cond-mat0307393/PRA.tex
1: 
2: %\documentstyle[amsmath,amssymb,aps,twocolumn,epsfig,amstex]{revtex}
3: %\documentstyle[amssymb,aps,prl,preprint,epsfig,amstex]{revtex}
4: %\usepackage{epsfig}
5: \documentclass[twocolumn,showpacs,superscriptaddress,preprintnumbers,amsmath,amssymb]{revtex4}
6: \usepackage{graphicx} % Include figure files
7: \usepackage{dcolumn} % Align table columns on decimal point
8: \usepackage{bm}
9: \usepackage{color}
10: \usepackage{times}
11: \begin{document}
12: \color{black}
13: %\preprint{APS/123-QED}
14: %\usepackage{epsfig}
15: \title{Dispersive Manipulation of Paired Superconducting Qubits}
16: \author{Xingxiang Zhou}
17: \affiliation{Department of Electrical and Computer Engineering,
18: University of Rochester, Rochester, NY 14627}
19: \affiliation{Department of Physics and Astronomy, University of
20: Rochester, Rochester, NY 14627}
21: \author{Michael Wulf}
22: \email{mikewulf@pas.rochester.edu}
23: \affiliation{Department of
24: Physics and Astronomy, University of
25: Rochester, Rochester, NY 14627} %\affiliation{Department of
26: %Electrical and Computer Engineering, University of Rochester,
27: %Rochester, NY 14627}
28: \author{Zhengwei Zhou}
29: \affiliation{Key Laboratory of Quantum Information, University of
30: Science and Technology of China, Chinese Academy of Sciences,
31: Hefei, Anhui, 230026, China}
32: \author{Guangcan Guo}
33: \affiliation{Key Laboratory of Quantum Information, University of
34: Science and Technology of China, Chinese Academy of Sciences,
35: Hefei, Anhui, 230026, China}
36: \author{Marc J. Feldman}
37: \affiliation{Department of Electrical and Computer Engineering,
38: University of Rochester, Rochester, NY 14627}
39: 
40: %Xingxiang Zhou, Michael Wulf, Zhengwei Zhou, Guangcan Guo, Marc J. Feldman
41: \date{March 9th, 2003}
42: 
43: \begin{abstract}
44: We combine the ideas of qubit encoding and dispersive dynamics to
45: enable robust and easy quantum information processing (QIP) on
46: paired superconducting charge boxes sharing a common bias lead. We
47: establish a decoherence free subspace on these and introduce
48: universal gates by dispersive interaction with a LC resonator and
49: inductive couplings between the encoded qubits. These gates
50: preserve the code space and only require the established local
51: symmetry and the control of the voltage bias.
52: 
53: %To explore the potential for large scale quantum information
54: %processing (QIP) of superconducting qubits, we must overcome their
55: %severe decoherence and lack of an easy and efficient coupling
56: %scheme. In this work, we employ the ideas of qubit encoding and
57: %dispersive dynamics to realize robust and easy QIP on
58: %paired superconducting charge boxes sharing the same bias lead. %For this purpose, a pair of
59: %closely placed charge boxes sharing the same bias lead were used
60: %as the encoded qubit to avoid collective noise.
61: %The encoding serves two separate purposes, simplifications of
62: %control and reduction of decoherence, by encoding into a
63: %decoherence free subspace (DFS).
64: %
65: %A simple encoding scheme allows us to simplify the control needed
66: %and reduce the decoherence simultaneously, by taking advantage of
67: %the decoherence free subspace and virtual transitions, induced by
68: %inductive
69: %couplings, we %between the charge boxes and the charge box and LC resonators,
70: %implement universal QIP without the encoded qubits ever leaving
71: %the DFS.
72: %The only control required is tuning the voltage biases of the
73: %encoded qubits.
74: 
75: %A pair of closely placed charge boxes sharing the same bias were
76: %used as the encoded qubit to avoid collective noise. By coupling
77: %to LC resonators and making use of its virtual excitations we can
78: %enforce single encoded qubit manipulations. Coupling between
79: %encoded is realized by inductive coupling between the qubits.
80: 
81: %To be useful, a Quantum Computation scheme has to be able to
82: %perform sufficiently many operations before decoherence destroys
83: %the outcome to allow the use of quantum error codes. Decoherence
84: %free subspaces offer a solution to this problem if the coupling to
85: %the environment is coherent over multiple qubits. Here we propose
86: %a realization of this concept of passive error avoidance for the
87: %recently demonstrated charge qubits.
88: \end{abstract}
89: 
90: \pacs{03.67.Lx, 74.50.+r}
91: 
92: \maketitle
93: 
94: %\begin{section}{Introduction}
95: Superconducting nano-circuits consisting of charge boxes (CB)
96: \cite{Charge} are among the most promising candidates for a
97: quantum computer. Coherent control of a single charge qubit
98: \cite{Nakamura}, long decoherence time \cite{Vion} and, more
99: recently, coupled two qubit systems \cite{Pashkin02} have been
100: demonstrated. But despite this encouraging experimental progress,
101: there are serious difficulties with superconducting QIP which may
102: appear insurmountable. The first is the severe decoherence
103: experienced by these macroscopic qubits, which are coupled to a
104: large number of degrees of freedom in their environment including
105: control circuitry \cite{Leggett87}. A few methods have been
106: employed to reduce the decoherence \cite{ChargeEcho,Caspar}, but
107: they usually require sophisticated manipulation or significant
108: overhead in the control circuitry. The second major difficulty
109: comes from the imperfect control realizable in solid state
110: systems. Specifically, one finds it difficult to achieve
111: controllable couplings between superconducting qubits, since the
112: commonly used hard-wired inductive or capacitive couplings are
113: untunable. Great effort has been exercised to realize controllable
114: couplings \cite{Makhlin99,You02,AV}. Schemes allowing to compute
115: with invariable couplings were also studied
116: \cite{Zhou02,Benjamin02}. Others have recently discussed to use an
117: LC circuit to actively mediate the interaction between
118: superconducting qubits \cite{Falci02,Blais02}.
119: %These schemes unfortunately introduce
120: %sources of decoherence by the addition of hardware and
121: %control-lines. An example is the decoherence due to the finite
122: %quality of the LC-resonator.
123: 
124: The requirement to reduce decoherence and the desire for the
125: easiest manipulation apply to all QIP implementations.
126: %, though these are most manifested in superconducting QIP.
127: Unfortunately, it is not always easy to accomplish both - actually
128: the goals are often contradictory - since reducing decoherence may
129: require extra complication in the manipulation.
130: 
131: %In this work we show how to achieve both goals. Using a pair of
132: %closely placed CBs sharing a common bias lead as the logic qubit,
133: %we can encode information in a fashion immune to collective noise,
134: %which is the dominating decoherence source in our setting. We
135: %introduce LC resonators inductively coupled to the paired charge
136: %boxes (PCB) whose virtual excitations allow us to manipulate the
137: %PCB dispersively. By inductively coupling the CBs and taking
138: %advantage of dispersive dynamics again, controlled phases can be
139: %induced between logical qubits.
140: 
141: In this work we show how to achieve both goals. Using a closely
142: placed pair of charge boxes (PCB) sharing a common bias lead as
143: the logic qubit, we can encode information in a fashion immune to
144: collective noise, which is the dominating decoherence source in
145: our setting. We introduce LC resonators inductively coupled to the
146: PCBs whose virtual excitations allow us to manipulate the PCB
147: dispersively; all interactions will be off resonance, without
148: energy transfer \cite{Haroche, Scully}, and thus a logical qubit
149: stays within its encoding space even during manipulation. By
150: inductively coupling the CBs and taking advantage of dispersive
151: dynamics again, controlled phases can be induced between logical
152: qubits.
153: 
154: Combining dispersive dynamics and encoding offers a new method for
155: QIP. It overcomes the major difficulties of superconducting CBs in
156: a realistic and very simple fashion.
157: %Our entire scheme only relies on dispersive manipulations, and t
158: The only control required after initialization is to change the
159: voltage bias of the PCB.
160: % Our
161: % method also provides a new approach
162: %to QIP based on noise symmetry in short ranges, which is realistic
163: %even for scalable QIP.
164: %
165: Another advantage of our method is that it relies only on noise
166: symmetry over short distance and so this is a realistic technique
167: for scalable QIP over large systems.
168: %
169: Though we focus on superconducting QIP in this paper,
170: % the value of our work lies in
171: the general principle of dispersive manipulation of encoded qubits
172: will be valuable for many other QIP systems.
173: 
174: {\em Charge qubits and the dominating noise. --} In its simplest
175: form, the charge qubit is just a superconducting island voltage
176: biased through a Josephson junction. The Hamiltonian for the CB
177: system is $E_c(n-n_g)^2-E_J\cos\varphi$ \cite{Charge}, where the
178: charging energy $E_c=(2e)^2/2C_t$, $C_t$ being the total
179: capacitance of the island, is much greater than the Josephson
180: energy $E_J=I_c\Phi_0/2\pi$, and $\varphi$ is the phase drop
181: across the junction. When biased close to $n_g=C_gV_g/2e=1/2$
182: ($C_g$ is the gate capacitance), it provides an effective
183: two-state system which can be used as a qubit. In the spin 1/2
184: notation,
185: \begin{equation}
186: H_{CB}=\frac{1}{2}B^z\sigma^z-\frac{1}{2}B^x\sigma^x,
187: \label{eq:H_CB}
188: \end{equation}
189: where the spin up or down states correspond to n=0 or n=1 excess
190: Cooper-pairs on the CB. The effective field $B^z=E_c(2n_g-1)$ can
191: be tuned by changing the gate voltage $V_g$. In order to control
192: $B^x=E_J$, we use a flux biased small dc-SQUID to replace the
193: junction, whose critical current is maximal $(I_c=I_c^0)$ when the
194: flux bias is off and vanishes $(I_c=0)$ when it is
195: $\Phi_0/2=\frac{h}{4e}$.
196: 
197: The dominating decoherence sources in this system are circuit
198: noise in the voltage bias \cite{Charge} and charge fluctuations in
199: the background (known as ``charge noise'')
200: \cite{Zorin96,Paladino02}, as indicated in Fig. \ref{fig:Charge}
201: (a). The circuit noise is described by the well known
202: ``Spin-Boson'' model \cite{Leggett87}. The charge noise is less
203: well understood, but it is now generally believed to be caused by
204: fluctuations of the impurity charges in the substrate
205: \cite{Zorin96}. Taking into account the noise sources, we have the
206: total Hamiltonian for the system as $H=H_{CB}+H_Z+H_B+H_{int}$,
207: where $H_Z$ and $H_B$ are the Hamiltonian of the circuit
208: fluctuations, modelled as a collection of harmonic oscillators
209: \cite{Makhlin01}, and the
210: Hamiltonian of the background charge. %For our purpose the specific
211: %nature of $H_Z$ and $H_B$ is irrelevant.
212: Since these noise perturb the voltage bias of the CB, the
213: interaction Hamiltonian $H_{int}$ has the form
214: $H_{int}=\sigma^z\hat{E}^Z+\sigma^z\hat{E}^B$,
215: %\begin{equation}
216: %H_{int}=\sigma^z\sum_i\lambda_i^Z\hat{E}_i^Z+\sigma^z\sum_j\lambda_j^B\hat{E}_j^B,
217: %H_{int}=\sigma^z\hat{E}^Z+\sigma^z\hat{E}^B,
218: %\end{equation}
219: where $\hat{E}^Z$ and $\hat{E}^B$ are environment operators (with
220: the coupling strengths included).
221: %and $\lambda_i^Z$ and $\lambda_^B$ are the coupling strengths.
222: Here we focus on the fluctuations in the voltage bias, the
223: dominating source of decoherence and neglect noise in the $B^x$
224: field (the critical current), as practised customarily
225: \cite{Leggett87, Makhlin01, Paladino02}. This is because charge
226: qubits are insensitive to flux noise and the effect of the
227: fluctuation in the $I_c$ suppression field is secondary to the
228: bias voltage variations discussed above. Detailed treatment of the
229: two voltage noise sources and their influence on the charge qubit
230: system can be found in the literature \cite{Paladino02,Makhlin01}.
231: For our purpose, the nature of the environment and specific form
232: of $\hat{E}^Z$ and $\hat{E}^B$ are not essential, therefore we do
233: not go into detail here.
234: \begin{figure}[h]
235:     \centering
236: %    \epsfig{file=aa1.eps, width=2.3in, height=1.2in}
237: %    \includegraphics[width=2.5in, height=2in]{dfs_xx_fig1.eps}
238:     \includegraphics[width=3in, height=2in]{Charge1.eps}
239:     \caption{%(a) A charge qubit under the influence of the circuit
240:     %noise, represented by the impedance $Z$, and the background
241:     %charge fluctuation schematically represented by $B$.
242:     (a) A pair
243:     of closely spaced CBs sharing a common lead used as
244:     the encoded qubit. %The Josephson junction is replaced by a low
245:     %self-inductance dc-SQUID to realize tunable critical current.
246:     %Due to the shared bias and close location the two charge boxes
247:     %experience the same circuit and charge noise.
248:     The SQUIDS of both CBs are inductively coupled to an
249:     LC resonator whose virtual excitation allows the two charge
250:     boxes to exchange energy quanta. The circuit noise and background charge fluctuations are
251:     represented by $Z$ and $B$ respectively.
252:     (b) An inductively coupled PCB array. For
253:     clarity the PCBs are schematically represented by two boxes,
254:     the inductive coupling is understood to be between the
255:     dc-SQUIDS of the PCBs.
256:     }
257:     \label{fig:Charge}
258: \end{figure}
259: 
260: {\em Paired charge boxes and DFS encoding. --} As shown in Fig.
261: \ref{fig:Charge} (a), we use two capacitively coupled identical
262: CBs ($a$ and $b$) with a common bias-lead as an encoded qubit. The
263: small capacitive coupling, $C_c\ll C_t$, is not essential for the
264: encoding, but necessary for the encoded two qubit gates. The
265: Hamiltonian of the PCB system is $H_{PCB}=\sum_{i=a,b}
266: \frac{1}{2}(B^z\sigma_i^z-
267: B_i^x\sigma_i^x)-\gamma\sigma_a^z\sigma_b^z$, where
268: $\gamma=\frac{C_c}{2C_t}E_c$ is the coupling energy.
269: 
270: Since the two CBs share the same lead, obviously they are biased
271: at the same voltage and they will experience the same circuit
272: noise. In addition the nano-scale charge islands are put
273: very close to each other. %This proximity leads to
274: %the two CBs being affected by charge-background fluctuations
275: %indeed symmetrically as we will show later.
276: Therefore, they will experience the same charge fluctuations too
277: (more discussion on this point will be given later). Hence the CBs
278: experience ``collective decoherence,'' meaning the noise sources
279: couple symmetrically to them, which naturally gives rise to
280: ``decoherence-free encoding.'' For a review of decoherence-free
281: subspace, see \cite{Kempe01} and references therein. Here we have
282: the simplest case of the DFS, with
283: $|0\rangle=|\downarrow_a\uparrow_b\rangle$ and
284: $|1\rangle=|\uparrow_a\downarrow_b\rangle$ as the decoherence-free
285: logical states. The way this works can easily be seen: as a
286: consequence of the collective decoherence the coupling between the
287: PCB and the environment is
288: $(\sigma_a^z+\sigma_b^z)*(\hat{E}^B+\hat{E}^Z)$,
289: %$\hat{E}$ being the environment operator,
290:  which annihilates the
291: two logical states given above. Therefore the PCB system will not
292: get entangled with the environment if it is initialized and kept
293: in the DFS. Physically, the encoded qubit's immunity to noise
294: stems from the fact that the CBs acquire random but opposite
295: phases.
296: 
297: To prepare the system in the DFS, we bias the PCB far off the
298: degeneracy point $n_g=1/2$. In the spin-1/2 picture, this
299: corresponds to applying a strong field in the $z$ direction. At
300: low temperatures, the spins will line up with the field, and the
301: PCB relaxes to the state $|\downarrow_a\downarrow_b\rangle$.
302: Keeping both $B^x$ off, we change $B^z$ (same for $a$ and $b$) to
303: $-2\gamma$. This cancels the bias of $a$ on $b$ making its total
304: $B^z_{tot}=B^z+2\gamma=0$. After that we turn on some $B_b^x$.
305: After a time $\pi/B_b^x$, the state of $b$ will become
306: $|\uparrow_b\rangle$, and the system is prepared in the DFS state
307: $|\downarrow_a\uparrow_b\rangle$. Now we turn off $B_b^x$ as well,
308: and from now on the $B^x$ fields for both CBs will remain off, by
309: biasing the dc-SQUIDS of the PCB at $\Phi_0/2$. Since $B^x$ fields
310: will remain off and need not be tuned after the initialization,
311: the leads tuning $I_c$ of the PCB could be heavily filtered to
312: keep out the noise once the system is initialized. Alternatively
313: we could make use of the noise free constant flux-bias techniques
314: such as that demonstrated in \cite{MooijAPL}. In practice it can
315: be difficult to suppress $I_c$ of the PCB precisely to 0 due to
316: the finite self inductance of the dc-SQUID. However as shown in
317: \cite{You01} if low self inductance dc-SQUIDS are used
318: ($\frac{LI_c}{\Phi_0}\ll 1$) the $B^x$ field at $\Phi_0/2$ bias
319: point is negligibly small compared to $B^z$ fields used for
320: computation and can thus be safely dropped. Many schemes
321: \cite{Charge, Makhlin99, You02} rely on this fact too. One notices
322: that the logical states $|0\rangle$ and $|1\rangle$ are always
323: degenerate regardless of the voltage bias $n_g$, therefore there
324: is no evolution in the idle mode, regardless of the voltage bias
325: or noise in it. To readout the state of the PCB, a measurement of
326: its CB $a$ or $b$ will suffice, which is readily accomplished with
327: developed techniques \cite{Makhlin01}.
328: 
329: As seen above, in realizing DFS encoding with the PCB, we lose
330: considerable freedom in manipulating the system. First, in order
331: to guarantee symmetrical coupling to the circuit noise, the two
332: CBs share the same lead and hence they are always biased at the
333: same voltage. More importantly, we must ensure that operations on
334: the PCB do not drive the system out of the DFS, otherwise the
335: immunity to noise is lost \cite{Kempe01}. This is why we must keep
336: the $B^x$ fields for the CBs off: they flip the states of a single
337: CB and hence do not preserve the DFS. Therefore the only control
338: left is the voltage bias of the PCB, which clearly is insufficient
339: for universal QIP on the PCB. To deal with this difficulty, we
340: introduce a LC resonator inductively coupled to the PCB system:
341: 
342: {\em The LC resonator inductively coupled to the PCB. --} As shown
343: in Fig. \ref{fig:Charge} (a), we inductively couple the dc-SQUIDS
344: of the two CBs in the PCB system symmetrically to an initially
345: unexcited LC circuit. Even in the ground state its vacuum
346: fluctuations bias the dc-SQUIDS of the PCB off $\Phi_0/2$ making
347: charge tunnelling possible. The Hamiltonian for the PCB-LC system
348: is $H=H_{PCB}+H_{LC}+H_{coup}$, where $H_{LC}=\omega a^\dagger a$,
349: and $H_{coup}=-ig(a-a^{\dagger})\sigma_x$ with $
350: g=\frac{1}{2}MI_c^0\sqrt{\frac{\hbar\omega}{2L}}$.
351: %
352: %$H_{coup}=-\frac{i}{2}I_c^0M\sqrt{\frac{\hbar\omega}{2L}}
353: %(a-a^{\dagger})\sigma_x=-ig(a-a^{\dagger})\sigma_x$ is the
354: %coupling term.
355: %in its lowest order.
356: Here $\omega$ and $L$ refer to frequency and inductance of the
357: LC-resonator; M is the mutual inductance between SQUID and
358: resonator. When the PCB and LC are far off resonance, the effect
359: of the LC can be neglected. On the other hand, when we tune the
360: bias of the PCB such that it is close to being in resonance with
361: the LC resonator, within the framework of Rotating Wave
362: approximation the above Hamiltonian becomes
363: %
364: % the
365: %Jaynes-Cummings model
366:  $H=H_{PCB}+H_{LC}-ig
367: \sum_{i=a,b}(\sigma_i^+a-\sigma_i^-a^\dagger)$,
368: %\begin{equation}
369: %H_{JC}=H_{PCB}+H_{LC}-ig\sum_{i=a,b}(\sigma_i^+a-\sigma_i^-a^\dagger),
370: %\label{eq:H_JC}
371: %\end{equation}
372: where $\sigma^\pm=(\sigma^x\pm i\sigma^y)/2$ are the ladder
373: operators. Notice that $H_{PCB}$ contains a coupling term , in
374: contrast to the standard Jaynes Cummings Model which modifies the
375: dynamics significantly, as will be seen below.
376: 
377: %In the following we choose $B_z$ to be larger than 0.
378: Let $\delta=B^z-\omega$ be the detuning. If we let the PCB and the
379: (initially un-excited) LC resonator interact right in resonance,
380: i.e., $|\delta-2\gamma|\ll g\ll \omega$ or $|\delta+2\gamma|\ll
381: g\ll \omega$ \cite{Note}, state transfer occurs between the PCB
382: and the LC resonator \cite{Falci02, Blais02}. This is not allowed
383: in our scheme, since it will drive the PCB out of the DFS.
384: Besides, once the LC resonator is excited, we will have additional
385: decoherence due to the finite quality of the LC resonator
386: \cite{Falci02, Blais02}. Therefore, we only work in the dispersive
387: region, $g\ll |\delta\pm 2\gamma| \ll \omega$. In this case, the
388: PCB and the LC resonator cannot exchange energy because of the
389: large detuning. However the virtual excitation of the LC resonator
390: gives rise to an effective interaction between the two CBs
391: \cite{Zheng},
392: \begin{equation}
393: H_{eff}=\frac{g^2}{\delta+2\gamma} +
394: \frac{g^2}{\delta+2\gamma}(\sigma_a^+\sigma_b^-
395: +\sigma_a^-\sigma_b^+), \label{eq:H_eff}
396: \end{equation}
397: where the first term describes the Stark shift, which can be
398: neglected since it is the same for both logical states,
399: %due to the
400: %emission and absorption of the virtual photon by the same CB,
401: and the second term is the effective exchange interaction
402: %between the two CBs
403: caused by the exchange of a virtual photon. It preserves the DFS
404: (it changes $|\uparrow_a\downarrow_b\rangle$ to
405: $|\downarrow_a\uparrow_b\rangle$ and vice versa) and acts as a
406: logical $X$ gate on the PCB. Starting from
407: $|0\rangle=|\downarrow_a\uparrow_b\rangle$, letting the PCB evolve
408: under the effective Hamiltonian (\ref{eq:H_eff}) for a time
409: $t=\frac{\pi(\delta+2\gamma)}{g^2}$ or $\frac{t}{2}$, we can swap
410: the states of $a$ and $b$ or generate a maximally entangled state
411: between them.
412: %
413: %swaps the state of the two CBs $a$ and $b$. A maximally entangled
414: %state between $a$ and $b$ results if the system evolves for a time
415: %$t=\frac{\pi(\delta+2\gamma)}{2g^2}$. Also,
416: The LC resonator is initially in the vacuum state and will not be
417: excited due to the dispersive interaction with the PCB; therefore
418: unlike in previous schemes \cite{Falci02, Blais02}we are not
419: subject to decoherence caused by its finite quality. Also, notice
420: that our dispersive scheme is fundamentally different from the
421: previous method of using the virtual excitation of a large LC
422: circuit capacitively coupled to all the qubits \cite{Makhlin99},
423: in which the LC-frequency is much larger than the CB-energies and
424: the coupling strength is tuned by changing the $E_J$'s of the
425: qubits. In our scheme the two energies are close (though the
426: detuning is large) and the $E_J$'s are always off; the coupling
427: strength is controlled by simply changing the detuning. Our scheme
428: takes advantage of the quantum exchange effect (Eq. (2)) assisted
429: by a virtually excited quantum state of the LC-qubit system
430: \cite{Zheng,Haroche01}. It offers noise protection with simplest
431: operation using very realistic parameters (see below), which was
432: not available in previous schemes.
433: 
434: To realize $SU(2)$ on the PCB, we still need a phase gate. This
435: can be accomplished by using another LC resonator, not depicted in
436: Fig. \ref{fig:Charge} (a) and at a frequency $\omega'$ very
437: different from $\omega$, inductively coupled to the dc-SQUID of
438: only $a$ (or $b$) in the PCB. Then when we tune the voltage bias
439: of the PCB such that it interacts with this LC resonator
440: dispersively ($g\ll |\delta'\pm 2\gamma|\ll \omega'$), a phase
441: gate will be obtained due to the Stark shift
442: $\frac{g^2}{\delta+2\gamma}|1\rangle \langle 1|$. The effect of
443: the previous LC resonator can be neglected because it is far off
444: resonance with the PCB. However for the purpose of universal QIP
445: this %phase gate realized by a
446: second LC resonator is not absolutely necessary
447: \cite{Shi02}, %as long as we have a quantum information processor
448: %in systems of more than 1 qubit, because another single PCB gate
449: %not commuting with the $X$ gate can be obtained by making use of
450: %the $X$ gate and the controlled phase gate
451: as will be shown below.
452: 
453: {\em Inductively coupled PCB arrays. --} To realize universal QIP
454: on the PCBs, we need a scheme to couple them. We notice that
455: different PCBs will experience different noise as they are biased
456: by different leads. Those far apart are susceptible to different
457: charge noise too. So we only have what we call ``local'' DFS; this
458: only relies on noise symmetry over a few, here two, physical
459: qubits, as is inevitably the only realistic case for scalable QIP,
460: and previously discussed methods \cite{Kempe01} do not apply.
461: %
462: %
463: Stringent restrictions are put on the two qubit coupling in order
464: to preserve the DFS. A capacitive coupling between $1b$ and $2a$,
465: for example, cannot be used, as this would cause the noise in
466: PCB1's lead to leak asymmetrically into PCB2. Furthermore neither
467: PCB may leave its DFS during its evolution.
468: 
469: Here we discuss a new approach that allows scalable QIP based on
470: local DFS.
471: % , work that focuses on situations of higher
472: %noise-symmetry \cite{Kempe01} does not apply.
473: We couple the PCBs
474: inductively, using a small mutual inductance $M'$ between their
475: dc-SQUIDS, as shown in Fig. \ref{fig:Charge} (b).
476:  As the dc-SQUIDS are biased
477: at $\Phi_0/2$ the coupling Hamiltonian is
478: $\lambda\sigma_{1b}^x\sigma_{2a}^x$ \cite{You01}, where the
479: coupling strength $\lambda=\frac{3}{4}M'I_{c,1b}^0I_{c,2a}^0$
480: (chosen much smaller than $E_J^0$, the unsuppressed Josephson
481: Energy of the SQUIDS).
482: 
483: Obviously the PCBs in the array in Fig. \ref{fig:Charge} (b) can
484: be initialized in their DFSs just as described before. Now, if all
485: the PCBs are biased at the same voltage, they will get out of
486: their DFSs due to resonant interactions by the above coupling
487: term. Therefore, we bias the odd numbered PCBs at the degenerate
488: point $n_g=1/2$, giving a $B^z=0$ and the even numbered PCBs at
489: some other point (but far off resonance with their LC resonators)
490: giving a large $B^z$. Because of the large detuning, $B^z\gg
491: \lambda$, the coupling Hamiltonian has no effect in the idle mode
492: \cite{Benjamin02}. One might worry that this different biasing
493: will introduce a difference between the phase frequencies of the
494: PCBs, but it is not the case since the DFS states always have the
495: energy $\gamma$ regardless of the voltage bias of the PCBs. When
496: we want to do a controlled phase gate between PCB 1 and 2, we tune
497: their biases near some common target value very different from
498: their previous values (such that they do not interact with other
499: neighboring PCBs) and the LC frequencies. Assuming the fields are
500: $B_1^z$ and $B_2^z$, we work in the dispersive region such that
501: the states of the PCBs do not change except that dispersive phases
502: are obtained due to the virtual transitions. For instance, the
503: energy of the state $|0_10_2\rangle$ will be shifted by
504: $\frac{\langle
505: 0_10_2|\lambda\sigma_{1b}^x\sigma_{2a}^x|m\rangle^2}{E_{|0_10_2\rangle}
506: -E_{|m\rangle}}=\frac{\lambda^2}{4\gamma+\Delta}$, where
507: $|m\rangle=|\downarrow_{1a}\downarrow_{1b}\uparrow_{2a}\uparrow_{2b}\rangle$
508: is the virtually excited intermediate state and
509: $\Delta=B_1^z-B_2^z$ is the detuning. Other phases can be
510: calculated too, giving in the basis $|0_10_2\rangle$,
511: $|0_11_2\rangle$, $|1_10_2\rangle$, $|1_11_2\rangle$ an effective
512: Hamiltonian $diag\{\frac{\lambda^2}{4\gamma+\Delta},
513: \frac{\lambda^2}{4\gamma+(B_1^z+B_2^z)},
514: \frac{\lambda^2}{4\gamma-(B_1^z+B_2^z)},
515: \frac{\lambda^2}{4\gamma-\Delta}\}$, which reduces to $diag\{0, 0,
516: 0, \frac{\lambda^2}{4\gamma-\Delta}\}$ in the dispersive region
517: $\lambda\ll |4\gamma-\Delta|\ll |4\gamma+\Delta|\ll |4\gamma\pm
518: (B_1^z+B_2^z)|$. This gives a CPHASE($\alpha$) gate $diag\{1, 1,
519: 1, e^{-i\alpha}\}$, where
520: $\alpha=\frac{\lambda^2t}{4\gamma-\Delta}$ with $t$ being the
521: evolution time.
522: 
523: Notice that in the above implementation of the CPHASE gate, the
524: capacitive coupling $\gamma$ plays an essential role. As can be
525: easily checked the above procedure does not give an entangling
526: gate if $a$ and $b$ do not bias each other ($\gamma=0$). This is
527: because $|0_10_2\rangle$ and $|1_11_2\rangle$ would acquire
528: opposite energy shifts $\pm \lambda^2/\Delta$, due to the fact
529: that the energy differences between the initial and intermediate
530: states are opposite for these two cases. Thus the role of the
531: capacitive coupling $\gamma$ can be understood by an interesting
532: ``parity argument:'' under the exchange of the states of $a$ and
533: $b$ for both PCBs, the energy mismatch (the denominator in the
534: perturbative calculation) due to the detuning $\Delta$ is odd.
535: This symmetry is broken by the coupling between $a$ and $b$, which
536: is even under this operation (as is obvious from the form
537: $-\gamma\sigma_a^z\sigma_b^z$).
538: 
539: Another point of interest is that, a phase gate on a PCB can be
540: implemented by using the CPHASE($\alpha$) and the X gate, as is
541: easily recognized by the identity
542: $e^{iZ_1\alpha}=e^{i\alpha}(X_2\cdot CPHASE(2\alpha))^2$.
543: Therefore, for the purpose of universal QIP on the PCB array the
544: LC resonator giving the phase gate is not absolutely necessary, as
545: long as the system has more than 1 qubit \cite{Shi02}. This is
546: potentially beneficial in reducing the hardware.
547: 
548: {\em Discussion. --} In the above we described our scheme of QIP
549: with PCBs based on encoded qubits and dispersive dynamics, which
550: requires only tuning the voltage bias of the PCBs. Our scheme can
551: prevent decoherence from collective noise. The circuit noise
552: obviously couples symmetrically to the CBs of the PCB. The charge
553: noise requires some caution, since its exact nature is still a
554: topic of debate \cite{Song95,Zorin96}. The early experiments in
555: \cite{Zorin96} clearly show that the charge noise on close by
556: islands are correlated. The conclusion drawn from this
557: observation, that the charge noise stems mostly from sources in
558: the substrate was further substantiated by \cite{Krupenin00}. This
559: has important consequences, because it suggests that it is
560: possible to engineer the environment for desired noise
561: configurations. Indeed, analysis in \cite{Zorin96} shows that high
562: noise correlations can be achieved for properly designed geometry
563: and layouts of the charge islands. Simple environment engineering
564: was already successful \cite{Krupenin00, Clark03} in various
565: contexts.
566: 
567: 
568: % Experiments clearly showed
569: %\cite{Zorin96} that the charge noise of two closely spaced charge
570: %islands are correlated, supporting the conjecture that
571: %fluctuations coming from the substrate dominate the charge noise.
572: %\color{red} {\it The noise correlation observed in the relatively
573: %primitive experiment \cite{Zorin96} was substantial enough to
574: %identify the major source of the charge noise. Furthermore, the
575: %analysis in \cite{Zorin96} showed that the noise correlation can
576: %be close to unity for properly designed geometry and layouts of
577: %the charge qubits. Indeed, the noise correlation was used to
578: %distinguish between true and false signals for charge measurements
579: %[the australian group's APL paper on charge measurement]. }
580: For our scheme to work, the CBs must be located within a distance
581: smaller than the wavelength of the background charge fluctuations,
582: so that they experience the same noise. This seems to be
583: realistic, since the advance in device fabrication allows to make
584: smaller structures, and more importantly the fact that the charge
585: noise originates from the substrate makes it possible to engineer
586: the environment for the desired noise symmetry
587: \cite{Zorin96,Krupenin00}. For instance, if we put the PCB on an
588: electrode instead of the substrate \cite{Krupenin00}, the charge
589: impurities will be located far away from the PCB, and thus couple
590: symmetrically to the CBs.
591: %
592: %Though we expect our scheme to reduce greatly the effect of the
593: %collective noise on the PCB, obviously we will not have an
594: %infinitely long decoherence time as in ideal DFS QIP. There are
595: %other non-collective noise in the system not dealt with by our
596: %prescription, for instance dissipation due to the finite impedance
597: %of the junction and the noise in its critical current. The effect
598: %of the virtually excited states and the fluctuation of the
599: %dispersive energies must be evaluated carefully too, though some
600: %results were obtained previously \cite{Sorensen99,Gea02}. A
601: %detailed calculation of the decoherence time for a realistic PCB
602: %system is beyond the scope of this work and will be reported
603: %elsewhere.
604: %
605: 
606: Though our scheme eliminates the effect of the collective charge
607: noise on the PCB, the decoherence time will be finite as there are
608: other non-collective noise in the system not dealt with by our
609: prescription, for instance dissipation due to the finite impedance
610: of the junction and the noise in its critical current. The effect
611: of the virtually excited states and the fluctuation of the
612: dispersive energies must be evaluated carefully too, though some
613: results were obtained previously \cite{Sorensen99,Gea02}.
614: Qualitatively, as shown by simple analysis based on Master
615: equations the number of operations allowed in our dispersive
616: scheme increases by $\Delta/g \gg 1$ (here, $\Delta$ the effective
617: detuning and g the coupling strength) as compared to the usual
618: scheme based on resonant Rabi manipulations, if the same coupling
619: strength $g$ is assumed \cite{Mike}. The PCBs do not experience
620: decoherence in the idle mode, which is favorable for a large
621: system in which only a fraction of the qubits undergo active
622: manipulation at the same time. Therefore, our scheme we can reduce
623: the error rate of the PCBs below the threshold for error
624: correction schemes \cite{QCbook} and thus make superconducting QIP
625: feasible. Detailed calculation of the decoherence time for a
626: realistic PCB system will be reported elsewhere \cite{Mike}.
627: 
628: Another practical concern is that the CBs in a PCB will not be
629: completely identical due to the imperfect fabrication. Because
630: only local symmetry is required, this problem is less significant
631: since fabrication variations tend to happen at large scales and
632: experiments show that closely spaced charge-boxes can be equal
633: beyond experimental resolution \cite{Zorin96}. In addition, the
634: error induced by non-identical qubits was shown to be higher order
635: in the symmetry breaking \cite{Lloyd}. Therefore, we conclude that
636: the technological problem of imperfect fabrication is already
637: solved to the extent needed for our scheme.
638: 
639: {\em Parameters. --} Finally, we give some parameters for the
640: experimental consideration. We use small CBs closely spaced with a
641: total capacitance $C_t\approx 0.16fF$ and charging energy
642: $E_c\approx 500GHz$. A mutual capacitance $C_c=5aF$ gives
643: $\gamma\approx 7.5GHz$. A mutual inductance $M=7pH$ between the
644: PCB and LC with $L=50pH$ and $\omega/2\pi=200GHz$ gives
645: $g=0.25GHz$ for $I_c=40nA$. Tuning the bias of the PCB close to
646: the LC frequency with a detuning $\delta\approx -12.5GHz$ results
647: in an exchange interaction with the strength
648: $\frac{g^2}{\delta+2\gamma}\approx 25MHz$, corresponding to a
649: period of $40ns$.
650: % Alternatively a large current biased junction
651: %could be used to realize the resonator.
652: During the idle mode, we
653: bias the odd numbered PCBs at $B^z=0$ and the even numbered ones
654: at $B^z\approx 100GHz$. Since low self inductance dc-SQUIDS should
655: be used, we choose $M'I_c\approx 10^{-3}\Phi_0$ \cite{You01}, and
656: $M'=120pH$ gives the coupling strength $\lambda \approx 0.22GHz$.
657: Tuning the biases of the neighboring PCBs both to about $400GHz$
658: with a detuning $\Delta\approx 28GHz$ gives a CPHASE gate at the
659: rate $\frac{\lambda^2}{4\gamma-\Delta}\approx 25MHz$. The above
660: parameters are well within the reach of the current technology
661: \cite{Jena}.
662: 
663: In conclusion, we have discussed a technique for robust and easy
664: superconducting QIP. By combining the ideas of encoding and
665: dispersive manipulations, we protect the charge qubits from the
666: dominating decoherence and realize universal QIP on the encoded
667: qubits with minimal control. Besides the great potential of
668: solving the fundamental difficulties in superconducting QIP, we
669: expect the general idea of dispersive manipulation of encoded
670: qubits to be of interest to other physical systems, such as atomic
671: and other solid state systems \cite{Others}.
672: 
673: %The authors acknowledge support from AFOSR, ARDA, and NFRP, NNSF
674: %and CAS of China.
675: 
676: X. Zhou, M. Wulf and M. J. Feldman were supported in part by AFOSR
677: and funded under the DoD DURINT program and by the ARDA. Z-W. Zhou
678: and G-C. Guo acknowledge funds from NFRP, NNSF, and CAS of China.
679: %
680: %X. Zhou, M. Wulf and M. J. Feldman were supported in part by AFOSR
681: %and funded under the DoD DURINT program and by the ARDA. Z-W. Zhou
682: %and G-C. Guo acknowledge funds from National Fundamental Research
683: %Program %(2001CB309300)
684: %, National Natural Science Foundation of China, and Chinese
685: %Academy of Sciences.
686: %note: the choice for g=0.5GHz requires something like:
687: %$E_j^0=24GHz$; M=L/20; L=10pH; $2\pi*\omega_{LC}=400GHz$ this
688: %implies a capacitance of  16 fF. This seems fairly reasonable,
689: %even for a LC-oscillator
690: 
691: \begin{references}
692: \bibitem{Charge}A. Shnirman, G. Sch\"{o}n and Z. Hermon,
693: Phys. Rev. Lett. {\bf 79}, 2371 (1997).
694: \bibitem{Nakamura} Y. Nakamura, Y. A. Pashkin,
695: J. S. Tsai, Nature {\bf398}, 786 (1999).
696: \bibitem{Vion}
697: D. Vion, A. Aassime, A. Cottet, P. Joyez, H. Pothier, C. Urbina,
698: D. Esteve, M. H. Devoret, Science {\bf 296}, 886 (2002).
699: \bibitem{Pashkin02}
700: Yu. A. Pashkin, T. Yamamoto, O. Astafiev, Y. Nakamura, D. V.
701: Averin, J. S. Tsai , Nature {\bf421}, 823 (2003).
702: %\bibitem{Universal}A. Barenco {\it et al.}, Phys. Rev. A {\bf 52}, 3457 (1995).
703: %\bibitem{Leggett86}A. J. Leggett, {\em Quantum Mechanics at the Macroscopic
704: %Level,} in ``Directions in Condensed Matter Physics,'' vol 1,
705: %World Scientific, Singapore, 1986.
706: \bibitem{Leggett87}A. J. Leggett, S. Chakravarty, A. T. Dorsey,
707: Matthew P. A. Fisher, Anupam Garg, W. Zwerger, Rev. Mod. Phys.
708: {\bf 59}, 1, (1987).
709: \bibitem{ChargeEcho} Y. Nakamura, Y. Pashkin, T. Yamamoto,
710: J. S. Tsai, Phys. Rev. Lett. {\bf 88}, 047901 (2002).
711: \bibitem{Caspar} Caspar H. van der Wal, F.K. Wilhelm,
712: C. J. P. M. Harmans, J. E. Mooij, Eur. Phys. J. B,
713: {\bf 31}, 111 (2003).
714: %\bibitem{Spintronics}D. P. Divincenzo {\it et al.}, Nature {\bf
715: %   408},339 (2000). D. A. Lidar and L-A. Wu, Phys. Rev. Lett.
716: %    {\bf 88}, 017905 (2002).
717: \bibitem{Makhlin99}Y. Makhlin, G. Sch\"{o}n and A. Shnirman,
718: Nature {\bf 398}, 305 (1999).
719: \bibitem{You02}J. Q. You, J. S. Tsai and F. Nori, Phys. Rev.
720: Lett. {\bf 89}, 197902, (2002).
721: \bibitem{AV} D.V. Averin, C. Bruder, cond-mat/0304166
722: \bibitem{Zhou02}X. Zhou, Z-W. Zhou, G-C. Guo and M. J. Feldman,
723: Phys. Rev. Lett. {\bf 89}, 197903, (2002).
724: \bibitem{Benjamin02}S. C. Benjamin, S. Bose, quant-ph/0210157.
725: \bibitem{Falci02}F. Plastina and G. Falci, cond-mat/0206586.
726: \bibitem{Blais02}A. Blais, A. M. van den Brink, A. M. Zagoskin
727: , Phys. Rev. Lett. {\bf90}, 127901 (2003).
728: \bibitem{Haroche} M. Brune, S. Haroche, J.M. Raimond, L.
729: Davidovich, N. Zagury, Phys. Rev. A {\bf45}, 5139 (1992).
730: \bibitem{Scully} M.O. Scully, M.S. Zubairy, \textit{Quantum
731: Optics}, Cambridge University Press, 1997.
732: \bibitem{Zorin96} A. B. Zorin, F.–J. Ahlers, J. Niemeyer, T. Weimann,
733: H. Wolf, V. A. Krupenin, S. V. Lotkhov, Phys. Rev. B {\bf 53},
734: 13682 (1996).
735: \bibitem{Paladino02} E. Paladino, L. Faoro, G. Falci, and R. Fazio
736: , Phys. Rev. Lett. {\bf 88}, 228304 (2002).
737: \bibitem{Makhlin01}Y. Makhlin, G. Sch\"{o}n and A. Shnirman,
738: Rev. Mod. Phys. {\bf 73}, 357 (2001).
739: %\bibitem{DFS}L-M. Duan and G-C. Guo, Phys. Rev. Lett. {\bf
740:  %   79}, 1953 (1997). P. Zanardi and M. Rasetti, Phys. Rev. Lett. {\bf
741:  %   79}, 3306 (1997). D. A. Lidar, I. L. Chuang and K. B. Whaley, Phys. Rev. Lett.
742:  %  {\bf 81}, 2594 (1998).
743: \bibitem{Kempe01}J. Kempe, D. Bacon, D. A. Lidar and K. B. Whaley,
744: Phys. Rev. A {\bf 63}, 042307 (2001).
745: \bibitem{MooijAPL} J. B. Majer, J. R. Butcher, and J. E. Mooij, Appl. Phys. Lett. {\bf80}, 3638
746: (2002).
747: \bibitem{You01}J-Q. You, C-H. Lam and H-Z. Zheng,
748: Phys. Rev. B {\bf 63}, 180501(R) (2000).
749: \bibitem{Note} These two conditions correspond to transitions to
750: $|\uparrow_a\uparrow_b\rangle$ and
751: $|\downarrow_a\downarrow_b\rangle$ respectively. Note here the
752: coupling energy $\gamma$ must be taken into account in determining
753: the in-resonance condition.
754: \bibitem{Zheng}S-B. Zheng and G-C. Guo,
755: Phys. Rev. Lett. {\bf 85}, 2392 (2000). S-B. Zheng, Phys. Rev.
756: Lett. {\bf 87}, 230404 (2001).
757: 
758: \bibitem{Haroche01} S.Osnaghi, P. Bertet, A. Auffeves, P. Maioli, M. Brune,
759:  J. M. Raimond, S. Haroche, Phys. Rev. Lett
760: {\bf87}, 037902 (2001).
761: \bibitem{Shi02}Y-Y. Shi, Quantum Information and Computation, Vol.
762: {\bf3}, No. {\bf1}, 84 (2003).
763: %\bibitem{Bacon00}D. Bacon, J. Kempe, D. A. Lidar and K.B. Whaley,
764: %Phys. Rev. Lett. {\bf 85}, 1758 (2000).
765: \bibitem{Song95}D. Song, A. Amar, C.J. Lobb, F.C. Wellstood,
766: IEEE Trans. Appl. Supercond. {\bf 5}, 3058 (1995).
767: \bibitem{Krupenin00}V. A. Krupenin, D. E. Presnov, A. B. Zorin,
768: J. Niemeyer, J. Low Temp. Phys. {\bf 118}, 287 (2000).
769: \bibitem{Clark03} T.M. Buehler, D.J. Reilly, R. Brenner, A.R.
770: Hamilton, A.S. Dzurak, R.G. Clark, Appl. Phys. Lett. {\bf82} 577,
771: (2003).
772: \bibitem{Sorensen99}A. So\!\!\!\slash rensen and K. Mo\!\!\!\slash lmer,
773: Phys. Rev. Lett. {\bf 82}, 1971 (1999).
774: \bibitem{Gea02}J. Gea-Banacloche,
775: Phys. Rev. A {\bf 65}, 022308 (2002).
776: \bibitem{Mike} M. Wulf et.al, in preparation.
777: \bibitem{QCbook} M.A. Nielsen, I.L. Chuang, \textit{Quantum
778: Computation and Quantum Information}, Cambridge University Press,
779: 2000.
780: \bibitem{Lloyd} D. Bacon, D.A. Lidar, K.B. Whaley, Phys. Rev. A {\bf60},
781:  1944 (1999), S. Lloyd, unpublished.
782: \bibitem{Jena}D. Born, T. Wagner, W. Krech, U. H\"{u}bner,
783: L. Fritzsch, IEEE Trans. on Appl. Supercond. {\bf11}, 373 (2001).
784: \bibitem{Others}X-X. Yi, X-H. Su and L. You, Phys. Rev. Lett. {\bf90}, 097902 (2003). D.
785: Petrosyan and G. Kurizki, Phys. Rev. Lett. {\bf89} 207902 (2002).
786: 
787: 
788: %\bibitem{ref:BBC+95} C.H. Bennet, G. Brassard, R. Cleve, D.P.
789: %DiVincenzo, N. Margolus, P. Shor, T. Sleator, J. Smolin, H.
790: %Weinfurter, Phys. Rev. A {\bf 52}, 3457 (1995)
791: %\bibitem{ref:haroche} M. Brune, S. Haroche, V. Lefevre, J.M.
792: %Raimond, N. Zagury, Phys. Rev. Lett. {\bf 65}, 976 (1990)
793: %\bibitem{ref:finiteL} J.Q. You, C. Lam, H.Z. Zheng, Phys. Rev. B
794: %{\bf 63}, 180501R (2001)
795: %\bibitem{ref:lidar-nov02} D.A. Lidar, L.A. Wu, quant-ph/0211088v1
796: %(2002)
797: \end{references}
798: 
799: 
800: \end{document}
801: