1: \documentclass[english,aps,prl,twocolumn]{revtex4}
2: \usepackage[T1]{fontenc}
3: \usepackage[latin1]{inputenc}
4: \usepackage{graphicx}
5: \usepackage{amssymb}
6:
7:
8: \begin{document}
9:
10: \title{Intrinsic Quantum Noise in Faraday Rotation Measurements of a Single
11: Electron Spin}
12:
13:
14: \author{Yanjun Ma and Jeremy Levy}
15:
16: \affiliation{Department of Physics and Astronomy, University of Pittsburgh, Pittsburgh, PA 15260}
17:
18: \date{2 September 2008}
19:
20: \begin{abstract}
21: Faraday rotation is one way to realize quantum non-demolition measurement
22: of electron spin in quantum dots. To describe Faraday rotation, semiclassical
23: models are typically used, based on quantized electron spin states
24: and classical electromagnetic fields. Such treatments neglect the
25: entanglement between electronic and photonic degrees of freedom that
26: produce intrinsic quantum noise, limiting the ultimate sensitivity
27: of this technique. We present a fully quantum-mechanical description
28: of Faraday rotation, and quantify this intrinsic noise. A method for
29: measuring the purity of a given spin state is suggested based on this
30: analysis.
31: \end{abstract}
32: \maketitle
33:
34: \section{Introduction}
35:
36: Because of the discovery of long-lived spin coherence in semiconductors such as GaAs\cite{kikkawa}, the essential requirement of manipulating spins for spintronics and quantum information is now possible. The first quantum computing
37: proposal of Loss and DiVincenzo used electron spin qubits in semiconductor
38: quantum dots (QD)\cite{divi} and forecast the importance of measuring
39: single electrons and their spins.
40:
41: The first step to realizing coherent manipulation of a single electron
42: spin is to orient the spin. Such orientation can be achieved optically
43: (by exciting with circularly polarized light)\cite{opti}, electrically
44: (by driving the electrons toward a ferromagnetic surface)\cite{asw}
45: or thermodynamically (by application of a uniform magnetic field at
46: low temperatures). Photoluminescence (PL) allows
47: for measurement of electron spin polarization through the relation
48: between the circular polarization of light and electron spin orientation.
49: However, PL is destructive in that it involves recombination of the
50: electron with a hole. PL measurements are intrinsically limited by
51: the lifetime of the state, and it is not possible to monitor electron
52: spin continuously. Furthermore, unless one uses a technique such as
53: time-resolved upconversion\cite{upc1,upc2,upc3} or streak camera
54: measurements, dynamical information is lost.
55:
56: Time-resolved Faraday and Kerr rotation methods (hereafter referred
57: to as Faraday rotation) have been extensively developed\cite{exp1,exp2,exp3},
58: and allow one to probe the spin dynamics of a single electron in a quantum
59: dot. Faraday rotation results from a fundamental interaction between
60: electronic and photonic degrees of freedom. Seigneur et al.\cite{spfe}have
61: proposed a scheme to implement quantum computation by using the single
62: photon Faraday effect. However, in most semiconductors the Faraday
63: effect is usually quite weak, corresponding to rotation angles $\theta_{F}\sim10^{-5}\mathrm{rad}$
64: for single electrons. Dynamic information is usually obtained using
65: pump-probe optical techniques: a circularly polarized pump beam creates
66: an initially spin-polarized electron population, and a probe beam subsequently
67: interrogates the spin state at a later time. The experiment is performed
68: repeatedly as a function of the delay to obtain a time-resolved signal
69: with an acceptably high signal-to-noise ratio. In the case of a single
70: electron in a quantum dot, spin coherence can be achieved in the following
71: manner: the quantum dot is configured (either through biasing or doping
72: ) to begin in a state that contains a single electron in the conduction
73: band and no holes in the valence band. The quantum dot is excited,
74: promoting a second electron into the conduction band and leaving behind
75: a hole in the valence band. This state is often referred to as a {}``trion''.
76: After one of the electrons recombines with the hole, the remaining
77: electron spin is partially polarized. A linearly polarized probe pulse
78: measures the spin of this electron via the Faraday effect. In most
79: cases, the electron neither begins in a pure state nor remains in
80: one. Hyperfine interactions with nuclear spins quickly produce a mixed
81: state on time scales \textasciitilde{}1-10 ns\cite{nuc1,nuc2,nuc3,nuc4,nuc5,nuc6,nuc7}.
82: In this paper, we study the noise introduced by the mixed quantum state of the electron spin analytically and numerically. In this paper, our previous analysis\cite{patr} about the noise is extended to a more formal quantum mechanical frame. Since it's from the spin state itself, we call it intrinsic noise.
83:
84: %
85: \begin{figure}
86: \centering \includegraphics[width=0.5\textwidth,height=0.3\textheight]{band}
87:
88: \caption{light induced interband transition}
89:
90:
91: \label{band}
92: \end{figure}
93:
94:
95:
96: \section{Theoretical Model}
97:
98: Here we discuss in detail the quantum-mechanical source of this noise
99: using a theoretical model that treats both the electron and light
100: field quantum mechanically. We model the interaction between a single
101: electron in a QD and a linearly polarized monochromatic probe laser
102: field. The Hamiltonian for the photon field can be written as
103:
104: \begin{equation}
105: H_{P}=\hbar\omega_{P}(a_{L}^{\dag}a_{L}+a_{R}^{\dag}a_{R}),
106: \end{equation}
107:
108: where $\omega_{P}$ is the optical frequency of the probe laser, $a_{L}^{\dag}$
109: and $a_{L}$ are creation and annihilation operators for left circularly
110: polarized (LCP) photons; $a_{R}^{\dag}$ and $a_{R}$ are creation
111: and annihilation operators for right circularly polarized (RCP) photons.
112: Due to optical selection rules\cite{opti} spin-up (spin-down) electrons
113: interact only with LCP (RCP) photons(See FIG.~\ref{band}). The raising
114: and lowering operators satisfy boson commutation relations \[
115: [a_{m},a_{n}^{\dag}]=\delta_{mn},(m,n=L,R);\]
116: \[
117: [a_{m},a_{n}]=0,[a_{m}^{\dag},a_{n}^{\dag}]=0.\]
118:
119:
120: The electron state is quantized as well. We assume that the electron
121: resides in the conduction band quantum-confined ground state in an
122: s orbital, which means it has total angular momentum $J=\frac{1}{2}$.
123: In the valence band, the electronic ground states are constructed from
124: p-orbitals, and hence the total angular momentum is $J=\frac{3}{2}$.
125: The Hamiltonian for the electron is given by \cite{yama}
126:
127: \begin{equation}
128: H_{e}=\hbar\omega_{e}(\sigma_{uz}+\sigma_{dz}),
129: \end{equation}
130:
131: where \[
132: \sigma_{uz}=b_{cu}^{\dag}b_{cu}-b_{vu}^{\dag}b_{vu},\]
133: \[
134: \sigma_{dz}=b_{cd}^{\dag}b_{cd}-b_{vd}^{\dag}b_{vd};\]
135: subscript \char`\"{}$c$\char`\"{} and \char`\"{}$v$\char`\"{} indicate
136: conduction band and valence band respectively; subscripts \char`\"{}$u$\char`\"{}
137: and \char`\"{}$d$\char`\"{} refer to spin-up or spin-down states
138: of the electron. The fermion operators satisfy anticommutation relations:
139: \[
140: \{b_{i\mu},b_{j\nu}^{\dag}\}=\delta_{ij}\delta_{\mu\nu},\]
141: \[
142: \{b_{i\mu},b_{j\nu}\}=0,\{b_{i\mu}^{\dag},b_{j\nu}^{\dag}\}=0,\]
143: where $i$ and $j$ indicate conduction band or valence band, and
144: $\mu$ and $\nu$ indicate spin-up or spin-down. Heavy-hole
145: and light-hole intermixing is neglected for simplicity and because it is not expected
146: to affect qualitatively our results. Only the heavy-hole subband is
147: accounted for in our calculation. A LCP photon couples to a transition
148: between $|+\frac{1}{2}>$ and $|+\frac{3}{2}>$, while a RCP photons
149: couples to a transition between $|-\frac{1}{2}>$ and $|-\frac{3}{2}>$.
150: The interaction Hamiltonian is given by
151:
152: \begin{equation}
153: H_{I}=\lambda_{Lu}(a_{L}\sigma_{u+}+a_{L}^{\dag}\sigma_{u-})+\lambda_{Rd}(a_{R}\sigma_{d+}+a_{R}^{\dag}\sigma_{d-}),
154: \end{equation}
155:
156: where \[
157: \lambda_{Lu}\propto<+\frac{1}{2}|x+iy|+\frac{3}{2}>,\]
158: \[
159: \lambda_{Rd}\propto<-\frac{1}{2}|x-iy|-\frac{3}{2}>,\]
160: \[
161: \sigma_{u+}=b_{cu}^{\dag}b_{vu},\sigma_{u-}=\sigma_{u+}^{\dag},\]
162: \[
163: \sigma_{d+}=b_{cd}^{\dag}b_{vd},\sigma_{d-}=\sigma_{d+}^{\dag}.\]
164: The full Hamiltonian of the entire system is given by \[
165: H=H_{P}+H_{e}+H_{I}\]
166: By applying the Wigner-Eckart theorem, it can be shown that the two
167: coupling strengths $\lambda_{Lu}$ and $\lambda_{Rd}$ must be equal
168: ($\lambda_{Lu}$=$\lambda_{Rd}$$\equiv$$\lambda$). Based on the
169: defining anticommutation relations, it can be explicitly shown that
170: $\sigma_{\mu z}$,$\sigma_{\mu+}$ and $\sigma_{\mu-}$ have the following
171: commutation relations \[
172: [\sigma_{\mu+},\sigma_{\nu-}]=\delta_{\mu\nu}\sigma_{\mu z},\]
173: \[
174: [\sigma_{\mu z},\sigma_{\nu+}]=2\delta_{\mu\nu}\sigma_{\mu+},\]
175: \[
176: [\sigma_{\mu z},\sigma_{\nu-}]=2\delta_{\mu\nu}\sigma_{\mu-}.\]
177: These commutation relations for $\sigma_{\mu z}$,$\sigma_{\mu+}$
178: and $\sigma_{\mu-}$ are formally identical to those for the Pauli operators, even
179: though they are actually products of fermionic creation and annihilation
180: operators. This feature makes it possible to find an analytic solution
181: within the Heisenberg picture\cite{solution}. In the limit where
182: the coupling strength is much smaller than the incident photon frequency
183: or the characteristic frequency of the electron, the approximate solution
184: for photon operators is as follows:
185: \begin{equation}
186: a_{L}^{\dag}(t)=e^{-it\Omega\sigma_{uz}}a_{L}^{\dag}+g(t)(\sigma_{u+}+\alpha\sigma_{uz}a_{L}^{\dag}),\label{eq:al}
187: \end{equation}
188:
189:
190: \begin{equation}
191: a_{R}^{\dag}(t)=e^{-it\Omega\sigma_{dz}}a_{R}^{\dag}+g(t)(\sigma_{d+}+\alpha\sigma_{dz}a_{R}^{\dag}),\label{eq:ar}
192: \end{equation}
193:
194: where \[
195: \alpha=\frac{\lambda}{\omega_{P}-\omega_{e}},\]
196: \[
197: \Omega=\lambda\alpha,\]
198: \[
199: g(t)=\alpha(1-e^{-i(\omega_{P}-\omega_{e})t}).\]
200: This approximate solution is correct only when the coupling strength $\lambda$ is much smaller than $\omega_{e}$ and $\omega_{P}$. In the section of RESULTS, one can see this criteria is satisfied in the sense that the coupling strength of our sample is in the order of \textasciitilde{}$10^{9}$Hz, but the frequency of laser and characteristic frequency of electron is in the order of \textasciitilde{}$10^{15}$Hz. This solution, therefore, is a very good approximation and based on this, one can derive Faraday rotation angle.
201:
202:
203:
204: \subsection{Faraday rotation operator}
205:
206: Quantum Stokes operators can be used to describe Faraday rotation.
207: They are the quantum-mechanical analogue of classical Stokes parameters.
208: Classical Stokes parameters are defined as the following\cite{stoke1}
209: \[
210: \left\{ \begin{array}{ll}
211: S_{0}=E_{x}^{*}E_{x}+E_{y}^{*}E_{y}\\
212: S_{1}=E_{x}^{*}E_{x}-E_{y}^{*}E_{y}\\
213: S_{2}=E_{x}^{*}E_{y}+E_{y}^{*}E_{x}\\
214: S_{3}=E_{x}^{*}E_{y}-E_{y}^{*}E_{x}\end{array}\right.\]
215: In electrodynamics, the polarization of light can be parameterized
216: by two angles $\varphi$ and $\chi$ in the polarization ellipse. There
217: is a one-to-one correspondence between the polarization-ellipse representation
218: and the Stokes representation (See FIG.~\ref{stokerep}). %
219: \begin{figure}
220: \centering \includegraphics[width=0.5\textwidth,height=0.3\textheight]{Stoke}
221:
222: \caption{The left figure shows the polarization ellipse in real space. The right figure is the Stokes representation of the same polarization.}
223:
224:
225: \label{stokerep}
226: \end{figure}
227:
228:
229: Once the light field is known, the Stokes parameters can be computed.
230: The physical interpretation of $S_{0}$ is the light intensity; hence,
231: all parameters can be normalized to $S_{0}$ (See FIG.~\ref{stokepolar}).
232:
233: %
234: \begin{figure}[h]
235: \centering \includegraphics[width=0.5\textwidth,height=0.3\textheight]{polarization}
236:
237: \caption{(a)-(f)Different polarization defined in terms of Stokes parameters. $S_{1}, S_{2}$ and $S_{3}$ have all been normalized to $S_{0}$.}
238: \label{stokepolar}
239:
240: \end{figure}
241:
242:
243: Quantum Stokes operators are defined in the following way\cite{stoke2,stoke3}
244: \[
245: \left\{ \begin{array}{l}
246: S_{0}=a_{L}^{\dag}a_{L}+a_{R}^{\dag}a_{R}\\
247: S_{1}=a_{L}^{\dag}a_{R}+a_{R}^{\dag}a_{L}\\
248: S_{2}=i(a_{L}^{\dag}a_{R}-a_{R}^{\dag}a_{L})\\
249: S_{3}=a_{R}^{\dag}a_{R}-a_{L}^{\dag}a_{L}\end{array}\right.\]
250: Information about polarization is obtained by calculating the expectation
251: values of these operators. In a typical Faraday experiment, the probe
252: light is linearly polarized at a $45^{\circ}$ angle with respect to a final
253: polarizing beamsplitter. After the interaction between the probe light
254: and the electron, the polarization of the transmitted light will be rotated
255: from its initial position by an angle $\theta_{F}$, known as the
256: Faraday rotation angle. In the Stokes representation, the initial
257: polarization vector lies along the positive $S_{2}$ axis. Faraday
258: rotation will result in a rotation of the vector within the $S_{1}-S_{2}$
259: plane (See FIG.~\ref{fangle}). This vector P is confined to the plane
260: as long as there is no circular dichroism that can lead to a non-zero
261: expectation value for $S_{3}$.
262:
263: %
264: \begin{figure}
265: \centering \includegraphics[width=0.5\textwidth,height=0.3\textheight]{fangle}
266:
267: \caption{Definition of Faraday rotation angle. P indicates the polarization vector of light field.}
268:
269:
270: \label{fangle}
271: \end{figure}
272:
273:
274: In our calculations, we aim to reproduce the overall magnitude of
275: the rotation angle that has been reported in experimental work\cite{exp1, exp2, exp3}.
276: The experimentally observed rotation angle is small: $\theta_{F}\sim10^{-5}\mathrm{rad}$.
277: Hence, it can be expressed as
278: \begin{equation}
279: \theta_{F}=\frac{1}{2}tan^{-1}(\frac{<S_{1}>}{<S_{2}>})\approx\frac{<S_{1}>}{2<S_{2}>}.
280: \end{equation}
281:
282:
283:
284: \section{Results}
285:
286: In our calculation, a coherent state $|\nu_{L},\nu_{R}>$is used for
287: the light field, where $|\nu_{L}^{2}|$ and $|\nu_{R}^{2}|$ are the
288: average number of left and right circularly polarized photons. These
289: states satisfy the canonical eigenvalue equations for the (non-Hermitian)
290: photon annihilation operators: \[
291: a_{L}|\nu_{L},\nu_{R}>=\nu_{L}|\nu_{L},\nu_{R}>,\]
292: \[
293: a_{R}|\nu_{L},\nu_{R}>=\nu_{R}|\nu_{L},\nu_{R}>,\]
294:
295:
296: Using the form $\nu_{L}=N_{L}e^{i\theta_{L}}$ and $\nu_{R}=N_{R}e^{i\theta_{R}}$,
297: the expectation value of Stokes operators in this coherent state can
298: be found
299: \begin{equation}
300: \left(\begin{array}{l}
301: <S_{0}>\\
302: <S_{1}>\\
303: <S_{2}>\\
304: <S_{3}>\end{array}\right)=\left(\begin{array}{l}
305: N_{L}^{2}+N_{R}^{2}\\
306: 2N_{L}N_{R}cos(\theta_{L}-\theta_{R})\\
307: 2N_{L}N_{R}sin(\theta_{L}-\theta_{R})\\
308: N_{R}^{2}-N_{L}^{2}\end{array}\right).
309: \end{equation}
310: In order to start with $+45^{\circ}$ linearly polarized light, the
311: following condition must be satisfied \[
312: \left\{ \begin{array}{l}
313: N_{L}^{2}=N_{R}^{2}\\
314: \theta_{L}-\theta_{R}=\frac{\pi}{2}\end{array}\right.\]
315:
316:
317: To describe the mixed state of electron, a density matrix formula
318: is employed. \begin{equation}
319: \rho_{e}=\tau|+\frac{1}{2}><+\frac{1}{2}|+(1-\tau)|-\frac{1}{2}><-\frac{1}{2}|\end{equation}
320:
321:
322: Here, $\tau$ is a parameter that varies between 0 and 1. For $\tau=0$
323: and $\tau=1$, one has a pure state, while $\tau=1/2$ corresponds
324: to a fully mixed (unpolarized) state. Because the electron Hamiltonian
325: is expressed in terms of creation and annihilation operators, caution
326: must be taken when applying those operators onto electron state. When
327: operators for spin-up electron are applied to the spin-up state, one
328: obtains \[
329: \sigma_{uz}|+\frac{1}{2}>=|+\frac{1}{2}>,\]
330: \[
331: \sigma_{uz}|+\frac{3}{2}>=-|+\frac{3}{2}>,\]
332: and \[
333: \sigma_{u+}|+\frac{1}{2}>=0,\]
334: \[
335: \sigma_{u+}|+\frac{3}{2}>=|+\frac{1}{2}>,\]
336: and \[
337: \sigma_{u-}|+\frac{1}{2}>=|+\frac{3}{2}>,\]
338: \[
339: \sigma_{u-}|+\frac{3}{2}>=0.\]
340:
341:
342: Spin-down operators have the same rules when applied to the spin-down
343: state. If a spin-up operator operates on a spin-down state, however,
344: one gets zero. For example, \[
345: \sigma_{uz}|-\frac{1}{2}>=(b_{cu}^{\dag}b_{cu}-b_{vu}^{\dag}b_{vu})|-\frac{1}{2}>=0.\]
346:
347:
348: The initial state of the whole system is then \begin{equation}
349: \rho_{0}(\tau)=|\nu_{L},\nu_{R}><\nu_{L},\nu_{R}|\otimes(\tau|\uparrow_{-}>+(1-\tau)|\downarrow_{-}>)\end{equation}
350:
351:
352: According to the solution (4) and (5), the analytical expression for
353: $S_{1}$ and $S_{2}$ can be obtained and the expectation values calculated
354: \[
355: <S_{1}>=Tr(S_{1}(t)\rho_{0}(\tau)),\]
356: \[
357: <S_{2}>=Tr(S_{1}(t)\rho_{0}(\tau)).\]
358:
359:
360: The rotation angle is given by
361: \begin{equation}
362: \theta_{F}(t,\tau)=\frac{Tr(S_{1}(t)\rho_{0}(\tau))}{2Tr(S_{2}(t)\rho_{0}(\tau))},
363: \end{equation}
364: After some algebra, one finds the following expression for the Faraday
365: rotation:
366: \begin{equation}
367: \theta_{F}(t,\tau)=(2\tau-1)(\frac{\lambda^{2}}{\delta^{2}}sin(\delta t)-sin(\frac{\lambda^{2}}{\delta}t)),
368: \end{equation}
369: where $\delta\equiv\omega_{P}-\omega_{e}.$ For initial pure spin-up
370: state $\tau=1$, the rotation angle is
371: \begin{equation}
372: \theta_{+}\equiv\theta_{F}(t,1)=(\frac{\lambda^{2}}{\delta^{2}}sin(\delta t)-sin(\frac{\lambda^{2}}{\delta}t)).
373: \end{equation}
374: For initial pure spin-down state $\tau=0$, the rotation angle is
375: \begin{equation}
376: \theta_{-}\equiv\theta_{F}(t,0)=-(\frac{\lambda^{2}}{\delta^{2}}sin(\delta t)-sin(\frac{\lambda^{2}}{\delta}t)).
377: \end{equation}
378:
379:
380: The fluctuation is given by
381: \begin{equation}
382: \Delta\theta_{F}(t,\tau)=\frac{\sqrt{Tr(S_{1}^{2}(t)\rho_{0}(\tau))-Tr(S_{1}(t)\rho_{0}(\tau))^{2}}}{2Tr(S_{2}(t)\rho_{0}(\tau))}.
383: \end{equation}
384:
385:
386: %
387: \begin{figure}
388: \centering \includegraphics[width=0.5\textwidth,height=0.3\textheight]{angle3d.JPG}
389: \hfill{}\includegraphics[width=0.5\textwidth,height=0.3\textheight]{angle}
390:
391: \caption{(a)Faraday rotation as a function of time and parameter $\tau$. (b)Faraday
392: rotation angle for two pure states. $\tau=1$ corresponds to a spin-up
393: state, while $\tau=0$ corresponds to a spin-down state. }
394:
395:
396: \label{angle}
397: \end{figure}
398:
399:
400: From (11), (12) and (13), the following intuitive result can be proven
401: very easily: \begin{equation}
402: \theta_{F}(t,\tau)=\tau\theta_{+}+(1-\tau)\theta_{-},\end{equation}
403:
404:
405: where $\theta_{+}$ ($\theta_{-}$) is the Faraday rotation angle
406: for an initial state which is a pure spin-up (spin-down) state.
407:
408:
409:
410: An analytical derivation shows that the fluctuation is a function
411: of both photon number and the initial electron state.
412: \begin{equation}
413: \Delta\theta_{F}(t,\tau)=\sqrt{\frac{1}{4N}+\tau(1-\tau)(\theta_{+}-\theta_{-})^{2}},
414: \end{equation}
415: The second term under the square root is the so-called intrinsic noise term.
416:
417:
418: Numerical simulation is done so that we compare our analytical calculation to recent Kerr rotation experimental results on single electrons. From Berezosky et al.\cite{exp1}, one finds from
419: a PL plot that the energy for a neutral exciton is about $1.633\mathrm{meV}$.
420: That corresponds to the band gap between the top of the valence band and
421: the bottom of the conduction band. From this number, the frequency $\omega_{e}=\frac{E}{\hbar}=2.48\times10^{15}Hz$.
422: Choosing probe light of wavelength $760\mathrm{nm}$, which means the
423: frequency is $\omega_{P}=2.47\times10^{15}Hz$. From the paper\cite{cpl},
424: one can take the value of the coupling strength to be $\lambda=\mathrm{98GHz}$. In our experiment, the probe power is about $1.57\mathrm{\mu W}$, the corresponding photon number is about $5\times10^{5}$. In the simulation, the interaction time between the spin and the photon is set to be $20\mathrm{ps}$. Notice that as expected, if the initial electron state is a pure state, the rotation angle has opposite values for the spin-up state and spin-down state,
425: respectively (See FIG.~\ref{angle}). For pure spin-up states or
426: spin-down states, the fluctuation (quantum noise) scales with photon
427: number $N$ as $N^{-1/2}$, as expected for shot noise. For mixed
428: states or superposition states, the fluctuation saturates even when
429: the photon number approaches infinity (See FIG.~\ref{noise}).
430:
431: \begin{figure}
432: \centering
433: \includegraphics[width=0.5\textwidth,height=0.3\textheight]{deviation}
434: \hfill{}
435: \includegraphics[width=0.5\textwidth,height=0.3\textheight]{noise}
436:
437: \caption{(a)Fluctuation of Faraday rotation angle as a function of time and parameter $\tau$. (b)Shot noise and intrinsic noise as a function of photon number N.Shot noise (black) is from a pure spin-up(spin-down) state, while intrinsic noise (green) is from a maximally mixed state. In the simulation, the interaction time is chosen to be $20\mathrm{ps}$ in order to make the splitting more obvious, in which case the intrinsic noise saturates at about $\Delta\theta_{F0}=20mrad$.}
438:
439:
440: \label{noise}
441: \end{figure}
442:
443:
444:
445:
446: One scheme to measure $\tau$ is proposed here. Suppose the photon number is so large that the shot noise term in Equation (16) could be neglected. Notice that $\theta_{+}=\theta_{-}$ and when the rotation
447: angle is zero, according to Equation (11) it means $\tau$ is $\frac{1}{2}$.
448: If the value $\tau=\frac{1}{2}$ is used in Equation (16), one obtains
449: $\Delta\theta_{F0}=\theta_{+}$. This result implies that one can
450: use the measured values of Faraday angle fluctuation at an extreme
451: value ($\Delta\theta_{F}$) and at a zero crossing ($\Delta\theta_{F0}$)
452: to measure the purity of the spin state as quantified by $\tau$:
453: \begin{equation}
454: \tau=\frac{1}{2}(1\pm\sqrt{\frac{\Delta\theta_{F0}^{2}-\Delta\theta_{F}^{2}}{\Delta\theta_{F0}^{2}-\frac{1}{4N}}}).
455: \end{equation}
456:
457:
458: In the limit of a large number of photons (i.e., where shot noise can
459: be neglected), the above expression simplifies further
460: \begin{equation}
461: \tau=\frac{1}{2}(1\pm\sqrt{1-\frac{\Delta\theta_{F}^{2}}{\Delta\theta_{F0}^{2}}}).
462: \end{equation}
463:
464:
465: The above analysis is based on the assumption that every device in
466: the experiment is perfect, and the noise is only introduced by quantum
467: state of the electron spin itself. This is, however, not the case in the
468: real experiment. Suppose the overall noise $\Delta\theta_{B}$ is
469: white noise for the bandwidth in which the experiment is done. It
470: serves as background noise and can be measured by detuning the probe
471: laser, for example. This background contribution can be subtracted
472: from the measured noise $\Delta\theta_{M}$ and $\Delta\theta_{M0}$,
473: where $\Delta\theta_{M}$ indicates the measured noise level at extreme
474: points while $\Delta\theta_{M0}$ represents the measured noise at
475: the zero-crossing point. It makes sense to assume the fluctuation due
476: to quantum states is not correlated with the white noise in the device,
477: therefore subtracting the background noise from the measured noise
478: gives the fluctuation due to quantum states \[
479: \Delta\theta_{F}^{2}=\Delta\theta_{M}^{2}-\Delta\theta_{B}^{2},\]
480: and \[
481: \Delta\theta_{F0}^{2}=\Delta\theta_{M0}^{2}-\Delta\theta_{B}^{2}.\]
482: The above equation therefore becomes \begin{equation}
483: \tau=\frac{1}{2}(1\pm\sqrt{1-\frac{\Delta\theta_{M}^{2}-\Delta\theta_{B}^{2}}{\Delta\theta_{M0}^{2}-\Delta\theta_{B}^{2}}}).\end{equation}
484:
485:
486: Notice that in the above equation, $\Delta\theta_{M}$ is the \char`\"{}real\char`\"{}
487: noise we see at the extreme points of the rotation angle in the actual
488: experiment. This noise has two sources: external noise, which is $\Delta\theta_{B}$
489: and intrinsic noise, which is $\Delta\theta_{F}$. In the actual experiment,
490: $\Delta\theta_{M}$ should be larger than $\Delta\theta_{B}$ due
491: to the fact that the pump for the electron spin is not perfect, therefore
492: the spin that interacts with photon is in a mixed state. However,
493: if $\Delta\theta_{M}=\Delta\theta_{B}$, that means the intrinsic
494: noise contribution is zero. From Eq. (16), one can see in the limit
495: of large photon number, $\Delta\theta_{F}=0$ indicates that $\tau$
496: is either 0 or 1, which is consistent with the result if one plugs
497: $\Delta\theta_{M}=\Delta\theta_{B}$ into Eq. (19). In other words,
498: if in the real experiment, one observes $\Delta\theta_{M}=\Delta\theta_{B}$,
499: then the pumped spin is in either pure spin-up or spin-down state
500: and one can also pin down the orientation of spin by looking at the
501: sign of measured rotation angle.
502:
503:
504: \section{Conclusion}
505:
506: Using a quantum-mechanical model of Faraday rotation, we find that
507: both the Faraday rotation angle and the fluctuation are functions of the initial
508: electron spin state. If the electron spin is initially in a mixed
509: state, intrinsic noise fluctuations will contain not only shot noise
510: but also intrinsic noise due to weak measurement of the electron's
511: spin state. The reason that this intrinsic noise appears in this scheme
512: is that the measurement done here is non-destructive, and differs
513: from a projective measurement, which causes the collapse of the electron
514: spin wave function to a certain spin direction. Analysis of the noise
515: spectrum should enable quantification of the purity of a given spin
516: state.
517:
518: \begin{thebibliography}{10}
519: \bibitem{kikkawa}J. M. Kikkawa and D. D. Awschalom, Phys. Rev. Lett.
520: \textbf{80}, 4313 (1998).
521:
522: \bibitem{divi}D. Loss and D. P. DiVincenzo, Phys.Rev.A \textbf{57},120
523: (1998).
524:
525: \bibitem{opti}F. Meier and B. P. Zakharchenya, Optical Orientation
526: (Elsevier Science Publishers B. V., Amsterdam, 1984).
527:
528: \bibitem{asw}R. J. Epstein, I. Malajovich, R. K. Kawakami, Y. Chye,
529: M. Hanson, P. M. Petroff, A. C. Gossard, and D. D. Awschalom, Phys.
530: Rev. B, \textbf{65}, 121202 (2002)
531:
532: \bibitem{upc1}J. F. Smyth, D. A. Tulchinsky, D. D. Awschalom, N.
533: Samarth, H. Luo, and J. K. Furdyna, Phys. Rev. Lett. \textbf{71},
534: 601 (1993).
535:
536: \bibitem{upc2}M. P. Hehlen, G. Frei, and H. U. Güdel, Phys. Rev.
537: B \textbf{50}, 16264 (1994).
538:
539: \bibitem{upc3}R. Kumar, A. S. Vengurlekar, S. S. Prabhu, J. Shah,
540: and L.N.Pfeiffer, Phys. Rev. B \textbf{54}, 4891 (1996).
541:
542: \bibitem{exp1}J. Berezovsky, M. H. Mikkelsen, O. Gywat, N. G. Stoltz,
543: L. A. Coldren, and D.D.Awschalom, Science \textbf{314}, 1916 (2006).
544:
545: \bibitem{exp2}M. Atature, J. Dreiser, A. Badolato, and A. Imamoglu,
546: Nature Physics \textbf{521}, 101 (2007).
547:
548: \bibitem{exp3}M. H. Mikkelsen, J. Berezovsky, N. G. Stoltz, L. A.
549: Coldren, and D.D.Awschalom, Nature Physics \textbf{736}, 1 (2007).
550:
551: \bibitem{spfe}H. P. Seigneur, M. N. Leuenberger and W. V. Schoenfeld,
552: J. Appl. Phys. \textbf{104}, 014307 (2008).
553:
554: \bibitem{nuc1}I. A. Merkulov, Al. L. Efros, and M.Rosen, Phys. Rev.
555: B \textbf{65}, 205309 (2002).
556:
557: \bibitem{nuc2}A. V. Khaetskii, D. Loss, and L.Glazman, Phys. Rev.
558: Lett \textbf{88}, 186802 (2002).
559:
560: \bibitem{nuc3}A. C. Johnson, J. R. Petta, J. M. Taylor, A. Yacoby,
561: M. D. Lukin, C. M. Marcus, M. P. Hanson, and A.C.Gossard, Nature \textbf{435},
562: 925-928 (2005)
563:
564: \bibitem{nuc4}A. S. Bracker, E. A. Stinaff, D. Gammon, M. E. Ware,
565: J. G. Tischler, A. Shabaev, Al. L. Efros, D. Park, D. Gershoni, V.
566: L. Korenev, and I. A. Merkulov, Phys. Rev. Lett. \textbf{94}, 047402
567: (2005).
568:
569: \bibitem{nuc5}P. F. Braun, X. Marie, L. Lombez, B. Urbaszek, T. Amand,
570: P. Renucci, V. K. Kalevich, K. V. Kavokin, O. Krebs, P. Voisin, and
571: Y. Masumoto, Phys. Rev. Lett. \textbf{94}, 116601 (2005).
572:
573: \bibitem{nuc6}F. H. L. Koppens, J. A. Folk, J. M. Elzerman, R. Hanson,
574: L. H. Willems van Beveren, I. T. Vink, H. P. Tranitz, W. Wegscheider,
575: L. P. Kouwenhoven, and L. M. K. Vandersypen, Science \textbf{309},
576: 1346-1350 (2005).
577:
578: \bibitem{nuc7}J. R. Petta, A. C. Johnson, J. M. Taylor, E. A. Laird,
579: A. Yacoby, M. D. Lukin, C. M. Marcus, M. P. Hanson, and A. C. Gossard,
580: Science \textbf{309}, 2180-2184 (2005).
581:
582: \bibitem{patr}P. Irvin, P. S. Fodor, and J. Levy, Optics Express
583: \textbf{15}, 11756 (2007).
584:
585: \bibitem{yama}M. Sugita, S. Machida, and Y. Yamamoto, arxiv:quant-ph/0301064v1
586: (2003).
587:
588: \bibitem{solution}J. R. Ackerhalt and K. Rzazewski, Phys. Rev. A
589: \textbf{12}, 2549 (1975).
590:
591: \bibitem{stoke1}L. Mandel and E. Wolf, Optical Coherence and Quantum
592: Optics (Cambridge Unversity Press, New York, 1995).
593:
594: \bibitem{stoke2}N. Korolkova, G. Leuchs, R. Loudon, T. C. Ralph,
595: and C. Silberhorn, Phys. Rev. A \textbf{65}, 052306 (2002).
596:
597: \bibitem{stoke3}Y. Takahashi, K. Honda, N. Tanaka, K. Toyoda, K.
598: Ishikawa, and T. Yabuzaki, Phys. Rev. A \textbf{60}, 4974 (1999).
599:
600: \bibitem{nuc8}A. V. Khaetskii, D. Loss, and L. Glazman, Phys. Rev.
601: Lett \textbf{88}, 186802 (2002)
602:
603: \bibitem{nuc9}J. Schliemann, A. V. Khaetskii, and D. Loss, Phys.
604: Rev. Lett \textbf{66}, 245303 (2002)
605:
606: \bibitem{nuc10}A. Khaetskii, D. Loss, and L. Glazman, Phys. Rev.
607: B \textbf{67}, 195329 (2003)
608:
609: \bibitem{nuc11}W. A. Coish and D. Loss, Phys. Rev. B \textbf{70},
610: 195340 (2004)
611:
612: \bibitem{nuc12}W. Zhang, V. V. Dobrovitski, K. A. Al-Hassanieh, E.
613: Dagotto, and B. N. Harmon, Phys. Rev. B \textbf{74}, 205313 (2006)
614:
615: \bibitem{cpl}G. Khitrova, H. M. Gibbs, M. Kira, S. W. Koch, and A.
616: Scherer, Nature Physics \textbf{2}, 81 (2006)
617: \end{thebibliography}
618:
619: \end{document}
620: