cond-mat0303474/hm.tex
1: \documentstyle[aps,prb,psfig,multicol]{revtex}
2: 
3: \begin{document}
4: 
5: \def\be{\begin{eqnarray}}
6: \def\ee{\end{eqnarray}}
7: \def\nn{\nonumber}
8: \def\cad{c_A^\dagger}
9: \def\ca{c_A}
10: \def\cbd{c_B^\dagger}
11: \def\cb{c_B}
12: 
13: \title{Spin-Orbit Coupling and Time-Reversal Symmetry in Quantum Gates}
14: 
15: \author{D. Stepanenko and N.E. Bonesteel} 
16: \address
17: %\affiliation
18: {Department of
19: Physics and National High Magnetic Field Laboratory, Florida State
20: University, Tallahassee, FL 32310} 
21: 
22: \author{D.P. DiVincenzo and G. Burkard}
23: \address
24: %\affiliation
25: {IBM Research Division, T.J. Watson Research Center, Yorktown
26: Heights, NY 10598}
27: 
28: \author{Daniel Loss}
29: \address
30: %\affiliation
31: {Department of Physics and Astronomy, University of Basel, Klingelbergstrasse
32: 82, CH-4056 Basel, Switzerland}
33: 
34: \maketitle
35: 
36: \begin{abstract}
37: We study the effect of spin-orbit coupling on quantum gates produced
38: by pulsing the exchange interaction between two single electron
39: quantum dots. Spin-orbit coupling enters as a small spin precession
40: when electrons tunnel between dots. For adiabatic pulses the resulting
41: gate is described by a unitary operator acting on the four-dimensional
42: Hilbert space of two qubits. If the precession axis is fixed,
43: time-symmetric pulsing constrains the set of possible gates to those
44: which, when combined with single qubit rotations, can be used in a
45: simple CNOT construction.  Deviations from time-symmetric pulsing
46: spoil this construction. The effect of time asymmetry is studied by
47: numerically integrating the Schr\"odinger equation using parameters
48: appropriate for GaAs quantum dots. Deviations of the implemented gate
49: from the desired form are shown to be proportional to dimensionless
50: measures of both spin-orbit coupling and time asymmetry of the pulse.
51: 
52: \end{abstract}
53: 
54: \pacs{}
55: 
56: 
57: \begin{multicols}{2}
58:  
59: 
60: \section{Introduction}
61: 
62: A promising proposal for building a solid-state quantum computer is
63: based on the notion of using electron spins trapped in quantum dots as
64: qubits.\cite{loss} In such a device, two-qubit quantum gates would be
65: carried out by turning on and off the exchange interaction between
66: spins on neighboring dots through suitable pulsing of gate voltages.
67: 
68: When performing such a quantum gate, if nonadiabatic errors
69: \cite{burkard99_1,burkard99_2,hu} can be safely ignored,
70: \cite{schliemann} both the initial and final states of the two dots
71: will be in the four-dimensional Hilbert space of two qubits. In the
72: absence of spin-orbit coupling, and neglecting the dipolar interaction
73: between spins, the unitary transformation resulting from such a pulsed
74: exchange gate will necessarily have the form
75: %
76: \be U = \exp -i \lambda {\bf S}_A \cdot {\bf S}_B,
77: \label{isotropic}
78: \ee 
79: %
80: where $\lambda$ is a dimensionless measure of the pulse strength.
81: This simple isotropic form is a consequence of symmetry --- if spin
82: and space decouple exactly, as they do in the nonrelativistic limit,
83: then the system is perfectly isotropic in spin space.  Up to an
84: irrelevant overall phase the gates (\ref{isotropic}) are the most
85: general unitary operators with this symmetry acting on a two-qubit
86: Hilbert space.
87: 
88: These isotropic exchange gates are useful for quantum computation.  In
89: conjunction with single qubit rotations, they can be used in a simple
90: construction of a controlled-not (CNOT) gate.\cite{loss} It has also
91: been shown that, even without single qubit rotations, isotropic
92: exchange gates can be used for universal quantum computing with proper
93: encoding of logical qubits.\cite{bacon,divincenzo}
94: 
95: When the effects of spin-orbit coupling are included, well-isolated
96: single electron dots will have a two-fold Kramers degeneracy and so
97: can still be used as qubits.  However, when carrying out a quantum
98: gate the total spin will no longer be a good quantum number. As a
99: result there will inevitably be corrections to the isotropic exchange
100: gates (\ref{isotropic}).  Motivated by this fact, a number of authors
101: have considered {\it anisotropic} gates of the form
102: %
103: \be 
104: U &=& \exp -i \lambda \bigl({\bf S}_A \cdot
105: {\bf S}_B + {\mbox{\boldmath{$\beta$}}} \cdot({\bf S}_A \times {\bf S}_B)\nn\\
106: &&~~~~~~~~~~+
107: \gamma( 
108: {\bf S}_A \cdot {\bf
109: S}_B
110: -
111: (\hat{\mbox{\boldmath{$\beta$}}}
112: \cdot{\bf S}_A)
113: (\hat{\mbox{\boldmath{$\beta$}}}\cdot{\bf S}_B)\bigr),
114: \label{axial} 
115: \ee 
116: %
117: and shown that they have several useful properties.  For example, in
118: Ref.~\onlinecite{burkard01} it was shown that the CNOT construction of
119: Ref.~\onlinecite{loss} is robust against anisotropic corrections of
120: the form appearing in (\ref{axial}). It has also been shown that, when
121: combined with a controllable Zeeman splitting, the gates (\ref{axial})
122: form a universal set.\cite{wu}
123: 
124: The anisotropic terms which appear in (\ref{axial}) are not the most
125: general corrections to (\ref{isotropic}) which can occur when carrying
126: out an exchange gate in the presence of spin-orbit coupling.  It is
127: therefore important to ask under what conditions these corrections can
128: be restricted to have this desired form. The key observation
129: motivating the present work is that, up to an irrelevant overall
130: phase, the gates (\ref{axial}) are the most general two-qubit quantum
131: gates which are both axially symmetric, i.e. symmetric under rotations
132: about an axis parallel to the vector ${\mbox{\boldmath{$\beta$}}}$ in
133: spin space, and symmetric under time reversal (${\bf S}_\mu
134: \rightarrow -{\bf S}_\mu$, $\mu = A,B$).  It follows that if these
135: symmetries can be maintained throughout the gate operation, and
136: provided nonadiabatic errors can be neglected, the resulting quantum
137: gate is {\it guaranteed} to have the form (\ref{axial}).  Of course
138: symmetry alone cannot determine the values of $\lambda$,
139: ${\mbox{\boldmath{$\beta$}}}$ and $\gamma$.  However, in practice we
140: envision these parameters will be determined through experimental
141: calibration rather than microscopic calculation.  Therefore we
142: emphasize symmetry as a useful guiding principle.
143: 
144: In this paper we study the effect of spin-orbit coupling on
145: exchange-based quantum gates.  For concreteness we consider a system
146: of two single-electron quantum dots in GaAs. The contribution of
147: spin-orbit coupling to the exchange interaction between localized
148: spins in GaAs has been studied by Kavokin\cite{kavokin} within the
149: Heitler-London approximation, and by Gor'kov and Krotkov\cite{gorkov}
150: who derived the exact asymptotic exchange interaction between
151: hydrogen-like bound states at large separation.
152: 
153: Here we follow Ref.~\onlinecite{burkard99_1} and work within the
154: Hund-Mulliken approximation, keeping one orbital per dot, and allowing
155: double occupancy.  In this approximation, the effect of spin-orbit
156: coupling is to induce a small spin precession whenever an electron
157: tunnels from one dot to another.  The Hamiltonian governing the
158: two-dot system is therefore axially symmetric in spin space with the
159: symmetry axis being the precession axis of the spin.  If the direction
160: of the precession axis does not change while the gate is being pulsed,
161: then the resulting quantum gate will also be axially symmetric.
162: 
163: An additional useful symmetry principle, first suggested in
164: Ref.~\onlinecite{bonesteel01}, is that any time-dependent Hamiltonian
165: $H_P(t)$ which is time-reversal symmetric at all times $t$, and which
166: is then pulsed in a time-symmetric way ($H_P(t) = H_P(-t)$) will lead
167: to a gate which can be described in terms of an effective
168: time-independent Hamiltonian $H$ which is also time-reversal
169: symmetric.  Here we give a proof of this result.
170: 
171: Taken together these two results imply that, within the Hund-Mulliken
172: approximation, if the spin-orbit precession axis is fixed and
173: nonadiabatic errors can be ignored, the unitary transformation
174: produced by pulsing the exchange interaction between two quantum dots
175: will necessarily have the desired form (\ref{axial}) provided the gate
176: is pulsed in a time-symmetric way.
177: 
178: This paper is organized as follows.  In Sec.~II we derive the
179: Hund-Mulliken Hamiltonian for a double quantum dot system in the
180: presence of spin-orbit coupling.  In Sec.~III we develop an effective
181: spin Hamiltonian description which can be applied to pulsing our
182: double dot system, and we review the robust CNOT construction of
183: Ref.~\onlinecite{burkard01}.  The implications of time-symmetric
184: pulsing are then studied in Sec.~IV, and in Sec.~V we present
185: numerical results showing the effect of small time asymmetry of the
186: pulse.  Finally, in Sec.~VI we summarize the results of the paper.
187: 
188: 
189: \section{Hund-Mulliken Hamiltonian}
190: 
191: We consider a system of two laterally confined quantum dots with one
192: electron in each dot.  For concreteness we assume the dots are formed
193: in a two-dimensional electron gas (2DEG) realized in a GaAs
194: heterostructure.
195: 
196: The system is modeled by the Hamiltonian
197: %
198: \be H = T + C + H_{SO}. 
199: \label{hamiltonian}
200: \ee
201: % 
202: Here $T+C$ is the Hamiltonian studied in
203: Ref.~\onlinecite{burkard99_1}, where $T = \sum_i h_i$ with
204: %
205: \be h_i = \frac{1}{2m} \left({\bf p}_i -
206: \frac{e}{c}{\bf A}({\bf r}_i)\right)^2 + V({\bf r}_i),  \ee
207: %
208: and $C = e^2/\epsilon|{\bf r}_1 - {\bf r}_2|$ is the Coulomb repulsion
209: between electrons.  We take the 2DEG the dots are formed in to lie in
210: the $xy$ plane, and for GaAs we take $m = 0.067 m_e$ and $\epsilon =
211: 13.1$.  For completeness we include a vector potential ${\bf A} =
212: (-y,x,0)B/2$ which couples the orbital motion of the electrons to a
213: uniform magnetic field ${\bf B} = B \hat{\bf z}$.  We will see in
214: Sec.~III that this orbital coupling does not affect any of our
215: arguments based on time-reversal symmetry, while a nonzero Zeeman
216: coupling does.
217: 
218: 
219: As in Ref.~\onlinecite{burkard99_1} lateral confinement of the dots is
220: modeled by the double-well potential,
221: %
222: \be
223: V(x,y) = \frac{m\omega_0^2}{2} \left(\frac{1}{4a^2}(x^2 - a^2)^2 +
224: y^2 \right).
225: \ee
226: %
227: This potential describes two quantum dots sitting at the points $(x,y)
228: = (\pm a, 0)$. In the limit of large separation the dots decouple into
229: two harmonic wells with frequency $\omega_0$.  
230: 
231: 
232: Spin-orbit coupling enters the Hamiltonian through the term
233: %
234: \be
235: H_{SO} = \sum_{i=1,2}{\bf \Omega}({\bf k}_i) \cdot {\bf S}_i,
236: \ee
237: %
238: where $\hbar{\bf k} = {\bf p} - \frac{e}{c} {\bf A}$. Time-reversal
239: symmetry requires that ${\bf \Omega}({\bf k})$ is an odd function of
240: ${\bf k}$, ${\bf \Omega}({\bf k}) = - {\bf \Omega}(-{\bf k})$.  Thus
241: ${\bf \Omega}$ is nonzero only in the absence of inversion symmetry.
242: 
243: 
244: For definiteness, we take the 2DEG in which the dots are formed to lie
245: in the plane perpendicular to the [001] structural direction, which
246: then points along the $z$-axis.  However, we allow the $x$-axis, which
247: is parallel to the displacement vector of the two dots, to have any
248: orientation with the respect to the [100] and [010] structural axes.
249: To describe the dependence of ${\bf \Omega}$ on ${\bf k}$ it is then
250: convenient to introduce unit vectors ${\hat {\bf e}}_{[110]}$ and
251: ${\hat {\bf e}}_{[{\overline 1}10]}$ which point in the $[110]$ and
252: $[\overline 110]$ structural directions, respectively, and define
253: $k_{[110]} = {\bf k} \cdot \hat {\bf e}_{[110]}$ and $k_{[\overline
254: 110]} = {\bf k} \cdot \hat {\bf e}_{[\overline 110]}$.  We then have,
255: following Kavokin,\cite{kavokin}
256: %
257: \begin{eqnarray}
258: {\bf \Omega}({\bf k}) \simeq 
259: (f_D - f_R) k_{[110]} \hat {\bf e}_{[\overline 1 1 0]}
260: +(f_D + f_R)  k_{[\overline 110]} 
261: \hat {\bf e}_{[110]}.
262: \label{omega}
263: \end{eqnarray}
264: %
265: Here $f_D$ is the Dresselhaus contribution\cite{dresselhaus,dyakonov}
266: due to the bulk inversion asymmetry of the zinc-blende crystal
267: structure of GaAs, and $f_R$ is the Rashba contribution\cite{rashba}
268: due to the inversion asymmetry of the quantum well used to form the
269: 2DEG.  These quantities depend on details of the 2DEG confining
270: potential and so will vary from system to system.
271: 
272: It was pointed out in Ref.~\onlinecite{schliemann03} that $H_{SO}$ has
273: a special symmetry when $f_D = \pm f_R$.  This can be seen directly
274: from (\ref{omega}).  When $f_D = f_R$ ($f_D = - f_R$) the direction of
275: ${\bf \Omega}$ is independent of ${\bf k}$ and is fixed to be parallel
276: to $\hat {\bf e}_{[110]}$ ($\hat {\bf e}_{[\overline 110]}$). The full
277: Hamiltonian (\ref{hamiltonian}) is then invariant under rotations in
278: spin space about this axis.  We will see below that this special case
279: has a number of attractive features.
280: 
281: In the limit of decoupled dots, and ignoring spin-orbit coupling, the
282: single electron ground states will be the Fock-Darwin ground states
283: centered at $(x,y) = (\pm a,0)$,
284: %
285: \be \phi_{\pm a}(x,y) = \sqrt{\frac{m\omega}{\pi\hbar}} e^{ -m\omega
286: \left((x \mp a)^2 + y^2\right)/2\hbar} e^{ \pm 
287: i a y/2 l_B^2}. \ee
288: %
289: Here $\omega^2 = \omega_0^2 + \omega_{L}^2$ is the frequency of the
290: magnetically squeezed oscillator where $\omega_{L} = eB/2mc$ is the
291: Larmor frequency and $l_B = \sqrt{\hbar c/eB}$ is the magnetic length.
292: In zero magnetic field, the size of these wave functions is set by the
293: effective ``Bohr radius'' $a_B = \sqrt{\hbar/m\omega_0}$.
294: 
295: The Fock-Darwin states can be orthogonalized to obtain the Wannier
296: states
297: %
298: \be
299: \Phi_A &=& \frac{1}{\sqrt{1-2Sg -g^2}} (\phi_a - g \phi_{-a}),\\
300: \Phi_B &=& \frac{1}{\sqrt{1-2Sg -g^2}} (\phi_{-a} - g \phi_{a}),
301: \ee
302: %
303: where $S = \langle \phi_{-a} | \phi_{a} \rangle$ and $g =
304: (1-\sqrt{1-S^2})/S$.  We can then introduce second quantized operators
305: ${c^\dagger_A}_\alpha$ (${c_A}_\alpha$) and ${c^\dagger_B}_\alpha$
306: (${c_B}_\alpha$) which create (annihilate) electrons in the states
307: $\Phi_A$ and $\Phi_B$ with spin $\alpha = \uparrow,\downarrow$.
308: 
309: 
310: In the Hund-Mulliken approximation we keep one orbital per dot and
311: allow for double occupancy.  This amounts to restricting the full
312: Hilbert space of the problem to the six-dimensional Hilbert space
313: spanned by the states
314: %
315: \be |S_1 \rangle &=& \frac{1}{\sqrt{2}} (c^\dagger_{A\uparrow}
316: c^\dagger_{B\downarrow} -c^\dagger_{A\downarrow}
317: c^\dagger_{B\uparrow}) |0\rangle,\\ |S_2 \rangle &=& \frac{1}{\sqrt{2}}
318: (c^\dagger_{A\uparrow} c^\dagger_{A\downarrow}
319: +c^\dagger_{B\downarrow} c^\dagger_{B\uparrow}) |0\rangle,\\
320: |S_3\rangle &=& \frac{1}{\sqrt{2}} (c^\dagger_{A\uparrow}
321: c^\dagger_{A\downarrow} -c^\dagger_{B\downarrow}
322: c^\dagger_{B\uparrow})|0\rangle,\\ 
323: |T_-\rangle &=&
324: c^\dagger_{A\downarrow}c^\dagger_{B\downarrow} |0\rangle,\\ |T_0\rangle
325: &=& \frac{1}{\sqrt{2}} (c^\dagger_{A\uparrow} c^\dagger_{B\downarrow}
326: +c^\dagger_{A\downarrow} c^\dagger_{B\uparrow}) |0\rangle,\\
327: |T_+\rangle &=& c^\dagger_{A\uparrow} c^\dagger_{B\uparrow} |0
328: \rangle.  \ee
329: %
330: 
331: In terms of second quantized operators, the Hund-Mulliken Hamiltonian
332: acting in this space, up to an irrelevant overall additive
333: constant, can be written
334: %
335: \be H_{HM} &=& \sum_{\alpha,\beta = \uparrow,\downarrow}
336: -\left({c^\dagger_A}_\alpha (t_H \delta_{\alpha\beta} + i {\bf P} \cdot
337: {\mbox{\boldmath{$\sigma$}}}_{\alpha\beta}){c_B}_\beta+H.c.\right)\nn\\
338: &&+V\left({\bf S}_A \cdot {\bf S}_B+3/4\right)\nn\\ &&+ U_H
339: (n_{A\uparrow} n_{A\downarrow} + n_{B\uparrow} n_{B\downarrow}).  \ee
340: %
341: \noindent
342: Here
343: %
344: \be
345: {\bf S}_{\mu} = \frac{1}{2} \sum_{\alpha,\beta = \uparrow,\downarrow}
346: {c^\dagger_\mu}_\alpha {\mbox{\boldmath$\sigma$}}_{\alpha\beta} {c_\mu}_\beta
347: \ee
348: %
349: is the spin operator on site $\mu = A,B$, 
350: \be
351: V &=& \langle S_1 | C | S_1 \rangle - \langle T | C| T \rangle
352: \ee
353: %
354: is the ferromagnetic direct exchange,
355: %
356: \be
357: U_H &=& \langle S_2 | C| S_2 \rangle - \langle S_1 | C | S_1 \rangle
358: \ee
359: %
360: is the Coulomb energy cost of doubly occupying a dot,
361: and 
362: %
363: \be
364: t_H &=& \langle \Phi_A | h | \Phi_B \rangle
365: \ee
366: %
367: is the interdot tunneling amplitude.  
368: 
369: The only contribution from spin-orbit coupling is the matrix element
370: %
371: \be i{\bf P} = \langle \Phi_A | {\bf \Omega}({\bf k}) | \Phi_B
372: \rangle  = \langle \Phi_A | (p_x - \frac{e}{c}A_x)/\hbar | \Phi_B
373: \rangle
374: \mbox{\boldmath{$\eta$}}, \ee
375: %
376: where 
377: %
378: \be
379: \mbox{\boldmath{$\eta$}} &=& (f_D - f_R) \cos \theta \hat {\bf e}_{[\overline 110]}+ (f_D + f_R ) \sin \theta \hat {\bf e}_{[110]}.
380: \ee
381: %  
382: Here $\theta$ is the angle the $x$-axis makes with the [110]
383: structural direction.  This term introduces a small spin precession
384: about an axis parallel to ${\bf P}$ through an angle $\phi = 2
385: \arctan(P/t_H)$ when an electron tunnels between dots.
386: 
387: It is convenient to express the spin-orbit matrix element as ${\bf
388: P}=s{\bf l}_{SO}$ where
389: %
390: \be s=\frac{\sqrt{(f_D-f_R)^2 \cos^2 \theta + (f_D+f_R)^2 \sin^2
391: \theta }}{a_B~\hbar\omega_{0}} 
392: \label{sdef}
393: \ee
394: %
395: is a dimensionless measure of the strength of spin-orbit coupling.  As
396: stated above, $f_D$ and $f_R$ depend on details of the potential
397: confining the electron to the 2DEG. Thus $\theta$, $f_D$ and $f_R$ are
398: all parameters that in principle can be engineered to control the
399: value of $s$.  For example, if $\theta = 0$ then $s = |f_D - f_R|/(a_B
400: \hbar\omega_0)$.  Thus, for this orientation of the dots, if it is
401: possible to design a system in which $f_D = f_R$, $s$ can be made to
402: vanish. Even if such perfect cancellation cannot be achieved,
403: minimizing the difference $f_D - f_R$ will reduce $s$.
404: 
405: In what follows we leave $s$ as a free parameter.  We estimate that
406: for GaAs quantum dots $s < 0.1$ for typical parameters.\cite{kavokin}
407: The remaining contribution to the matrix element ${\bf P}$ is then
408: %
409: \be {\bf l}_{SO}=\frac{\hbar\omega_0}{2} \frac{1-g^2}{1-2Sg+g^2} \frac{d}{b}
410: e^{-d^2 b(2-1/b^2)} \hat{\mbox{\boldmath{$\eta$}}}, \ee
411: %
412: where $d = a/a_B$ is a dimensionless measure of the distance between
413: dots, $b = \sqrt{1+\omega_L^2/\omega_0^2}$, and
414: $\hat{\mbox{\boldmath{$\eta$}}} = \mbox{\boldmath{$\eta$}}/\eta$.  The
415: geometry of our model system is shown schematically in
416: Fig.~\ref{system}.
417: 
418: In what follows we envision pulsing quantum gates by varying the
419: distance $d$ between dots as a function of time.  In doing this, we
420: will assume that throughout the pulse the values of $f_D$ and $f_R$ do
421: not change.  If this is the case $s$ will be constant and all of the
422: time dependence of ${\bf P}$ will be due to ${\bf l}_{SO}$.  In
423: addition, the direction of the vector ${\bf P}$ will not change as a
424: function of time.  The Hamiltonian $H_{HM}$ will therefore be
425: invariant under rotations in spin space about a single fixed axis
426: parallel to ${\bf P}$ throughout the pulse.  We will refer to such a
427: pulse as having axial symmetry.
428:  
429: \
430: \vskip .2in
431: 
432: \begin{figure}[h]
433: \centerline{\psfig{figure=system-3.eps,height=1.2in,angle=0}}\vskip .15in
434: \caption{Sketch of the GaAs double quantum dot system considered in
435: this paper.  There is one electron per dot, and the dot separation is
436: $2a$.  The dots are taken to lie in the plane perpendicular to both
437: the [001] axis and an applied magnetic field $B$.  The displacement of
438: the dots makes an angle $\theta$ with the [110] axis.  Due to
439: spin-orbit coupling electron spins precess about an axis parallel to
440: ${\bf P}$ when tunneling between dots.}
441: \label{system}
442: \end{figure}
443: 
444: It is important to note that this axial symmetry is approximate.  In
445: general $f_D$ and $f_R$ will depend on time as the gate is pulsed,
446: though in principle the system can be engineered to minimize this
447: effect.  Also, for general $f_D$ and $f_R$ the appearance of only one
448: vector in spin space is a consequence of restricting the Hilbert space
449: to one orbital per dot.  If more orbitals are included then more
450: spin-orbit matrix elements will appear in the Hamiltonian,
451: corresponding typically to different spin-precession axes, thus
452: breaking the axial symmetry.  However, as shown above, if $f_D = \pm
453: f_R$ then the full Hamiltonian (\ref{hamiltonian}) is axially
454: symmetric -- thus for this special case all spin precession axes will
455: be parallel and axial symmetry will not just be an artifact of the
456: Hund-Mulliken approximation.  In Sec.~V we discuss the effect
457: deviations from axial symmetry will have on our results.
458: 
459: Given an axially symmetric pulse, it is convenient to take the
460: $z$-axis in spin space to be parallel to ${\bf P}$.  For this choice,
461: the states $|T_+\rangle$ and $|T_-\rangle$ decouple, each having
462: energy $V$.
463: 
464: Another useful symmetry of $H_{HM}$ is invariance under $c_{A\alpha}
465: \rightarrow c_{B,-\alpha}$ and $c_{B\alpha} \rightarrow
466: c_{A,-\alpha}$.  This transformation changes the sign of the states
467: $|S_1\rangle,\ |S_2\rangle$ and $|T_0\rangle$, while leaving
468: $|S_3\rangle$ invariant.  It follows that the state $|S_3\rangle$ also
469: decouples with energy $U_H$.  The matrix representation of $H_{HM}$ in
470: the remaining nontrivial $|T_0\rangle$, $|S_1\rangle$, $| S_2\rangle$
471: basis is then
472: %
473: \be H_{HM} = \left(
474: \begin{array}{ccc}
475: V & 0 & -2iP \\
476: 0 & 0 & -2t_H \\
477: 2iP & -2t_H & U_H
478: \end{array}
479: \right).
480: \label{hmatrix}
481: \ee 
482: 
483: \section{Effective Spin Hamiltonian}
484: 
485: We now consider pulsing the Hamiltonian $H_{HM}$ by varying the
486: distance between the dots, the barrier height, or some combination of
487: the two, in such a way that the two electron spins interact for a
488: finite period of time, but are well separated at the beginning and end
489: of the pulse.  We assume the initial state of the system is in the
490: four-dimensional Hilbert space describing two qubits, i.e. the space
491: spanned by the singly occupied states $|S_1\rangle, |T_0\rangle,
492: |T_-\rangle$ and $|T_+\rangle$.  As the pulse is carried out, the
493: eigenstates of $H_{HM}$ at any given instant in time can be grouped
494: into four low-energy states separated by a gap of order $U_H$ from two
495: high-energy states.  If the pulse is sufficiently adiabatic on a time
496: scale set by $\sim \hbar/U_H$, the amplitude for nonadiabatic
497: transitions which would leave the system in the excited state
498: $|S_2\rangle$ at the end of the pulse can be made negligibly small.
499: \cite{schliemann} If this condition holds, the final state of the
500: system can also be assumed to be in the four-dimensional Hilbert space
501: of two qubits.  We will see that this condition is easily achieved in
502: Sec.~V.
503: 
504: One way to theoretically study the effect of such a pulse would be to
505: first reduce $H_{HM}$ to an effective anisotropic spin Hamiltonian
506: acting on the four-dimensional low-energy Hilbert space and then
507: consider pulsing this effective model.\cite{bonesteel01} The problem
508: with this approach is that any such effective spin Hamiltonian will
509: only be valid if the pulse is adiabatic, not only on the time scale
510: $\hbar/U_H$, but also on the much longer time scale set by the inverse
511: of the small energy splittings within the low-energy space due to the
512: spin-orbit induced anisotropic terms.  However, it is precisely the
513: nonadiabatic transitions induced by these terms which give rise to the
514: quantum gate corrections we would like to compute.
515: 
516: Although we may not be able to define an instantaneous effective spin
517: Hamiltonian during the pulse, we can define one which describes the
518: net effect of a full pulse.  This definition amounts to parameterizing
519: the quantum gate produced by the pulse as
520: %
521: \be U = e^{ - i \tau H}, 
522: \ee
523: %
524: where $U$ acts on the four-dimensional Hilbert space of the initial
525: and final spin states.  $H$ is then an effective spin Hamiltonian,
526: i.e. it can be expressed entirely in terms of the spin operators ${\bf
527: S}_A$ and ${\bf S}_B$, and $\tau$ is a measure of the pulse duration.
528: Note the definition of $\tau$ is arbitrary because it is the product
529: $\tau H$ which determines $U$.  Here, and in the remainder of this
530: paper, we work in units in which $\hbar = 1$.
531: 
532: If we assume exact axial symmetry throughout the pulse, the effective
533: spin Hamiltonian must be invariant under rotations about the $z$-axis
534: in spin space and must also leave the states $|T_+\rangle$ and
535: $|T_-\rangle$ degenerate.  The most general such spin Hamiltonian, up
536: to an irrelevant additive term proportional to the identity operator,
537: is
538: \begin{eqnarray}
539: \tau { H}
540: (\lambda;\alpha,\beta,\gamma) &=& \lambda\bigl({\bf S}_A \cdot {\bf S}_B + 
541: \frac{\alpha}{2}({S_A}_z- {S_B}_z)\nonumber\\
542: &&~~~~~+{
543: {\beta}} ( {S_A}_x {S_B}_y - {S_A}_y {S_B}_x)\nonumber\\ 
544: &&~~~~~+{\gamma} ( {S_A}_x {S_B}_x + {S_A}_y {S_B}_y)\bigr),
545: \label{hlabg}
546: \end{eqnarray}
547: and we denote the corresponding quantum gate as
548: %
549: \be
550: U(\lambda;\alpha,\beta,\gamma) = e^{-i\tau H
551: (\lambda;\alpha,\beta,\gamma)}.
552: \label{axialnt}
553: \ee
554: %
555: When $\alpha = 0$, this is precisely the gate (\ref{axial}) for
556: $\mbox{\boldmath{$\beta$}}\parallel \hat {\bf z}$.
557: 
558: %
559: 
560: The CNOT construction originally proposed in Ref.~\onlinecite{loss} is
561: based on the sequence of gates
562: %
563: \be
564: U_g = U(\pi/2; 0,0,0) e^{i\pi {S_A}_z} U(\pi/2; 0,0,0),
565: \label{cnot}
566: \ee
567: %
568: where $U(\pi/2; 0,0,0) = \exp-i[(\pi/2) {\bf S}_A \cdot {\bf S}_B]$ is
569: a square-root of swap gate.  The CNOT gate is then
570: %
571: \be
572: U_{CNOT} =
573: e^{i(\pi/2){S_A}_z} e^{i(\pi/2){S_B}_z} U_g.
574: \ee
575: %
576: Remarkably, it was shown in Ref.~\onlinecite{burkard01} that if
577: $\lambda = \pi/2$ and $\alpha = 0$ this construction is robust against
578: the $\beta$ and $\gamma$ corrections, i.e. the gate
579: %
580: \be
581: U_g = U(\pi/2; 0,\beta,\gamma) e^{i\pi {S_A}_z} U(\pi/2; 0,\beta,\gamma)
582: \ee
583: %
584: is independent of $\beta$ and $\gamma$.
585: 
586: For completeness, we briefly review the arguments of
587: Ref.~\onlinecite{burkard01}.  Due to axial symmetry, the action of the
588: gate $U(\lambda;\alpha,\beta,\gamma)$ on the states $|T_+\rangle$ and
589: $|T_-\rangle$ is trivial and independent of $\alpha, \beta$ and
590: $\gamma$,
591: %
592: \be
593: U(\lambda;\alpha,\beta,\gamma) |T_\pm\rangle =  e^{-i\lambda/4} |T_\pm\rangle.
594: \ee
595: %
596: We can then introduce a pseudospin description of the remaining space,
597: where $|S_1\rangle$ is pseudospin down and $|T_0\rangle$ is pseudospin
598: up. The action of the gate $U(\lambda;\alpha,\beta,\gamma)$ on this
599: pseudospin space is a simple rotation,
600: %
601: \be U(\lambda;\alpha,\beta,\gamma) \Rightarrow e^{i\lambda/4} 
602: e^{-i {{\bf b}}\cdot {\mbox{\boldmath$\tau$}}/2}, \ee
603: %
604: where ${\bf b} = \lambda (\alpha,\beta,\gamma+1)$ and the components
605: of ${\mbox{\boldmath{$\tau$}}} = (\tau_x,\tau_y,\tau_z)$ are pseudospin
606: Pauli matrices. At the same time, the action of the single qubit
607: rotation entering $U_g$ is
608: \be
609: e^{i\pi {S_A}_z} \Rightarrow i \tau_x.
610: \ee
611: %
612: 
613: 
614: Thus to show that the CNOT construction is independent of $\beta$ and
615: $\gamma$ if $\alpha = 0$ we need only show that the product
616: %
617: \be e^{ -i {\bf b}\cdot {\mbox{\boldmath$\tau$}}/2} \tau_x 
618: e^{-i {\bf b}\cdot \mbox{\boldmath$\tau$}/2} \ee
619: %
620: is independent of $\beta$ and $\gamma$ if $\alpha = 0$. This condition
621: has a simple geometric interpretation. It is the requirement that a
622: rotation about an axis parallel to ${\bf b}$, followed by a 180$^o$
623: rotation about the $x$-axis, and then a repeat of the initial rotation
624: must be equivalent to a simple 180$^o$ rotation about the $x$-axis.
625: This will trivially be the case if the vector ${\bf b} =
626: \lambda(\alpha,\beta,\gamma+1)$ lies in the $yz$ plane.  Thus, if
627: $\alpha =0$, this condition is satisfied and the CNOT construction is
628: exact.  Conversely, if $\alpha \ne 0$ the construction is spoiled.
629: 
630: 
631: \section{Time-Reversal Symmetry}
632: 
633: In this section we prove the following general result. Any
634: time-dependent Hamiltonian $H_P(t)$ which is time-reversal symmetric
635: for all $t$, and for which the time dependence is itself symmetric,
636: i.e. $H_P(t_0-t) = H_P(t_0+t)$ for all $t$, will generate a unitary
637: evolution operator $U = \exp -i \tau H$ where $H$ is a
638: time-independent effective Hamiltonian which is also time-reversal
639: symmetric.  We then show that this theorem implies that the parameter
640: $\alpha$, which spoils the CNOT construction described in Sec.~III, is
641: equal to zero for time-symmetric pulsing.
642: 
643: The time-reversal operation for any quantum system can be represented
644: by an antiunitary operator $\Theta$.\cite{gottfried} An orthonormal
645: basis $\{|M_i\rangle\}$ for the Hilbert space of this system is then
646: said to be a time-symmetric basis if
647: %
648: \be \Theta | M_i\rangle = |M_i\rangle 
649: \ee 
650: %
651: for all $i$.
652: 
653: For any Hamiltonian $H$ acting on a state $|M_i\rangle$ in this basis
654: we can write
655: %
656: \be H |M_i\rangle = \sum_j \langle M_j | H | M_i\rangle
657: |M_j \rangle. 
658: \label{hm1}
659: \ee 
660: %
661: Under time reversal $H$ is transformed into $\Theta H \Theta^{-1}$.
662: Using the invariance of the $\{|M_i\rangle\}$ basis and the antiunitarity
663: of $\Theta$ we can then also write
664: %
665: \be \Theta H \Theta^{-1} |M_i\rangle &=&
666: \Theta H |M_i\rangle \\ &=& \Theta \sum_j \langle M_j | H | M_i\rangle
667: |M_j \rangle \\ &=& \sum_j \langle M_j | H | M_i\rangle^* |M_j\rangle.
668: \label{hm2}
669: \ee 
670: %
671: Comparing (\ref{hm1}) and (\ref{hm2}) leads to the conclusion that if
672: $H$ is time-reversal symmetric, i.e. $H = \Theta H \Theta^{-1}$, then
673: the Hamiltonian matrix is purely real in the $\{|M_i\rangle\}$ basis,
674: while if $H$ is antisymmetric under $\Theta$, i.e. $H = - \Theta H
675: \Theta^{-1}$, then the Hamiltonian matrix is purely imaginary.
676: 
677: Since $H$ is real in the $\{|M_i\rangle\}$ basis if and only if $H$ is
678: time-reversal symmetric it follows that the unitary operator $U = \exp
679: - i\tau H$ is self-transpose, i.e. $U = U^{T}$, if and only if $H$ is
680: time-reversal symmetric.
681: 
682: Now consider a time-dependent pulse described by the Hamiltonian
683: $H_P(t)$.  We assume that $H_P(t)$ is time-reversal symmetric at all
684: times, i.e. $H_P(t) = \Theta H_P(t) \Theta^{-1}$ for all $t$.  The
685: corresponding unitary evolution operator $U$ which evolves the system
686: from time $t_I$ to $t_F$ can be written
687: \begin{eqnarray}
688: U = \lim_{N\rightarrow\infty}U(t_N) U(t_{N-1}) \cdots U(t_2) U(t_1)
689: \end{eqnarray}
690: where 
691: \begin{eqnarray}
692: U(t_i) = e^{- i \Delta t H_P(t_i)},
693: \end{eqnarray}
694: with $\Delta t = (t_F- t_I)/N$ and $t_1 \equiv t_I$ and $t_N \equiv
695: t_F$.
696: 
697: Since $H_P(t_i)$ is time-reversal symmetric the above arguments imply
698: $U^{T}(t_i) = U(t_i)$ when $U(t_i)$ is expressed in the
699: time-symmetric basis $\{|M_i\rangle\}$. Thus, in this basis, we have
700: \begin{eqnarray}
701: U^{T} &=& \lim_{N\rightarrow\infty}
702: \left( U(t_N) U(t_{N-1}) \cdots U(t_{2}) U(t_{1})
703: \right)^{T} \\ &=& 
704: \lim_{N\rightarrow\infty}
705: U^{T}(t_1) U^{T}(t_{2}) \cdots U^{T}(t_{N-1})
706: U^{T}(t_{N}) \\ &=& 
707: \lim_{N\rightarrow\infty}
708: U(t_1) U(t_{2}) \cdots U(t_{N-1}) U(t_{N}).
709: \label{transpose}
710: \end{eqnarray}
711: 
712: For a time-symmetric pulse $H_P(t_i) = H_P(t_{N+1-i})$ and so $U(t_i)
713: = U(t_{N+1-i})$.  This allows us to reverse the order of the operators
714: in (\ref{transpose}) which then implies
715: \begin{eqnarray}
716: U^{T} = U.
717: \end{eqnarray}
718: Thus if we write $U$ in terms of an effective Hamiltonian, 
719: %
720: \be U = e^{
721: -i \tau {H} },
722: \ee 
723: %
724: the matrix elements of $H$ must be real in the time-symmetric basis.
725: $H$ must therefore be time-reversal symmetric, i.e. $H = \Theta H
726: \Theta^{-1}$.
727: 
728: To apply this theorem to the present problem we take the time-reversal
729: operator for our two-electron system  to be
730: %
731: \be
732: \Theta = e^{i\pi {S_A}_y} e^{i\pi {S_B}_y}  K.
733: \ee
734: %
735: Here the antiunitary operator $K$ is defined so that when acting on a
736: given state it takes the complex conjugate of the amplitudes of that
737: state when expressed in the Hund-Mulliken basis defined in Sec.~II.
738: Note that this basis is constructed using the Fock-Darwin states, and
739: if a magnetic field is present these states will be necessarily
740: complex valued when expressed in the position basis.  As defined here,
741: the antiunitary operator $K$ only takes the complex conjugates of the
742: amplitudes in the Hund-Mulliken basis, {\it it does not take the
743: complex conjugate of the Fock-Darwin states themselves}. Thus, if a
744: magnetic field is present, $\Theta$ should be viewed as an {\it
745: effective} time-reversal symmetry operator.  This is a technical point
746: which does not affect any of our conclusions (provided the Zeeman
747: coupling can be ignored --- see below).  The key property that we will
748: need in what follows is that spin changes sign under time reversal,
749: and it is readily verified that for our definition of $\Theta$,
750: %
751: \be
752: \Theta {\bf S}_\mu \Theta^{-1} = - {\bf S}_\mu
753: \ee
754: %
755: for $\mu=A,B$ even in the presence of a magnetic field.
756: 
757: Under $\Theta$, the Hund-Mulliken basis states transform as follows,
758: %
759: \be
760: \Theta |S_i\rangle &=& \phantom{-}|S_i\rangle\ \ \ \ \ {\rm for\ }i=1,2,3,\\
761: \Theta |T_0\rangle &=& -|T_0\rangle, \\
762: \Theta |T_+\rangle &=& \phantom{-}|T_-\rangle, \\
763: \Theta |T_-\rangle &=& \phantom{-}|T_+\rangle.
764: \ee
765: %
766: The states $|S_i\rangle$ therefore form a time-symmetric basis for the
767: singlet states.  A time-symmetric basis for the triplet states is
768: given by
769: %
770: \be 
771: |\tilde T_0 \rangle &=& i|T_0\rangle, \\
772: |\tilde T_a \rangle &=& \frac{1}{\sqrt{2}} (|T_+\rangle + |T_-\rangle), \\
773: |\tilde T_b \rangle &=& \frac{i}{\sqrt{2}}(|T_-\rangle - |T_+\rangle),
774: \ee  
775: %
776: all of which are eigenstates of $\Theta$ with eigenvalue $+1$.
777: 
778: 
779: The matrix representation of $H_{HM}$ in the time-reversal invariant
780: $|\tilde T_0\rangle$, $| S_1\rangle$, $| S_2\rangle$ basis is
781: %
782: \be H_{HM} = \left(
783: \begin{array}{ccc}
784: V & 0 & -2P \\
785: 0 & 0 & -2t_H \\
786: -2P & -2t_H & U_H
787: \end{array}
788: \right),
789: \label{hmatrix2}
790: \ee 
791: %
792: which is real, reflecting the effective time-reversal symmetry of
793: $H_{HM}$.  Note that this would not be the case if $H_{HM}$ included
794: the Zeeman coupling of electron spins to an external magnetic field.
795: While for typical field strengths the Zeeman coupling is
796: small,\cite{burkard99_1} for some parameters it can be comparable to
797: the spin-orbit corrections considered here.  If this is the case our
798: conclusions following from effective time-reversal symmetry will no
799: longer be valid. Of course in zero magnetic field exact time-reversal
800: symmetry is guaranteed.
801: 
802: We now consider pulsing a time dependent $H_{HM}(t)$ adiabatically so
803: that, according to the arguments of Sec.~III, the resulting gate can
804: be parametrized by an effective spin Hamiltonian $H$.  Since at all
805: times $t$ the Hund-Mulliken Hamiltonian is time-reversal symmetric, if
806: the pulse itself is time symmetric, i.e. $H_{HM}(t) = H_{HM}(-t)$
807: where we take the center of the pulse to be at $t=0$, then the above
808: theorem implies that the effective spin Hamiltonian $H$ will also be
809: time-reversal symmetric. Thus $H = \Theta H \Theta^{-1}$, and since
810: $\Theta {\bf S}_\mu \Theta^{-1} = - {\bf S}_\mu$ this implies $H$ must
811: be quadratic in the spin operators, and so $\alpha = 0$.  The
812: resulting gate will therefore have the desired form (\ref{axial}).
813: 
814: 
815: For completeness we also consider here the case of time-antisymmetric
816: pulsing.  If $H_P(t) = - H_P(-t)$ then 
817: 
818: \end{multicols}
819: 
820: \begin{figure}[h]
821: \centerline{\psfig{figure=dpulse.eps,height=5.0in,angle=0}}\vskip .15in
822: \caption{Time dependence of matrix elements appearing in the
823: Hund-Mulliken description of a double quantum dot when the
824: displacement of the dots is varied according to (\ref{doft}) with $d_0
825: = 1$.  Results are for GaAs parameters in zero magnetic field with
826: $\hbar\omega_0 = 3$ meV and are plotted vs. the dimensionless quantity
827: $t/\tau$ for two values of the time-asymmetry parameter, $r = 0$
828: (solid line) and $r=0.1$ (dashed line).}
829: \label{dpulse}
830: \end{figure}
831: 
832: \begin{multicols}{2}
833: 
834: \begin{eqnarray}
835: U(t) = e^{- i \Delta t H_P(t) } = e^{i \Delta t H_P(-t)}  = U(-t)^{-1},
836: \end{eqnarray}
837: and the resulting quantum gate is
838: \begin{eqnarray}
839: U &=& \lim_{N\rightarrow\infty}U(t_1) U(t_2) \cdots U(t_{N/2})\nn\\
840: &&~~~~~~~~~\times U(t_{N/2})^{-1} \cdots U(t_2)^{-1} U(t_1)^{-1} = 1.
841: \end{eqnarray}
842: The net effect of any time-antisymmetric pulse is thus simply the
843: identity transformation.
844: 
845: 
846: \section{Model Calculations}
847: 
848: We have seen from symmetry arguments that time-symmetric pulsing of an
849: axially symmetric Hamiltonian, such as $H_{HM}$ when $f_D$ and $f_R$
850: are constant, which is itself time-reversal symmetric at all times,
851: will automatically produce a gate of the form (\ref{axial}), provided
852: the pulse is adiabatic so that the initial and final states of the
853: system are in the four-dimensional Hilbert space of two qubits.  It is
854: natural to then ask what the effect of the inevitable deviations from
855: time-symmetric pulsing will be on the resulting gate.  To investigate
856: this we have performed some simple numerical simulations of coupled
857: quantum dots.
858: 
859: 
860: 
861: In our calculations, we imagine pulsing the dots by varying the
862: dimensionless distance $d$ between them according to
863: %
864: \be
865: d(t) = d_0 + \left(\frac{t}{\tau + r t}\right)^2.
866: \label{doft}
867: \ee
868: %
869: Here $d_0$ is the distance at the point of closest approach, $\tau$ is
870: a measure of the pulse duration, and $r$ is a dimensionless measure of
871: the time asymmetry of the pulse.  This form describes the generic
872: behavior of any pulse for times near the pulse maximum ($t = 0$).
873: Note that for large $|t|$, and for $r \ne 0$, the distance $d(t)$ will
874: saturate, and has a singularity for negative $t$.  We have taken $r$
875: to be small enough so that the dots decouple long before this leads to
876: any difficulty.
877: 
878: \end{multicols}
879: 
880: \begin{figure}[t]
881: \centerline{\psfig{figure=heff.eps,height=5.0in,angle=0}}\vskip.15in
882: \caption{Parameters appearing in the effective spin Hamiltonian
883: derived from pulses depicted in Fig.~\ref{dpulse}.  The parameters
884: $\alpha$, $\beta$ and $\gamma$ are shown as functions of $s$ for the
885: case $r=0$ (time-symmetric pulses).  For $\alpha$ the quantity
886: $\alpha/s$ is plotted vs. $r$.  We have verified that the ratio
887: $\alpha/s$ is essentially independent of $s$ for all values we have
888: considered ($|s| \le 0.1$).}
889: \label{heff}
890: \end{figure}
891: 
892: \begin{multicols}{2}
893: 
894: For our calculations, we work in zero magnetic field and take
895: $\hbar\omega_0 = 3$ meV and $d_0 = 1$, corresponding to $a \simeq 20$
896: nm at closest approach.  The resulting time dependences of the
897: parameters in $H_{HM}$ are shown in Fig.~\ref{dpulse}.  Note that the
898: spin-orbit matrix element plotted in this figure is $l_{SO}$, while
899: the spin-orbit matrix element appearing in $H_{HM}$ is ${\bf P} = s
900: l_{SO} \hat z $ where $s$ is the dimensionless measure of spin-orbit
901: coupling introduced in Sec.~II.
902: 
903: For a given pulse $H_{HM}(t)$ we integrate the time-dependent
904: Schr\"odinger equation to obtain the evolution operator $U$ for the
905: full pulse.  If the pulse is adiabatic then the matrix elements of $U$
906: which couple the singly occupied states $|S_1\rangle$ and
907: $|T_0\rangle$ to the doubly occupied state $|S_2\rangle$ can be made
908: negligibly small.\cite{schliemann} The quantum gate is then obtained
909: by simply truncating $U$ to the $4\times 4$ matrix acting on the
910: two-qubit Hilbert space.  By taking the $\log$ of this matrix we
911: obtain $\tau H = i \log U$ and thus the parameters $\lambda, \alpha,
912: \beta, \gamma$.  Note that when calculating $\log U$, there are branch
913: cuts associated with each eigenvalue of $U$, and as a consequence
914: $\tau H$ is not uniquely determined.  We resolve this ambiguity by
915: requiring that as the pulse height is reduced to zero and $U$ goes
916: continuously to the identity that $\tau H \rightarrow 0$ without
917: crossing any branch cuts.
918: 
919: We fix the pulse width $\tau$ by requiring that if we turn off
920: spin-orbit coupling ($s=0$) we obtain a $\lambda = \pi/2$ pulse,
921: i.e. a square-root of swap.  For the parameters used here we find this
922: corresponds to taking $\tau = 23.9/\omega_0 \simeq 5$ ps.  We have
923: checked that these pulses are well into the adiabatic regime.  The
924: magnitudes of the matrix elements coupling singly occupied states to
925: the doubly occupied state $|S_2\rangle$ are on the order of $|\langle
926: S_1 | U | S_2\rangle| \sim 10^{-6}$ and $|\langle T_0 | U |
927: S_2\rangle| \sim s 10^{-6}$.
928: 
929: \vskip 0.1in
930: 
931: \begin{table}[h]
932: \vskip 0.3in
933: \caption{Symmetry properties of the pulse parameters $r$ and $s$, and
934: gate parameters $\lambda$, $\alpha$, $\beta$ and $\gamma$ under parity
935: $P$ and time reversal $T$.}
936: \begin{tabular}{ccc|ccc|ccccc}
937: &  & & $r$  & $s$ & & $\lambda$ & $\alpha$ & $\beta$ & $\gamma$ \\  
938: \tableline
939: & P &  & $+$  & $-$ & & $+$  & $-$  &  $-$  &  $+$ \\ 
940: & T &  & $-$  & $+$ & & $+$  & $-$  &  $+$  &  $+$
941: \end{tabular}
942: \end{table}
943: 
944: 
945: Once $\tau$ is fixed, there are two parameters characterizing each
946: pulse, $s$ and $r$, and four parameters characterizing the resulting
947: gate, $\lambda$, $\alpha$, $\beta$, and $\gamma$.  The transformation
948: properties of these parameters under parity ($P$) and time reversal
949: ($T$) are summarized in Table I.  These properties follow from the
950: fact that (i) under time reversal ${\bf S}_\mu \rightarrow - {\bf
951: S}_\mu$ and $r \rightarrow -r$, while ${\bf P} = s l_{SO}{\hat {\bf
952: z}}$ is invariant, and (ii) under parity ${\bf S}_A \leftrightarrow
953: {\bf S}_B$ and ${\bf P} \rightarrow -{\bf P}$, while $r$ is invariant.
954: Note that, as defined in Sec.~II, the parameter $s$ is positive. Here
955: we allow $s$ to change sign when the direction of the vector ${\bf P}$
956: is reversed, thus under parity $s \rightarrow -s$.
957: 
958: 
959: 
960: These symmetry properties imply that if $s$ and $r$ are small, the
961: parameters of the effective Hamiltonian will be given approximately by
962: %
963: \be
964: \alpha &\simeq& C_\alpha r s, \\
965: \beta &\simeq& C_\beta s, \\
966: \gamma &\simeq& C_\gamma s^2, \\
967: \lambda &\simeq& \lambda_0 + C_\lambda s^2,
968: \ee
969: %
970: where the coefficients should be of order 1. For the pulses we
971: consider here $\lambda_0 = \pi/2$.
972: 
973: 
974: The results of our calculations are shown in Fig.~\ref{heff}.  Each
975: point corresponds to a separate numerical run.  The plots for
976: $\lambda$, $\beta$ and $\gamma$ show their dependence on $s$ when
977: $r=0$.  The dependence of the parameter $\alpha$ on pulse asymmetry is
978: shown by plotting $\alpha/s$ vs. $r$.  For the $s$ values we have
979: studied, up to $|s| = 0.1$, the numerical results for $\alpha/s$ are
980: essentially independent of $s$ for a given $r$.  These results are
981: clearly consistent with the above symmetry analysis.
982: 
983: 
984: Now consider carrying out a CNOT gate using the scheme reviewed in
985: Sec.~III.  For this construction to work it is necessary that $\lambda
986: = \pi/2$. In our calculations we have fixed $\tau$ so that $\lambda =
987: \pi/2$ for $s=r=0$.  Thus, when spin-orbit coupling is included
988: %
989: \be
990: \lambda \simeq \pi/2 + C_\lambda s^2.
991: \ee
992: %
993: In order to keep $\lambda = \pi/2$ it will therefore be necessary to
994: adjust the pulse width $\tau$ slightly to correct for spin-orbit
995: effects.  
996: 
997: The central result of this paper is summarized by the equation
998: %
999: \be
1000: \alpha \simeq C_\alpha r s.
1001: \label{alpha}
1002: \ee
1003: % 
1004: As shown in Sec.~III, any nonzero $\alpha$ will lead to corrections to
1005: the CNOT construction. For time-symmetric pulses $r=0$ and these
1006: corrections will vanish.  Equation (\ref{alpha}) can then be used to
1007: estimate the errors due to any time asymmetry of the pulse, and to put
1008: design restrictions on the allowed tolerance for such asymmetry.
1009: 
1010: It is important to note that while the results presented here are for
1011: a specific model, all of the key arguments are based on symmetry and
1012: so are quite general.  Given any time-reversal invariant two-qubit
1013: system with axial symmetry, if pulsed adiabatically in a
1014: time-symmetric way the resulting gate will have the form
1015: (\ref{axial}).
1016: 
1017: 
1018: If the pulse is not axially symmetric, e.g. if the ratio $f_D/f_R$
1019: varies during the pulse, then time-symmetric pulsing will still
1020: restrict the resulting gate to be invariant under time reversal.
1021: Thus, up to an irrelevant overall phase, this gate will necessarily
1022: have the form
1023: %
1024: \be
1025: U = \exp -i \lambda ({\bf S}_A \cdot {\bf S}_B
1026: +\mbox{\boldmath{$\beta$}} \cdot {\bf S}_A \times {\bf S}_B
1027: + {\bf S}_A \cdot {\bf I}\!{\mbox{\boldmath{$\Gamma$}}} \cdot {\bf S}_B).
1028: \ee
1029: %
1030: Here ${\bf I}\!{\mbox{\boldmath{$\Gamma$}}}$ is a symmetric tensor
1031: which will, in general, deviate from the axial form of the $\gamma$
1032: term in (\ref{axial}) leading to corrections to the CNOT construction.
1033: However, because ${\bf I}\!{\mbox{\boldmath{$\Gamma$}}}$ is even under
1034: parity it will still be second order in spin-orbit
1035: coupling,\cite{bonesteel01} and thus the deviations from (\ref{axial})
1036: will also be second order.  We conclude that even in the absence of
1037: axial symmetry, the corrections to the CNOT construction will be
1038: second order in spin-orbit coupling, rather than first order.
1039: 
1040: \section{Conclusions}
1041: 
1042: In this paper we have studied spin-orbit corrections to exchange-based
1043: quantum gates, emphasizing symmetry arguments.  In particular, we have
1044: shown that adiabatic time-symmetric pulsing of any Hamiltonian which
1045: (i) describes two well defined spin-1/2 qubits at the beginning and
1046: end of the pulse, (ii) is time-reversal symmetric at all times during
1047: the pulse, and (iii) is axially symmetric in spin space with a fixed
1048: symmetry axis, will automatically produce a gate of the form
1049: (\ref{axial}).  Together with single qubit rotations, for $\lambda =
1050: \pi/2$ this gate can then be used in a simple CNOT construction. This
1051: result is quite general.
1052: 
1053: As a specific example we have studied a GaAs double quantum dot system
1054: within the Hund-Mulliken approximation.  In this approximation
1055: spin-orbit coupling enters as a small spin precession when an electron
1056: tunnels between dots.  If the direction of this precession axis is
1057: constant throughout the pulse the resulting gate will be axially
1058: symmetric and have the form (\ref{axialnt}).  The deviation of this
1059: gate from the desired gate (\ref{axial}) is then characterized by a
1060: single dimensionless parameter $\alpha$ which spoils the CNOT
1061: construction.  Using symmetry arguments, as well as numerical
1062: calculations, we have shown that $\alpha \simeq C_\alpha s r$ where $s$ and
1063: $r$ are, respectively, dimensionless measures of spin-orbit coupling
1064: and time asymmetry of the pulse.  Thus time-symmetric pulsing ($r=0$)
1065: ensures the anisotropic corrections will have the desired form.
1066: 
1067: In any system without spatial inversion symmetry, spin-orbit coupling
1068: will inevitably lead to anisotropic corrections to the exchange
1069: interaction between spins.  According to current
1070: estimates,\cite{aharonov} fault-tolerant quantum computation will
1071: require realizing quantum gates with an accuracy of one part in
1072: $10^4$.  Thus, even if spin-orbit coupling is weak, the design of any
1073: future quantum computer which uses the exchange interaction will have
1074: to take these anisotropic corrections into account.  We believe the
1075: symmetry based analysis presented in this paper provides a useful
1076: framework for studying these effects.
1077: 
1078: 
1079: \acknowledgements DS and NEB acknowledge support from the National
1080: Science Foundation through NIRT Grant No.\ DMR-0103034.  DPDV is
1081: supported in part by the National Security Agency and the Advanced
1082: Research and Development Activity through Army Research Office
1083: contract number DAAD19-01-C-0056.  He thanks the Institute for Quantum
1084: Information at Cal Tech (supported by the National Science Foundation
1085: under Grant. No. EIA-0086038) for its hospitality during the initial
1086: stages of this work.  DL thanks Swiss NSF, NCCR Nanoscience, DARPA and
1087: ARO.
1088: 
1089: 
1090: \begin{references}
1091: 
1092: \bibitem{loss} D. Loss and D.P. DiVincenzo, Phys. Rev. A {\bf 57},
1093: 120 (1998).
1094: 
1095: \bibitem{burkard99_1} G. Burkard, D. Loss, and D.P. DiVincenzo,
1096: Phys. Rev. B {\bf 59}, 2070 (1999).
1097: 
1098: \bibitem{burkard99_2} G. Burkard, D. Loss, D.P. DiVincenzo, and
1099: J.A. Smolin, Phys. Rev. B {\bf 60}, 11404 (1999).
1100: 
1101: \bibitem{hu} X. Hu and S. Das Sarma, Phys. Rev. A {\bf 61}, 062301
1102: (2000).
1103: 
1104: \bibitem{schliemann} J. Schliemann, D. Loss and A.H. MacDonald,
1105: Phys. Rev. B {\bf 63}, 085311 (2001).
1106: 
1107: \bibitem{bacon} D. Bacon, J. Kempe, D.A. Lidar, and K.B. Whaley,
1108: Phys. Rev. Lett. {\bf 85}, 1758 (2000).
1109: 
1110: \bibitem{divincenzo} D.P. DiVincenzo, D. Bacon, J. Kempe, G. Burkard
1111: and K.B. Whaley, Nature {\bf 408}, 339 (2000).
1112: 
1113: \bibitem{burkard01} G. Burkard and D. Loss,  Phys. Rev. Lett. {\bf 88}, 047903
1114: (2002).
1115: 
1116: \bibitem{wu} L.-A. Wu and D. Lidar, Phys. Rev. A {\bf 66}, 062314
1117: (2002).
1118: 
1119: \bibitem{kavokin} K.V. Kavokin, Phys. Rev. B {\bf 64} 075305 (2001);
1120: cond-mat/0212347.
1121: 
1122: \bibitem{gorkov} L.P. Gor'kov and P.L. Krotkov, Phys. Rev. B {\bf 67},
1123: 033203 (2003).
1124: 
1125: \bibitem{bonesteel01} N.E. Bonesteel, D. Stepanenko, and
1126: D.P. DiVincenzo, Phys. Rev. Lett. {\bf 87}, 207901 (2001).
1127: 
1128: \bibitem{dresselhaus} G. Dresselhaus, Phys. Rev. {\bf 100}, 580
1129: (1955).
1130: 
1131: \bibitem{dyakonov} M. I. Dyakonov and V. Y. Kachorovskii, Sov. Phys.
1132: Semicond. {\bf 20}, 110 (1986).
1133: 
1134: \bibitem{rashba} E. L. Rashba, Fiz. Tv. Tela (Leningrad) {\bf 2}, 
1135: 1224 (1960) [Sov. Phys. Solid State {\bf 2}, 1109 (1960)]; Y.A. Bychkov
1136: and E.I. Rashba, J. Phys. C {\bf 17}, 6039 (1984).
1137: 
1138: \bibitem{schliemann03} J. Schliemann, J.C. Egues, and D. Loss,
1139: Phys. Rev. Lett. {\bf 90}, 146801 (2003).
1140: 
1141: \bibitem{gottfried} For an excellent discussion of time-reversal
1142: symmetry see, ``Quantum Mechanics'', K. Gottfried, Addison-Wesley,
1143: 1989, pp. 314-322.
1144: 
1145: \bibitem{aharonov} D. Aharonov and M. Ben-Or, quant-ph/9906129.
1146: 
1147: \end{references}
1148: 
1149: \end{multicols}
1150: 
1151: \end{document}
1152: 
1153: 
1154:  
1155: 
1156: 
1157: