1: \documentclass[pra,onecolumn, superscriptaddress, preprint]{revtex4}
2: \usepackage{amstext}
3: \usepackage{amsmath}
4: \usepackage{amssymb}
5:
6: \newcommand{\ud}{\mathrm{d}}
7: \newcommand{\bra}[1]{\langle #1|}
8: \newcommand{\ket}[1]{| #1\rangle}
9: \newcommand{\vect}[1]{\boldsymbol{#1}}
10: \newcommand{\eq}[1]{Eq.~\eqref{#1}}
11: \newcommand{\fig}[1]{Fig.~\ref{#1}}
12: \newcommand{\stn}[1]{Sec.~\ref{#1}}
13: \newcommand{\be}{\begin{equation}}
14: \newcommand{\ee}{\end{equation}}
15: \newcommand{\ba}{\begin{align}}
16: \newcommand{\ea}{\end{align}}
17: \newcommand{\ti}[1]{\text{#1}}
18: \newcommand{\mc}[1]{\mathcal{#1}}
19: \newcommand{\w}{\omega}
20: \newcommand{\mean}[1]{\langle #1 \rangle}
21: \newcommand{\me}[3]{\langle #1 | #2 | #3 \rangle}
22:
23: \begin{document}
24:
25: \title{Minimum requirements for laser-induced
26: symmetry breaking in quantum and classical mechanics}
27: \author{Ignacio Franco and Paul Brumer\\
28: Chemical Physics Theory Group, Department of Chemistry, and \\
29: Center for Quantum Information and Quantum Control,\\ University
30: of Toronto, Toronto, Ontario M5S 3H6, Canada.}
31:
32:
33: \date{\today}
34: \vspace{0.5in}
35: \begin{abstract}
36: Necessary conditions for generating phase controllable asymmetry in spatially
37: symmetric systems using lasers are identified and are shown to be identical in
38: quantum and classical mechanics. First, by studying the exact dynamics of
39: harmonic systems in the presence of an arbitrary radiation field, it is demonstrated that
40: anharmonicities in the system's potential are a necessary requirement for phase controllability.
41: Then, by analyzing the space-time symmetries of the laser-driven
42: Liouville dynamics for classical and quantum systems, a common set of
43: temporal symmetries for the driving field that need to be violated to induce transport are identified. The conditions apply to continuous wave lasers and to symmetry breaking effects that do not rely on the control of the absolute phase of the field.
44: Known examples of laser fields that can induce transport in symmetric systems are
45: seen to be particular cases of these symmetry constraints.
46: \end{abstract}
47:
48: \maketitle
49:
50: \section{Introduction}
51:
52: Recent years have witnessed the birth and rapid development of
53: the coherent control field ~\cite{paul,ricebook,bergmann,ricerev,dantusrev,nuernbergerrev}, in
54: which the coherence properties of applied laser fields are
55: employed to steer a given quantum dynamical process in a desired
56: direction. Of the different control schemes that have so far
57: been developed, there is a general class that has the ability to
58: induce phase controllable transport in spatially symmetric
59: systems without introducing a bias voltage in the potential, a
60: phenomenon that is referred to as laser-induced symmetry
61: breaking.
62:
63: This symmetry breaking effect is typically achieved by driving
64: the system with AC fields with frequency components $n\w$ and
65: $m\w$, where $n$ and $m$ are integers of different
66: parity~\cite{paul}. The nonlinear response of the system to such
67: fields results in net dipoles or currents whose magnitude and
68: sign can be manipulated by varying the relative phase between the
69: two frequency components of the radiation~\cite{francoprl}. For
70: the popular case of $n=1$ and $m=2$ the rectification effect
71: first appears in the third order response of the system to the
72: incident radiation. At this order the system mixes the
73: frequencies and harmonics of the field, generating a
74: phase-controllable zero-harmonic (DC) component in the response.
75:
76: Laser-induced symmetry breaking has been demonstrated in a wide
77: variety of systems ranging from atoms to solid state samples.
78: Experimentally it has been implemented for generating anisotropy
79: in atomic photoionization~\cite{yin_1vs2}, symmetry breaking effects in
80: molecular photodissociation~\cite{sheehy_1vs2} (see also Ref.~\onlinecite{charron_1vs2}), photocurrents in
81: quantum wells~\cite{dupont_1vs2} and intrinsic
82: semiconductors~\cite{hache_1vs2}, as well as directed diffusion
83: in symmetric optical lattices~\cite{schiavoni}. Theoretically, it
84: has been studied for generating transport in
85: doped~\cite{paul_1vs2} and bulk semiconductors through
86: interband~\cite{atanasov_1vs2} and intraband~\cite{pronin_1vs2}
87: excitations, in graphene and carbon nanotubes~\cite{mele_1vs2},
88: and molecular wires~\cite{prlwire, lehmann_1vs2}, among others.
89: The scenario is of interest since, with current laser
90: technology, it can be employed to generate transport on a
91: femtosecond timescale.
92:
93: An interesting feature of this laser control scenario is that it
94: has both a quantum~\cite{paul_1vs2, hanggi_harmonic_mixing,
95: dupont_1vs2, hache_1vs2} and a classical~\cite{flach, denisov}
96: manifestation. Further, the two versions of the effect correspond
97: to the same physical phenomenon~\cite{francoprl}, arising from
98: the nonlinear response of material systems to symmetry breaking
99: radiation fields. In this contribution, we isolate minimum
100: conditions on the driving field and on the system that is being
101: driven that are necessary for the symmetry breaking effect
102: to occur in quantum and classical mechanics. As shown, the
103: minimum requirements in both cases are identical and, further,
104: the effect can be accounted for in both mechanics through a
105: single symmetry analysis of the equations of motion.
106:
107: Specifically, in \stn{sec:harmonic} we demonstrate that laser
108: rectification can only occur in systems with anharmonic
109: potentials. Subsequently, in \stn{sec:symmetry} temporal
110: symmetries of the field that need to be violated to induce
111: transport are isolated. This is done by studying the space-time
112: symmetries of the Liouville equations of motion for
113: laser-driven quantum and classical systems, and by isolating
114: symmetries of the field that rule out any nonzero average
115: currents or dipoles at asymptotic times. In doing so we
116: considerably extend a previous analysis~\cite{flach} that
117: identified conditions on the field necessary for
118: laser-rectification in classical ergodic systems. It applies to
119: both quantum and classical systems and makes no assumption of
120: ergodicity.
121:
122:
123: \section{Conditions on the system}
124: \label{sec:harmonic}
125:
126: Consider first the exact solution for the dynamics of a harmonic
127: oscillator in the presence of an arbitrary space-homogeneous
128: radiation field $E(t)$. The Hamiltonian of the system reads
129: %----------------------------------------------------------------%
130: \begin{equation}
131: H(x,p) = \frac{p^2}{2m} + \frac{1}{2} m \omega_0^2 x^2 - q E(t) x,
132: \end{equation}
133: %----------------------------------------------------------------%
134: where $x$ and $p$ are the position and momentum of the particle
135: of mass $m$ and charge $q$, and $\w_0$ is the natural frequency
136: of the oscillator. Symmetry breaking here would correspond to
137: the production of a net dipole moment. In the quantum case, the
138: dynamics of the position $\hat{x}_\ti{H}(t)$ and momentum
139: operators $\hat{p}_\ti{H}(t)$ in Heisenberg picture is dictated
140: by the Heisenberg equations of motion
141: %----------------------------------------------------------------%
142: \begin{subequations}
143: \label{eq:qset}
144: \begin{align}
145: \frac{\ud \hat{x}_\ti{H}(t)}{\ud t} = \frac{1}{i \hbar}[\hat{x}_\ti{H}(t), \hat{H}_\ti{H}(t)] & = \frac{1}{m} \hat{p}_\ti{H}(t), \\
146: \frac{\ud \hat{p}_\ti{H}(t)}{\ud t} = \frac{1}{i \hbar}[\hat{p}_\ti{H}(t), \hat{H}_\ti{H}(t)] & = -m \omega_0^2 \hat{x}_\ti{H}(t) + q E(t),
147: \end{align}
148: \end{subequations}
149: %----------------------------------------------------------------%
150: where $\hat{H}_\ti{H}$ is the Hamiltonian operator in Heisenberg
151: picture and $[\hat{f}, \hat{H}_\ti{H}]= \hat{f}\hat{H}_\ti{H} -
152: \hat{H}_\ti{H} \hat{f}$ for any operator $\hat{f}$. In the
153: classical case, the position $x(t)$ and momentum $p(t)$
154: variables obey Hamilton's equations
155: %----------------------------------------------------------------%
156: \begin{subequations}
157: \label{eq:cset}
158: \begin{align}
159: \frac{\ud x(t)}{\ud t} = \left\{x(t), H(t) \right\} & = \frac{1}{m} p(t), \\
160: \frac{\ud p(t)}{\ud t} = \left\{p(t), H(t)\right\} & = -m \omega_0^2 x(t) + q E(t),
161: \end{align}
162: \end{subequations}
163: %----------------------------------------------------------------%
164: where $\{f, H \}= \frac{\partial f}{\partial x}\frac{\partial
165: H}{\partial p} - \frac{\partial f}{\partial p}\frac{\partial
166: H}{\partial q}$ is the Poisson bracket. The difference between
167: Eqs.~\eqref{eq:qset} and~\eqref{eq:cset} is that the former is a
168: differential equation for operators, with operator initial
169: conditions $\hat{x}_\ti{H}(0)=\hat{x}$ and
170: $\hat{p}_\ti{H}(0)=\hat{p}$, while the latter is an equation for
171: functions.
172:
173: These two sets of equations can be solved exactly using Laplace
174: transforms. In fact, for a general external field of the form
175: %----------------------------------------------------------------%
176: \begin{equation}
177: \label{eq:field2}
178: E(t)= \int_{-\infty}^\infty \, \ud \omega \,\epsilon({\omega}) e^{i\omega t},
179: \end{equation}
180: the usual procedure~\cite{arfken} yields:
181: %---------------------------------------------------------------------------------------------
182: \begin{multline}
183: \label{eq:quantum_pos}
184: \hat{x}_\ti{H}(t) =
185: \hat{x}_\ti{H}(0) \cos(\omega_0 t) +
186: \frac{\hat{p}_\ti{H}(0)}{m\omega_0} \sin(\omega_0 t) \\
187: +
188: \int_{-\infty}^\infty \ud\omega\, \frac{q \epsilon(\omega)}{m\omega_0}
189: \frac{i\omega \sin({{\omega }_0}t) +
190: {{\omega }_0}\cos({{\omega }_0}t) - {{\omega}_0}
191: e^{i \omega t}}
192: {{\omega }^2 - {{{\omega }_0}}^2};
193: \end{multline}
194: \begin{multline}
195: \label{eq:classical_pos}
196: x(t) = x(0) \cos(\omega_0 t) + \frac{p(0)}{m\omega_0} \sin(\omega_0 t) \\
197: + \int_{-\infty}^\infty \ud\omega\, \frac{q \epsilon(\omega)}{m\omega_0}
198: \frac{i\omega \sin({{\omega }_0}t) +
199: {{\omega }_0}\cos({{\omega }_0}t) - {{\omega}_0}
200: e^{i \omega t}}
201: {{\omega }^2 - {{{\omega }_0}}^2}.
202: \end{multline}
203: %----------------------------------------------------------------%
204: The first two terms in Eqs.~\eqref{eq:quantum_pos}
205: and~\eqref{eq:classical_pos} describe the field-free evolution of
206: the oscillator, while the third one characterizes the influence
207: of $E(t)$ on the dynamics.
208:
209: Note that a driven harmonic system can only oscillate at its
210: natural frequency $\omega_0$ and at the frequency of the field
211: $\omega$. That is, there are no frequency components of the
212: dipole that oscillate at multiples or combinations of the
213: frequencies of the field. Hence, if $E(t)$ is unbiased
214: ($\epsilon(0) =0$) no net dipole can be induced. Thus, we
215: conclude that a necessary requirement for symmetry breaking in
216: quantum and classical mechanics is that the potential of the
217: system is anharmonic. As seen below, the anharmonicities permit
218: the nonlinear response of the system to the incident radiation
219: that mixes the frequencies and harmonics of the field and, for a
220: certain class of radiation sources isolated below, can lead to
221: the generation of a phase-controllable zero-harmonic (dc)
222: component in the response.
223:
224: \section{Conditions on the field}
225: \label{sec:symmetry}
226:
227:
228: We now isolate those temporal symmetries of the field that need
229: to be violated to induce transport in both quantum and classical
230: mechanics. To do so we consider a symmetric one-dimensional
231: system composed of $N$ charged particles coupled to an
232: external field $E(t)$ in the dipole approximation. This is
233: done without loss of generality since the polarization of the
234: field effectively defines an axis along which symmetry breaking
235: can arise. The system's Hamiltonian is then:
236: %----------------------------------------------------------------%
237: \begin{equation}
238: \label{eq:hamgen}
239: H = \sum_{j=1}^N \frac{p_j^2}{2m_j} + V(\vect{x}) -\sum_{j=1}^N q_j x_j E(t+ \alpha\tfrac{T}{2\pi}),
240: \end{equation}
241: %----------------------------------------------------------------%
242: where $x_j$, $p_j$, $m_j$ and $q_j$ denote the coordinate,
243: momenta, mass and charge of the $j$-th particle and $\vect{x}
244: \equiv (x_1, x_2, \cdots, x_N)$. The systems of interest have a
245: potential $V(\vect{x})$ that is invariant under coordinate
246: inversion $V(-\vect{x}) = V(\vect{x})$, and the driving field
247: $E(t+ \alpha\frac{T}{2\pi})$ is an arbitrary time-periodic
248: zero-mean function, with period $T$ and global phase $\alpha$.
249:
250: In order to keep a close analogy between the quantum and
251: classical case we frame this analysis in phase space and adopt
252: the Wigner representation of quantum mechanics~\cite{ tatarskii,
253: wigner}. In it, the state of the quantum system is described by
254: the Wigner distribution function
255: $\rho_{\ti{W}}(\vect{x},\vect{p},t)$, which constitutes a map of
256: the system density matrix $\hat{\rho}$ in the phase space of
257: position $\vect{x}$ and momentum $\vect{p}$ variables.
258: For $N$-particle one-dimensional systems it is defined
259: by~\cite{wigner}
260: %----------------------------------------------------------------------
261: \be
262: \label{eq:wigner}
263: \rho_{\ti{W}}(\vect{x},\vect{p},t) =
264: \left(\frac{1}{2\pi\hbar}\right)^N
265: \int_{-\infty}^{\infty}\cdots\int_{-\infty}^{\infty}
266: \ud u_1 \cdots \ud u_N \,
267: e^{ \frac{i}{\hbar} \vect{p}\cdot \vect{u} }
268: \bra{\vect{x} -\vect{u}/2} \hat{\rho}(t)\ket{\vect{x} + \vect{u}/2},
269: \ee
270: %----------------------------------------------------------------------
271: where $\ket{\vect{x}}\equiv \ket{x_1}\ket{x_2}\cdots \ket{x_N}$ and $\vect{p} \cdot \vect{u} = \sum_{i=1}^N p_i u_i$. In this way the quantum or classical Liouville evolution can be expressed as
272: \begin{subequations}
273: \label{eq:dynamics}
274: \be
275: \label{eq:eqdynamics}
276: \mathcal{D}_{\beta} \rho_\beta(\vect{x}, \vect{p}, t)=0, \ee
277: where the label $\beta$ indicates either classical
278: ($\beta=\ti{c}$) or quantum ($\beta=\ti{W}$), with
279: $\rho_{\ti{c}}(\vect{x}, \vect{p},t)$ denoting the classical
280: phase space density. For the Hamiltonian in \eq{eq:hamgen}, the
281: operator $\mathcal{D}_\beta$ determining the dynamics is given
282: by~\cite{wigner, serimaa}
283: %--------------------------------------------------------------------
284: \begin{eqnarray}
285: \label{eq:classicaldynamics}
286: \mathcal{D}_{\text{c}} & = &
287: \frac{\partial}{\partial t} -
288: \sum_{j=1}^{N} \left[-\frac{p_j}{m_j} \frac{\partial}{\partial x_j}
289: + \left( \frac{\partial V(\vect{x})}{\partial x_j} -q_j E(t + \alpha\tfrac{T}{2\pi})\right)
290: \frac{\partial}{\partial p_j}\right], \\
291: \label{eq:wignerdynamics}
292: \mathcal{D}_{\ti{W}} & = &
293: \mathcal{D}_{\ti{c}} - \mspace{-25.0mu}
294: \sum_{\substack{\lambda_1,\ldots,\lambda_N \\
295: \lambda_1+\cdots+\lambda_N=3,5,\ldots}} \mspace{-25.0mu}
296: \frac{\left(i \hbar/2\right)^{\lambda_1+\cdots+\lambda_N-1}}
297: {\lambda_1! \cdots \lambda_N!}
298: \frac{\partial^{\lambda_1+\cdots+\lambda_N }V(\vect{x})}
299: {\partial x_1^{\lambda_1}\cdots{\partial x_N^{\lambda_N}}}
300: \frac{\partial^{\lambda_1+\cdots+\lambda_N}}
301: {\partial p_1^{\lambda_1}\cdots{\partial p_N^{\lambda_N}}},
302: \end{eqnarray}
303: \end{subequations}
304: where the last summation in $\mc{D}_\ti{W}$ runs over all
305: positive integer values of $\lambda_1,\ldots,\lambda_N$ for
306: which the sum $\lambda_1+\lambda_2+\cdots+\lambda_N$ is odd and
307: greater than one. In phase space the formal structure of the
308: quantum and classical evolution is remarkably
309: similar~\cite{wilkie1, wilkie2}. In the limit $\hbar\to 0$ the
310: second term in \eq{eq:wignerdynamics} vanishes and the quantum
311: equation of motion reduces to the classical evolution
312: ($\mathcal{D}_\text{w}\to \mathcal{D}_\text{c}$). Note that
313: Eqs.~\eqref{eq:wigner} and~\eqref{eq:dynamics} are fully
314: consistent with the Hamiltonian in \eq{eq:hamgen}. However, when
315: the radiation-matter interaction term in the Hamiltonian goes
316: beyond the dipole approximation both of them need to be modified
317: in order to ensure gauge invariance~\cite{serimaa}.
318:
319:
320: In the absence of an external field the equations of motion
321: [\eq{eq:dynamics}] are invariant under reflection ($\vect{x}\to
322: -\vect{x}$, $\vect{p}\to -\vect{p}$). Hence, if the system is
323: initially prepared with a given phase-space symmetry this initial
324: symmetry is preserved at all times during the subsequent
325: dynamics. Symmetry breaking is achieved by coupling the
326: system to $E(t)$. However, if $E(t)$ has a zero temporal mean (AC
327: field) then not every $E(t)$ will generate transport. As shown
328: below, by properly lowering the temporal symmetry of $E(t)$ it is
329: possible to induce rectification in the response. Furthermore,
330: the resulting symmetry constraints on $E(t)$ are identical for
331: the classical and quantum case. As will become evident, this is
332: a consequence of the important fact that the quantum corrections
333: in $\mc{D}_\ti{W}$ have the same symmetry properties as
334: $\frac{\partial V(\vect{x})}{\partial x_j}
335: \frac{\partial}{\partial p_j}$ under inversion of position and
336: momentum coordinates.
337:
338:
339: We focus on rectification effects that survive time averaging
340: and that are independent of the global phase $\alpha$ of the
341: laser beam. Typically, symmetry breaking effects that depend on
342: $\alpha$ are difficult to control (although not
343: impossible~\cite{kling}) since this requires an experimental
344: setup that both locks the absolute phase of the laser and has
345: control over the center of mass motion with respect to the
346: laboratory frame. Hence, the quantities of interest are the
347: mean position and momentum averaged over time and over $\alpha$:
348: %--------------------------------------------------------------------%
349: \begin{subequations}
350: \label{eq:mean}
351: \begin{align}
352: \langle \overline{\overline{\vect{x}}} \rangle_{\beta} &
353: = \lim_{\tau\to\infty}
354: \int_{-\tau/2}^{\tau/2} \frac{\ud t}{\tau}
355: \int_{0}^{2\pi}
356: \frac{\ud\alpha}{2\pi} \text{Tr}(\vect{x} \rho_{\beta}(\vect{x},\vect{p},t)); \\
357: \langle \overline{\overline{\vect{p}}} \rangle_\beta &
358: = \lim_{\tau\to\infty}
359: \int_{-\tau/2}^{\tau/2} \frac{\ud t}{\tau}
360: \int_{0}^{2\pi}
361: \frac{\ud\alpha}{2\pi} \text{Tr}(\vect{p}\rho_{\beta}(\vect{x},\vect{p},t));
362: \end{align}
363: \end{subequations}
364: %----------------------------------------------------------------%
365: where the double overbar indicates this kind of averaging. Here
366: the notation $\langle \cdots \rangle_{\beta}$ denotes the
367: classical ensemble average ($\beta=c$) or quantum expectation
368: value ($\beta=\ti{W}$), and the trace is an integration over the
369: $2N$-dimensional phase-space $(\vect{x}, \vect{p})$. When the
370: symmetry of the system is not broken, both $\langle
371: \overline{\overline{\vect{x}}} \rangle_{\beta}$ and $\langle
372: \overline{\overline{\vect{p}}} \rangle_{\beta}$ are zero. Below
373: we determine symmetries of the field and of the initial condition
374: that guarantee that this is indeed the case. When these
375: symmetries are violated a net dipole or current is expected to
376: appear.
377:
378: The fact that we are only interested in $\alpha$-independent
379: properties eliminates the necessity to invoke ergodicity in the
380: analysis. The average over $\alpha$ is sufficient to obviate any
381: initial-time preparation effects, which is the main role of the
382: ergodicity assumption in the purely classical analysis of
383: Ref.~\cite{flach}.
384:
385: We now tabulate the symmetries of the field that will be
386: relevant for our purposes. The field may change sign every half
387: a period $T$,
388: %----------------------------------------------------------------%
389: \begin{subequations}
390: \label{eq:sym}
391: \begin{align}
392: E(t + T/2) = &- E(t);\label{eq:sym1}\\
393: \intertext{or be symmetric or antisymmetric with respect to some time $t'$}
394: E(t - t') = &+ E(-(t-t')); \label{eq:sym2}\\
395: E(t - t') = &- E(-(t-t')). \label{eq:sym3}
396: \end{align}
397: \end{subequations}
398: %----------------------------------------------------------------%
399: Each of the symmetries in \eq{eq:sym} leads to a
400: transformation that leaves the equations of motion invariant
401: while changing the sign of the position and/or momentum
402: variables. They are identical for the quantum and classical
403: case. For example, if $E(t)$ satisfies \eq{eq:sym1}, then
404: $\mathcal{D}_{\beta}$ is invariant under $\mathcal{T}_1$ defined
405: as:
406: %----------------------------------------------------------------%
407: \begin{subequations}
408: \label{eq:trans}
409: \begin{align}
410: \label{eq:trans1}
411: \mathcal{T}_1&: \quad t \to t+T/2;\quad \vect{x} \to -\vect{x}; \quad \vect{p} \to -\vect{p}; \\
412: \intertext{where we have taken into account that under inversion of position and momenta,
413: $\frac{\partial^{\lambda_1+\cdots+\lambda_N }}
414: {\partial x_1^{\lambda_1}\cdots{\partial x_N^{\lambda_N}}}
415: \to -
416: \frac{\partial^{\lambda_1+\cdots+\lambda_N }}
417: {\partial x_1^{\lambda_1}\cdots{\partial x_N^{\lambda_N}}}$ and
418: $\frac{\partial^{\lambda_1+\cdots+\lambda_N }}
419: {\partial p_1^{\lambda_1}\cdots{\partial p_N^{\lambda_N}}}
420: \to -
421: \frac{\partial^{\lambda_1+\cdots+\lambda_N }}
422: {\partial p_1^{\lambda_1}\cdots{\partial p_N^{\lambda_N}}}$ since $\lambda_1 + \lambda_2 + \cdots +\lambda_N$ in \eq{eq:wignerdynamics} is odd.
423: Similarly, if $E(t)$ satisfies \eq{eq:sym2} [or \eq{eq:sym3}] then $\mathcal{D}_{\beta}$ is invariant under $\mathcal{T}_2$ [or $\mathcal{T}_3$], where}
424: \label{eq:trans2}
425: \mathcal{T}_2&:\, \quad t-t' \to -(t-t'); \quad \vect{x} \to \vect{x}; \quad \vect{p} \to -\vect{p}; \\
426: \label{eq:trans3}
427: \mathcal{T}_3&:\,\quad t-t' \to -(t-t'); \quad \vect{x} \to -\vect{x}; \quad \vect{p} \to \vect{p}.
428: \end{align}
429: \end{subequations}
430: %----------------------------------------------------------------%
431: Other temporal symmetries of the field exist but play no role in this analysis since they do not lead to invariance transformations that change the sign of the position and/or momentum variables.
432:
433: Now, given a solution to \eq{eq:dynamics}, $\rho_\beta(\vect{x}, \vect{p}, t)$, if $\mc{D}_\beta$ is invariant under $\mc{T}_\alpha$ one can generate another solution to the same equation by applying $\mc{T}_\alpha$ to $\rho_\beta(\vect{x}, \vect{p}, t)$. The new solutions $\rho_{\beta}^{(\alpha)}(\vect{x}, \vect{p}, t) = \mathcal{T}_\alpha \rho_{\beta}(\vect{x}, \vect{p}, t)$ generated by the invariance transformations in \eq{eq:trans} are:
434: %----------------------------------------------------------------%
435: \begin{subequations}
436: \label{eq:transfrho}
437: \begin{eqnarray}
438: \label{eq:transfrho1}
439: \rho_{\beta}^{(1)}(\vect{x}, \vect{p}, t) = &
440: \mathcal{T}_1 \rho_{\beta}(\vect{x}, \vect{p},t)
441: = & \rho_{\beta}(-\vect{x}, -\vect{p},t+T/2); \\
442: \rho_{\beta}^{(2)}(\vect{x}, \vect{p}, t) = & \mathcal{T}_2 \rho_{\beta}(\vect{x}, \vect{p}, t-t') = & \rho_{\beta}(\vect{x}, -\vect{p},-(t-t')); \\
443: \rho_{\beta}^{(3)}(\vect{x}, \vect{p}, t) = & \mathcal{T}_3 \rho_{\beta}(\vect{x}, \vect{p},t-t')
444: = & \rho_{\beta}(-\vect{x}, \vect{p},-(t-t')).
445: \end{eqnarray}
446: \end{subequations}
447: %----------------------------------------------------------------%
448: Further, if the original solution $\rho_{\beta}(\vect{x}, \vect{p}, t)$ predicts an average position $\langle \overline{\overline{\vect{x}}}\rangle_\beta$ and momenta $\langle \overline{\overline{\vect{p}}}\rangle_\beta$, the transformed solutions $\rho_{\beta}^{(\alpha)}(\vect{x}, \vect{p}, t)$ will predict a mean position $\langle \overline{\overline{\vect{x}}}\rangle_\beta^{(\alpha)}$ and/or momenta $\langle \overline{\overline{\vect{p}}}\rangle_\beta^{(\alpha)}$ that has the same magnitude but is opposite in sign:
449: %----------------------------------------------------------------%
450: \begin{subequations}
451: \label{eq:expectt}
452: \begin{eqnarray}
453: \label{eq:expect1}
454: \langle \overline{\overline{\vect{x}}}\rangle_\beta^{(1)} = -
455: \langle \overline{\overline{\vect{x}}}\rangle_\beta;
456: &\langle \overline{\overline{\vect{p}}}\rangle_\beta^{(1)} = -
457: \langle \overline{\overline{\vect{p}}}\rangle_\beta; \\
458: \label{eq:expect2}
459: \langle \overline{\overline{\vect{x}}}\rangle_\beta^{(2)} = +
460: \langle \overline{\overline{\vect{x}}}\rangle_\beta;
461: &
462: \langle \overline{\overline{\vect{p}}}\rangle_\beta^{(2)} = -
463: \langle \overline{\overline{\vect{p}}}\rangle_\beta; \\
464: \label{eq:expect3}
465: \langle \overline{\overline{\vect{x}}}\rangle_\beta^{(3)} = -
466: \langle \overline{\overline{\vect{x}}}\rangle_\beta;
467: & \langle \overline{\overline{\vect{p}}}\rangle_\beta^{(3)} = +
468: \langle \overline{\overline{\vect{p}}}\rangle_\beta.
469: \end{eqnarray}
470: \end{subequations}
471: %----------------------------------------------------------------%
472: % But, from \eq{eq:expect} it follows that the only way that this can happen is if symmetry breaking does not occur so that $\langle \overline{\overline{\vect{x}}}\rangle_\beta=0$ and/or $\langle \overline{\overline{\vect{p}}}\rangle_\beta=0$.
473:
474: The argument is completed by showing that the average position
475: and momenta predicted by $\rho_{\beta}(\vect{x}, \vect{p}, t)$
476: and $\rho_{\beta}^{(\alpha)}(\vect{x}, \vect{p}, t)$ coincide.
477: If this is the case, it follows from \eq{eq:expectt} that
478: symmetry breaking cannot occur. For this we exploit the
479: possible symmetries of the initial state:
480: \begin{subequations}
481: \label{eq:initialsym}
482: \begin{align}
483: \label{eq:initialsym1}
484: \rho_{\beta}(\vect{x}, \vect{p}, t_0) & = \rho_{\beta}(-\vect{x}, -\vect{p}, t_0);\\
485: \label{eq:initialsym2}
486: \rho_{\beta}(\vect{x}, \vect{p}, t_0) & = \rho_{\beta}(\vect{x}, -\vect{p}, t_0); \\
487: \label{eq:initialsym3}
488: \rho_{\beta}(\vect{x}, \vect{p}, t_0) & = \rho_{\beta}(-\vect{x}, \vect{p}, t_0).
489: \end{align}
490: \end{subequations}
491: %----------------------------------------------------------------%
492: The first one corresponds to a state with zero mean position and
493: momenta, while the second and third describe an initial state
494: with either zero mean momenta or zero mean position,
495: respectively.
496:
497:
498: Consider the case in which the equations of motion are $\mathcal{T}_1$ invariant.
499: The distributions $\rho_{\beta}(\vect{x}, \vect{p}, t)$ and $\rho_{\beta}^{(1)}(\vect{x}, \vect{p}, t)$
500: satisfy the same equation of motion but do not, in general, coincide. However,
501: if the initial condition for the original solution $\rho_{\beta}(\vect{x}, \vect{p}, t_0)$
502: is invariant under reflection [\eq{eq:initialsym1}] then
503: \be
504: \label{eq:inter1}
505: \rho_{\beta}^{(1)}(\vect{x}, \vect{p}, t_0 - T/2)=
506: \rho_{\beta}(-\vect{x}, -\vect{p}, t_0) = \rho_{\beta}(\vect{x},
507: \vect{p}, t_0) = \rho_{\beta}^{(1)}(-\vect{x}, -\vect{p}, t_0 -
508: T/2). \ee That is, the original and transformed solutions start
509: from the same initial distribution but at initial time they
510: experience a different value for the global phase of the field,
511: $E(t_0 +\alpha\frac{T}{2\pi})$ and
512: $E(t_0+(\alpha-\pi)\frac{T}{2\pi})=-E(t_0+ \alpha\frac{T}{2\pi})$
513: respectively. Since the averages in \eq{eq:mean} are independent
514: of $\alpha$, they coincide for the two solutions. Hence, no
515: rectification can be induced when the field satisfies
516: \eq{eq:sym1} and the initial condition satisfies
517: \eq{eq:initialsym1}.
518:
519: %In order to show this result explicitly it is convenient to express the Liouville evolution as a linear mapping among phase-space distributions $\Theta_\beta(t,t_0; E(t+ \alpha\tfrac{T}{2\pi})): \rho_\beta(\vect{x}, \vect{p}, t_0) \to \rho_\beta(\vect{x}, \vect{p}, t)$. This mapping can be written in the functional form
520: %\be
521: %\label{eq:funcmap}
522: %\Theta_\beta(t,t_0; E(t+ \alpha\tfrac{T}{2\pi}))[\rho_\beta(\vect{x}, \vect{p}, t_0)] \hspace{-0.1cm} = \hspace{-0.1cm} \rho_\beta(\vect{x}, \vect{p}, t).
523: %\ee
524: %With it one can express the net momentum of the transformed solution as
525: %%----------------------------------------------------------------%
526: %\begin{align*}
527: %\langle \overline{\overline{\vect{p}}} \rangle_{\beta}^{(1)} &
528: %= \lim_{\tau\to\infty}
529: %\int_{-\tau/2}^{\tau/2} \frac{\ud t}{\tau} \int_{0}^{2\pi} \frac{\ud\alpha}{2\pi} \text{Tr}\{\vect{p}\rho_{\beta}^{(1)}(\vect{x},\vect{p},t)\}\\
530: %& = \lim_{\tau\to\infty}
531: %\int_{-\tau/2}^{\tau/2} \frac{\ud t}{\tau} \int_{0}^{2\pi} \frac{\ud\alpha}{2\pi}
532: %\text{Tr}\left\{\vect{p}
533: %\Theta_\beta(t,t_0; E(t+ (\alpha-\pi)\tfrac{T}{2\pi}))[\rho_\beta(\vect{x}, \vect{p}, t_0)]\right\} \\
534: %\intertext{where we have used Eqs.~\eqref{eq:funcmap} and~\eqref{eq:inter1}. Letting $\alpha'=\alpha -\pi$ we obtain, }
535: %\langle \overline{\overline{\vect{p}}} \rangle_{\beta}^{(1)} &
536: %= \lim_{\tau\to\infty}
537: %\int_{-\tau/2}^{\tau/2} \frac{\ud t}{\tau}
538: %\left[ \int_{-\pi}^{0}\frac{\ud\alpha'}{2\pi}
539: %+ \int_{0}^{\pi}\frac{\ud\alpha'}{2\pi} \right]
540: % \text{Tr}\left\{\vect{p}
541: %\Theta_\beta(t,t_0; E(t+ \alpha'\tfrac{T}{2\pi}))[\rho_\beta(\vect{x}, \vect{p}, t_0)]\right\}.\\
542: %\intertext{Last, changing $\alpha = \alpha' +2\pi$ in the first integral over $\alpha'$, relabeling dummy variables and, using the periodicity of the field, it follows that }
543: %\langle \overline{\overline{\vect{p}}} \rangle_{\beta}^{(1)} & = \lim_{\tau\to\infty}
544: %\int_{-\tau/2}^{\tau/2} \frac{\ud t}{\tau} \int_{0}^{2\pi}\frac{\ud\alpha}{2\pi}
545: % \text{Tr}\left\{\vect{p}
546: %\Theta_\beta(t,t_0; E(t+ \alpha\tfrac{T}{2\pi}))[\rho_\beta(\vect{x}, \vect{p}, t_0)]\right\} \\
547: %& = \langle \overline{\overline{\vect{p}}} \rangle_\beta,
548: %\end{align*}
549: %as stated. An identical argument holds for the average position and $\langle \overline{\overline{\vect{x}}} \rangle_{\beta}^{(1)} = \langle \overline{\overline{\vect{x}}} \rangle_{\beta}$. Taking into account \eq{eq:expect1}, we conclude that $\langle \overline{\overline{\vect{x}}}\rangle_{\beta}=0$ and $\langle \overline{\overline{\vect{p}}}\rangle_{\beta}=0$ if the field satisfies \eq{eq:sym1} and the initial distribution satisfies \eq{eq:initialsym1}.
550:
551:
552: The argument for the two other cases is very similar. If the
553: field satisfies \eq{eq:sym2} [or \eq{eq:sym3}], the equation of
554: motion is $\mc{T}_2$ (or $\mc{T}_3$) invariant. Even when the
555: original $\rho_{\beta}(\vect{x}, \vect{p}, t)$ and transformed
556: $\rho_{\beta}^{(2)}(\vect{x}, \vect{p}, t)$ [or
557: $\rho_{\beta}^{(3)}(\vect{x}, \vect{p}, t)$] solutions obey the
558: same equation of motion, they do not need to coincide. However,
559: if the initial condition of the original solution satisfies the
560: symmetry in \eq{eq:initialsym2} [or \eq{eq:initialsym3}], then
561: \begin{align}
562: \rho_{\beta}^{(2)}(\vect{x}, \vect{p}, -t_0+t')&= \rho_{\beta}(\vect{x}, -\vect{p}, t_0) = \rho_{\beta}(\vect{x}, \vect{p}, t_0) = \rho_{\beta}^{(2)}(\vect{x}, -\vect{p}, -t_0+t')\\
563: \rho_{\beta}^{(3)}(\vect{x}, \vect{p}, -t_0+t')&= \rho_{\beta}(-\vect{x}, \vect{p}, t_0) = \rho_{\beta}(\vect{x}, \vect{p}, t_0) = \rho_{\beta}^{(3)}(-\vect{x}, \vect{p}, -t_0+t')
564: \end{align}
565: The transformed solution has the same initial condition as
566: the original one but as we had prepared the system a time $2t_0 - t'$ before.
567: The difference between the two solutions is that they experience a different
568: global laser phase at preparation time. Since we are not interested in effects
569: that depend on the global laser phase, the mean position and momenta
570: in \eq{eq:mean} for the original and transformed solution need to coincide.
571: However, from \eq{eq:expect2} [or \eq{eq:expect3}] we conclude that this
572: can only happen if $\langle \overline{\overline{\vect{p}}} \rangle_\beta = 0$
573: [or $\langle \overline{\overline{\vect{x}}} \rangle_\beta = 0$].
574:
575:
576: In summary, for spatially symmetric \emph{classical or quantum}
577: systems initially prepared in a symmetric state that satisfies
578: \eq{eq:initialsym}, net transport using time-periodic external
579: fields with zero temporal mean can only be generated if the field
580: violates the temporal symmetries in \eq{eq:sym}. Further, any
581: symmetry breaking effect that may be achieved with a field that
582: satisfies \eq{eq:sym} is necessarily due to an effect that
583: depends on the global phase of the laser (cf. Ref.~\cite{kling}).
584:
585:
586: It is natural to ask what kind of fields have sufficiently low
587: temporal symmetry to induce net transport. Monochromatic
588: sources satisfy all the symmetries in \eq{eq:sym} and, as a
589: consequence, cannot be used to induce symmetry breaking.
590: However, by adding a second frequency component to a
591: monochromatic source it is possible to lower the symmetry of the
592: field and induce symmetry breaking. For instance, a field like
593: %----------------------------------------------------------------%
594: \begin{equation}
595: \label{eq:field}
596: E(t) = \epsilon_{n\w} \cos(n \omega t + \phi_{n\omega}) +
597: \epsilon_{m\w} \cos(m \omega t + \phi_{m\omega}),
598: \end{equation}
599: %----------------------------------------------------------------%
600: where $n$ and $m$ are coprime integers so that $E(t)$ has a period $T= 2\pi/\omega$, satisfies \eq{eq:sym} only under special conditions. It satisfies \eqref{eq:sym1} only if $n$ and $m$ are odd. Thus, a field with $n=3$ and $m=1$, like the one used in the 1 vs. 3 photon control scenario~\cite{paul}, will not be symmetry breaking.
601: % symmetry~\eqref{eq:sym2} only when $\phi_{n\omega} - \frac{n}{m}\phi_{m\omega} = \pi (a + \frac{n}{m}b)$ and last; symmetry \eqref{eq:sym3} only when $\phi_{n\omega} - \frac{n}{m}\phi_{m\omega} = \frac{\pi}{2} ((1 + 2a) + \frac{n}{m}(1+2b))$, where $a,b=0, \pm 1, \pm 2, \cdots$.
602: However, a field with $n=2$ and $m=1$, like the one employed in the 1 vs. 2 scenario, does not satisfy \eq{eq:sym} and is expected to induce net dipoles and currents. These dipoles and currents are phase-controllable since, by varying the relative phase between the two components of the beam, the $\w+2\w$ field may satisfy \eq{eq:sym2} or~\eqref{eq:sym3} and thus rule out the possibility of inducing currents or dipoles, respectively. Explicitly, when $\phi_{2\omega} - 2\phi_{\omega}= 0, \pm\pi, \pm 2\pi, \cdots$ an $\w+2\w$ field satisfies \eq{eq:sym2} and zero currents are expected. Similarly, when $\phi_{2\omega} - 2\phi_{\omega}= \pm\frac{\pi}{2}, \pm\frac{3\pi}{2}, \cdots$ it fulfills symmetry \eqref{eq:sym3} and no dipoles can be induced. For all other choices of the phases an $\w+2\w$ field is expected to induce symmetry breaking.
603:
604: \section{Conclusions}
605: \label{sec:conclusions}
606:
607:
608: In conclusion, we have shown that the minimum conditions for the
609: generation of phase controllable asymmetry in spatially symmetric
610: quantum and classical systems using time-periodic external fields
611: with zero temporal mean are identical: anharmonicities in the
612: system's potential are required as is a driving field that violates
613: the temporal symmetries in \eq{eq:sym}. These conditions refer to
614: symmetry breaking effects that do not rely on the control of the absolute
615: phase of the field. The derived results provide necessary conditions for
616: the generation of asymmetry, applicable to all systems.
617: Additional sufficient conditions may be required, but these depend upon
618: the specific system under consideration.
619:
620: Further, we have shown that both quantum and
621: classical versions of the symmetry breaking effect can be accounted for
622: through a single space-time symmetry analysis of the equations of motion.
623: The anharmonicities in the potential permit the nonlinear
624: response of the system to the incident radiation that, through
625: harmonic mixing, and for fields that violate \eq{eq:sym}, can lead
626: to the generation of a phase-controllable DC component in the
627: photoinduced dipoles or currents.
628:
629: \section{Acknowledgments} This work was supported by NSERC Canada.
630:
631: \bibliography{symmetry_analysis}
632:
633: \end{document}
634:
635: