\ref{rsCDA}
895: or as in Eq.~\ref{rsLDA} for the LDA.
896: Both CDA and LDA require a knowledge of the $E_{xc}(r_s,a)$
897: for quasi-2D uniform systems with a layer width $a$. Parametrized forms
898: of the exchange energy and the correlation energy
899: of uniform quasi-2D electron fluids have been given by
900: Dharma-wardana\cite{ssc}. The exchange energy $E_x(r_s,a)$
901: is given in the form $E_x(r_s,\zeta)Q(r_s,a,\zeta)$ where
902: $E_x(r_s,\zeta)$ is the well-known exchange energy of the
903: $\delta$-thin system, while $Q(r_s,a,\zeta)$ is a form factor.
904: The correlation energy of quasi-2D layers in HIGFETS
905: is given in Ref.~\cite{ssc}
906: as an interpolation involving a form for electron ``rods'' interacting
907: via a logarithmic potential (as is the
908: case for small $r_s$), and for 3D-like electrons when $r_s$,
909: and the thickness $a$
910: become large. The Wigner crystallization regime involves
911: the latter regime. For details of these parametrizations,
912: the reader is referred to Ref.~\cite{ssc}. Since the WT involves
913: small energy differences, we have actually done an explicit
914: calculation instead of using the fits.
915: %
916:
917: \begin{figure}
918: \includegraphics*[width=8cm,height=6cm]{wpt.eps}
919: \caption{(Color online) Comparison of liquid and solid-phase energies.
920: $(E-E_M){r_s}^{3/2}$ where
921: $E_M=-1.106103/r_s$ and $E$ is the unpolarized or
922: fully-polarized fluid energy, or the solid energy
923: $E_{cda}$, $E_{lda}$ or from QMC.}
924: \label{figwpt}
925: \end{figure}
926:
927: %
928: \subsection{Minimization of the total energy $E_T$}
929:
930: As in Sec.\ref{perfmin}, we minimize the total energy as a function of
931: $\alpha$ for a given $r_s$. Here, the
932: energy is more sensitive to the spin-polarization
933: $\zeta$ than in the perfect crystal even
934: if the difference is very small. The unpolarized crystal is more stable
935: than the polarized one. So we present results for the unpolarized
936: system.
937: The values of $\alpha$ which minimizes the energy can also be fitted by the
938: same form as in Sec.\ref{perfmin}. We obtain
939: \begin{equation}
940: \alpha_{cda}=\frac{0.619}{{r_s}^{3/2}} \quad\mathrm{and}\quad
941: \alpha_{lda}=\frac{0.627}{{r_s}^{3/2}}
942: \end{equation}
943: We have also fitted the total energy. Here the Madelung energy is
944: the $E_{Cou}$ given in Eq.(\ref{thick-madelung}) and we use the
945: usual expansion in inverse $r_s^{3/2}$ etc.
946: %{\setlength\arraycolsep{2pt}
947: \begin{eqnarray}
948: E_{T}^{cda} & = &
949: E_{Cou}(r_s)+\frac{0.68597}{{r_s}^{3/2}}+\frac{0.321652}{{r_s}^2} \\
950: E_{T}^{lda} & = &
951: E_{Cou}(r_s)+\frac{0.708977}{{r_s}^{3/2}}+\frac{0.357242}{{r_s}^2}
952: \end{eqnarray}
953: We remark that the total energy has a minimum as a function of $r_s$. This
954: minimum is located around $r_s\sim 26$ and its value is $\sim -0.011$ a.u.
955: A comparison of liquid phase and Wigner crystal in HIGFETS is given in
956: Table~\ref{tabpol2}. These total energies include the $2a/3r_s^2$ correction
957: arsing from the interaction of the quasi-2D layer with the unifrom background,
958: as discussed in subsection~\ref{gau-thin}. Since this depends on the
959: layer thickness $a$, this correction does not occur in the ideal 2D
960: system.
961: %
962: \begin{table}
963: \caption{Results of energy minimization for a hexagonal
964: lattice and comparison with the unpolarized
965: liquid phase energy $E_L$.
966: The energies are measured in $10^{-3}$ atomic
967: units.}
968: \label{tabpol2}
969: \begin{ruledtabular}
970: \begin{tabular}{ccccccc}
971: \hline
972: $r_s$ \rule{0pt}{2.6ex} &15& $20$ & 30& $50$ \\
973: \hline
974: $E_{cda}$ \rule{0pt}{2.6ex}&-6.7255 & -10.1581&-10.8576 & -9.1112
975: \\
976: $E_{lda}$ \rule{0pt}{2.6ex}&-6.5036 & -9.8169&-10.7782 & -9.0306
977: \\
978: $E_{L}$ \rule{0pt}{2.6ex}&-7.1324 & -10.0249 &-10.5995& -8.8939
979: \\
980: \end{tabular}
981: \end{ruledtabular}
982: \end{table}
983: %
984: %
985: %
986: \subsection{Phase transition Liquid$\to$ Wigner crystal.}
987: %\label{transperf}
988: \label{wpt-sec}
989: According to quantum Monte Carlo simulations, the phase transition between a
990: $\delta$-thin 2D electron liquid and a 2D electron Wigner crystal occurs
991: around $r_s=
992: 35\pm5$.
993: With our methods we find a transition for $r_s=32$ using the CDA and
994: $r_s=38$ using the LDA.
995: %
996:
997: In Figure \ref{figwpt}, we show the phase diagram of the system (in
998: order to have a clear display we present $(E-E_M){r_s}^{3/2}$ where
999: $E_M=-1.106103/r_s$ the Madelung energy). The fluid phase energy is calculated
1000: using the fit given by\cite{attac}. Unlike in the $\delta$-thin
1001: 2D system, the total energy contains the term $\Delta_{be}=2a/3r_s^2$
1002: arising from the interaction with the unifrom background,
1003: (see subsection~\ref{gau-thin}). This has been removed from both the
1004: liquid and the solid phase
1005: energies as this improves the clarity of the display.
1006:
1007:
1008: \begin{figure}
1009: %\begin{center}
1010: %\includegraphics*[width=7.5cm,height=6cm]{phaset.eps}
1011: \includegraphics*[width=6cm,height=7.5cm]{r1p5.eps}
1012: \caption{(Color online) $(E-E_M){r_s}^{3/2}$ where
1013: $E_M=-1.106103/r_s$ and $E$ is the unpolarized or
1014: fully-polarized fluid energy, or the solid energy
1015: $E_{cda}$ or $E_{lda}$. The spin-phase transition in
1016: the fluid is labeled SPT. The onset of the Wigner crystal
1017: in CDA and LDA are indicated by arrows.
1018: }
1019: \label{phaset}
1020: %\end{center}
1021: \end{figure}
1022:
1023:
1024: These results
1025: tend
1026: to show that both LDA and CDA provide an adequate evaluation of $E_{xc}$ for
1027: electron solids, especially when we recall that the $E_c$ at $r_s$
1028: values (200-8000),
1029: way outside the fit range ($r_s\le 40$), are needed in the CDA and the
1030: LDA evaluations.
1031: Figure~\ref{phaset} displays the phase transitions
1032: in the quasi-2D HIGFET system. The Wigner transition occurs at
1033: $r_s=15.5$ for the CDA, and $r_s=20.5$ for the LDA, i.e,
1034: before the spin-phase transition (marked SPT in the figure)
1035: of the liquid phase. Since the $\delta$-thin 2D layer is
1036: expected to have a WT near $r_s\sim 35$, the thickness effect
1037: has brought the WT to smaller $r_s$ values. It should be noted
1038: that if correlation effects are neglected, the WT occurs at very low
1039: $r_s$. Hence the shift of the WT to low $r_s$ is a consequence
1040: of the reduced correlations in the quasi-2D system.
1041:
1042:
1043:
1044: \section{Conclusion}
1045:
1046: We have investigated the crystallization of electrons under their
1047: own interaction in infinitely-thin 2D electron layers, known
1048: as Wigner crystallization, as well as
1049: in 2D layers with a finite width. The case of
1050: infinitely-thin 2D electron layers has been extensively studied
1051: in the past, and provides a bench mark to test our methods
1052: for replacing inhomogeneous electron densities by equivalent
1053: uniform-density models. Detailed Hartree-Fock
1054: calculations with large plane-wave basis sets are shown to
1055: be closely equivalent to the single-gaussian harmonic
1056: lattice calculations. We show that the constant-density approximation,
1057: CDA, as well as the local-density approximation, LDA, successfully
1058: recover the total energy as well as the correlation energies of
1059: the infinitely-thin 2D electron crystal. In particular, these
1060: models predict a liquid$\to$solid phase-transition in the
1061: range $30<r_s<40$, in good agreement with Quantum Monte Carlo
1062: simulations. When these methods are applied to quasi-2D
1063: layers with the thickness as in HIGFETS, the weakened
1064: Coulomb correlations move the Wigner transition towards
1065: high densities. The LDA and the CDA predict a
1066: Wigner transition near $r_s\sim$15 to 21.
1067:
1068:
1069: \appendix*
1070: \section{Evaluation of the electrostatic energy and lattice sums.}
1071: The expression for the electron-electron Coulomb energy, Eq.~\ref{delta-ee}
1072: can be rewritten, using the $\theta$ Jacobi function technique as,
1073: {\setlength\arraycolsep{2pt}
1074: \begin{eqnarray}\label{eq1}
1075: E_{ee}(q) & = & \frac{N}{2\sqrt{\pi}} \sum_{j}\int_{0}^{\infty}
1076: dy\,y^{-1/2}f(y)e^{-y|\mathbf{\rho_j}|^2-i\mathbf{q\cdot\rho_j}}\nonumber \\
1077: & & \left(\prod_{\alpha} \theta\left [ \frac{a_{\alpha}}{2\pi}
1078: (2\rho_{j}^{\alpha}y+iq_{\alpha});\frac{y {a_{\alpha}}^2}
1079: {\pi}\right] - \delta_{i,0}\right)\nonumber
1080: \end{eqnarray}}
1081: where $\delta_{i,0}$ is the Kronecker symbol,
1082: because when $\mathbf{\rho_j}=0$, we must have $(m,n)\ne (m',n')$.
1083:
1084: \noindent The advantage of introducing the
1085: Jacobi function $\theta(z,X)$ is that it
1086: converges well for large $X$
1087: and we are also able to obtain convenient
1088: well-convergent formulas for the small-X
1089: region by applying the transformation (Eq.(\ref{theta2}))
1090: from which the Coulomb singular part at $q\rightarrow 0$
1091: can be rigorously extracted. Thus, $E_{ee}(q)$ obtained in Eq.(\ref{eq1}) can be
1092: separated into a large $y$ part and a small $y$ part given by
1093: \begin{equation}
1094: E_{ee}(q)=E_{ee}^{>}(q)+E_{ee}^{<}(q)
1095: \end{equation}
1096: where
1097: \begin{widetext}
1098: \begin{equation}\label{eee>}
1099: E_{ee}^{>}(q) = \frac{N}{2\sqrt{\pi}} \sum_{j}
1100: \int_{y_0}^{\infty}dy\,y^{-1/2}
1101: f(y)e^{-y|\mathbf{\rho_j}|^2-i\mathbf{q\cdot\rho_j}}
1102: \left(\prod_{\alpha=x,y} \theta\left [ \frac{a_{\alpha}}
1103: {2\pi}(2\rho_{j}^{\alpha}y+iq_{\alpha});
1104: \frac{y {a_{\alpha}}^2}{\pi}\right] - \delta_{i,0}\right)
1105: \end{equation}
1106: and
1107: {\setlength\arraycolsep{2pt}
1108: \begin{eqnarray}\label{eee<}
1109: E_{ee}^{<}(q) & = & \frac{N}{2\sqrt{\pi}}
1110: \sum_{j} \int_{0}^{y_0}dy\,y^{-1/2}
1111: f(y)e^{-y|\mathbf{\rho_j}|^2-i\mathbf{q\cdot\rho_j}}
1112: \left(\prod_{\alpha=x,y} \theta\left [ \frac{a_{\alpha}}
1113: {2\pi}(2\rho_{j}^{\alpha}y+iq_{\alpha});\frac{y {a_{\alpha}}^2}{\pi}
1114: \right] - \delta_{i,0}\right) \nonumber \\
1115: & = & \frac{N\sqrt{\pi}}{2 a_1
1116: a_2}\sum_{j}\int_{0}^{y_0}dy\,y^{-3/2}f(y)e^{-\frac{q^2}{4y}}
1117: \left(\prod_{\alpha=x,y} \theta\left
1118: [ -i\frac{2\rho_{j}^{\alpha}y+iq_{\alpha}}
1119: {2a_{\alpha}y};\frac{\pi}{y{a_{\alpha}}^2}
1120: \right]-1+1\right)\nonumber\\
1121: & & -\frac{N}{2\sqrt{\pi}}\int_{0}^{y_0} dy
1122: \,y^{-1/2}f(y) \nonumber\\
1123: & = & \frac{N\sqrt{\pi}}{2 a_1
1124: a_2}\sum_{j}\int_{0}^{y_0}dy\,y^{-3/2}f(y)
1125: e^{-\frac{q^2}{4y}} \left(\prod_{\alpha=x,y}
1126: \theta\left [ -i\frac{2\rho_{j}^{\alpha}y+iq_{\alpha}}
1127: {2a_{\alpha}y};\frac{\pi}{y{a_{\alpha}}^2}\right ]-1\right)\nonumber\\
1128: & & -\frac{n_l N\sqrt{\pi}}{2 a_1
1129: a_2}\int_{y_0}^{\infty}dy\,y^{-3/2}f(y)
1130: e^{-\frac{q^2}{4y}} -\frac{N}{2\sqrt{\pi}}\int_{0}^{y_0}
1131: dy\,y^{-1/2}f(y)+E_{ee}^{hom}(q)
1132: \end{eqnarray}}
1133: \end{widetext}
1134: where $n_l$ is the number of sublattices ($a_1 a_2 /n_l = \pi {r_s}^2$). The
1135: results of the calculation are independent of the value of $y_0>0$;
1136: nevertheless, we choose it such
1137: that the sums $\theta$ converge fast, and we have
1138: \begin{equation}\label{eeehom}
1139: E_{ee}^{hom}(q)=\frac{n_l N\sqrt{\pi}}{2 a_1 a_2}
1140: \int_{0}^{\infty}dy\,y^{-3/2}f(y)e^{-\frac{q^2}{4y}}
1141: \end{equation}
1142: In order to complete the calculation of $E_{Cou}$, we need to discuss the
1143: form of $\eta(z)$. In their article \cite{fuji2}, Fujiki and Geldart use the
1144: Fang-Howard density $\eta_{fh}(z)=\frac{1}{2 b^3}z^2 e^{-\frac{z}{b}}$ (Figure
1145: \ref{fang}). As already discussed we replace the FH distribution
1146: by the equivalent CDA, i.e., we use a constant density
1147: slab of width $a=16b/3$.
1148: \noindent With this homogeneous form of density
1149: {\setlength\arraycolsep{2pt}
1150: \begin{eqnarray}
1151: f(y) & = & {\bar{\eta}}^2 \int_{0}^{a} dz \int_{0}^{a} dz'
1152: e^{-y(z-z')^2}\nonumber\\
1153: & = & {\bar{\eta}}^2 \frac{(e^{-a^2 y}+a\sqrt{\pi
1154: y}\,\mathrm{erf} (a\sqrt{y})-1)}{y}
1155: \end{eqnarray}}
1156: We see here an advantage of the constant density
1157: mapping to density $\bar{\eta}$, the analytic expression
1158: of $f(y)$ being quite simple.
1159:
1160: In Eq.(\ref{eed}), we replace $\eta(z')$ by its expression
1161: {\setlength\arraycolsep{2pt}
1162: \begin{eqnarray}\label{eed2}
1163: E_{ed}(q) & = & -\frac{N}{q{r_s}^2}\frac{e^{-q b_d}}{a
1164: q}(1-e^{-aq})\nonumber\\
1165: E_{ed}(q\rightarrow 0) & = & -\frac{N}{{r_s}^2}
1166: \left(\frac{1}{q}-\frac{a}{2}-b_d+O(q)\right)
1167: \end{eqnarray}}
1168: We recall that $b_d$ is positive or zero, and
1169: gives the location of the donor layer at $z=-b_d$.
1170: Now, in Eq.(\ref{eeehom}), we use the definition of $f(y)$.
1171: {\setlength\arraycolsep{2pt}
1172: \begin{eqnarray}\label{eeehom2}
1173: E_{ee}^{hom}(q) & = & \frac{N}{q{r_s}^2}\frac{2}{q
1174: a^2}\left(a-\frac{1}{q}(1-e^{-aq})\right)\nonumber\\
1175: E_{ee}^{hom}(q\rightarrow 0) & = & \frac{N}{{r_s}^2}
1176: \left(\frac{1}{q}-\frac{a}{3}+O(q)\right)
1177: \end{eqnarray}}
1178:
1179: Now, we will use Eq.(\ref{edd}), (\ref{eed2}),
1180: (\ref{eee>}), (\ref{eee<}),
1181: (\ref{eeehom2}) in Eq.(\ref{ecoul}) to calculate
1182: the Coulomb energy for different
1183: types of lattices and for different thicknesses.
1184: We can see that the expression of $E_{Cou}$ is
1185: dependent on the parameter $b_d$. Since this is a constant
1186: contribution, we set $b_d=0$ and focus on the part
1187: which depends only on the geometry
1188: of the lattice and on its thickness.
1189:
1190:
1191:
1192: \begin{thebibliography}{99}
1193: \bibitem{ando} T. Ando, A. B. Fowler, F. Stern,
1194: Rev. Mod. Phys. \textbf{54}, 437(1982)
1195: \bibitem{krav}
1196: S. V. Kravchenko and M. P. Sarachik, Rep. Prog. Phys. {\bf 67}, 1 (2004)
1197: \bibitem{attac}
1198: C. Attaccalite, S. Moroni, P. Gori-Giorgi, G. B. Bachelet, Phys. Rev. Lett
1199: \textbf{88}, 256601(2002)
1200: %
1201: \bibitem{rapisarda} F. Rapisarda, G. Senatore, Aust. J. Phys. \textbf{49},
1202: 161(1996)
1203: %
1204: \bibitem{tanatceper}
1205: B. Tanatar, D. M. Ceperley, Phys. Rev. B \textbf{39}, 5005 (1989)
1206: %
1207: \bibitem{ceper}
1208: D. Ceperley, Phys. Rev. B
1209: \textbf{18}, 3126 (1978)
1210: %
1211: \bibitem{bonsall} L. Bonsall, A. A. Maradudin, Phys. Rev. B \textbf{15},
1212: 1959 (1977)
1213: %
1214: \bibitem{iyakutti}
1215: R. Rajeswara Palanichamy and K. Iyakutti,
1216: Int. J. Quant. Chem. {\bf 102}, 112 (2005)
1217: %
1218: \bibitem{prl2}
1219: Fran\c{c}ois Perrot and M. W. C. Dharma-wardana, Phys. Rev. Lett. {\bf 87},
1220: 206404 (2001)
1221: %
1222: \bibitem{ssc} M. W. C. Dharma-wardana, Solid State Com. {\bf 136}, 76 (2005);
1223: cond-mat/503246;
1224: cond-mat/506816.
1225: \bibitem{tan}
1226: Y.-W Tan, J. Zhu, H. L. Stormer, L. N. Pfeiffer, K. N. Baldwin, and
1227: K. W. West, cond-mat/0412260; Phys. Rev. Lett. {\bf 94}, 16405, (2005).
1228: \bibitem{stormer}
1229: J. Zhu, H. L. Stormer, L. N. Pfeiffer, K. W. Baldwin, and K. W. West,
1230: Phys. Rev. Lett. {\bf 90}, 56805 (2003)
1231: \bibitem{asgari}
1232: R. Asgari, B. Davoudi, M. Polini, M. P. Tosi, G. F. Giuliani, and
1233: G. Vignale, Phys. Rev. B {\bf 71}, 45323 (2005);
1234: cond-mat/0412665
1235: \bibitem{zhang}
1236: Y. Zhang and S. Das Sarma, Phys. Rev. B {\bf 71},45322 (2005);
1237: cond-mat/0312565
1238: \bibitem{morsen}
1239: S. De Palo, M. Botti, S. Moroni, and G. Senatore,
1240: Phys. Rev. Lett. {\bf 94}, 226405 (2005); cond-mat/0410145
1241: %
1242: \bibitem{fuji2}
1243: N. M. Fujiki, D. J. W. Geldart, Phys. Rev. B
1244: \textbf{46}, 9634(1992)
1245: %
1246: \bibitem{trail} J. R. Trail, M. D. Towler, R. J. Needs,
1247: Phys. Rev. B \textbf{68}, 045107 (2003)
1248: \bibitem{nagy} I. Nagy, Phys. Rev. B \textbf{60}, 4404 (1999)
1249: \bibitem{polini} M. Polini, G. Sica, B. Davoudi, M. P. Tosi, J. Phys.
1250: Cond. Mat.
1251: {\bf 13}, 3591 (2001)
1252: \bibitem{ggsavin}
1253: P. Gori-Giorgi and A. Savin, Phys. Rev. A {\bf 71}
1254: 32513 (2005); cond-mat/0411179
1255:
1256: \bibitem{fuji1}
1257: N. M. Fujiki, K. De'Bell, D. J. W. Geldart,
1258: Phys. Rev. B \textbf{36}, 8512 (1987)
1259: \bibitem{kohn}
1260: W. Kohn, P. Vashishta, {\i Theory of the inhomogeneous electron gas},
1261: edited by S. Lundqvist and N. H. March, Plenum Press (1983)
1262: %
1263: \bibitem{thank-trail}
1264: We thank Dr. Trail et al., for kindly supplying
1265: us their tabulation of the Hartree-Fock energy of the 2D
1266: wigner crystal as a function of $r_s$.
1267: %
1268: \bibitem{trail2}
1269: N. D. Drummond, Z. Radnai, J. R. Trail, M. D. Towler, and
1270: R. J. Needs, Phys. Rev. B {\bf 69}, 85116 (2004); Drummond et al.
1271: have come to a similar conclusion about the Hartree-Fock energy
1272: of the 3D-Wigner crystal.
1273: %
1274: \bibitem{zhu} X. Zhu, S. G. Louie, Phys. Rev. B \textbf{52}, 5863 (1995)
1275: %\bibitem{nagy} I.Nagy, Phys. Rev. B \textbf{60}, 4404(1999)
1276: %\bibitem{trail} J.R.Trail, M.D.Towler, R.J.Needs, \emph{Unrestricted
1277: % Hartree-Fock theory of Wigner crystal}, Phys. Rev. B \textbf{68}, 045107(2003)
1278: \bibitem{gori} P. Gori-Giorgi, J. P. Perdew,
1279: Phys. Rev. B \textbf{69}, 041103 (2004)
1280: \bibitem{perdew} J. P. Perdew, Y. Wang, Phys. Rev. B
1281: \textbf{45}, 13244 (1992)
1282: \end{thebibliography}
1283:
1284:
1285:
1286:
1287:
1288: \end{document}
1289:
1290:
1291:
1292:
1293:
1294:
1295:
1296:
1297:
1298:
1299:
1300: