1: %% LyX 1.3 created this file. For more info, see http://www.lyx.org/.
2: %% Do not edit unless you really know what you are doing.
3: \documentclass[a4paper,twoside,twocolumn,english]{revtex4}
4: \usepackage[T1]{fontenc}
5: \usepackage[latin1]{inputenc}
6: \usepackage{amsmath}
7: \usepackage{graphicx}
8: \usepackage{amssymb}
9:
10: \makeatletter
11: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% User specified LaTeX commands.
12: \newcommand{\ket}[1]{| #1 \rangle}
13: \newcommand{\bra}[1]{\langle #1 |}
14: \newcommand{\braket}[2]{\langle #1 | #2 \rangle}
15:
16: \usepackage{babel}
17: \makeatother
18: \begin{document}
19:
20: \title{Fast simulation of a quantum phase transition in an ion-trap realisable unitary map}
21:
22:
23: \author{J.P. Barjaktarevic, G.J. Milburn and Ross H. McKenzie}
24:
25:
26: \address{Quantum Computer Technology Research Centre,\\
27: Department of Physics, The University of Queensland,QLD 4072 Australia.}
28:
29:
30: \date{\today{}}
31:
32: \begin{abstract}
33: We demonstrate a method of exploring the quantum critical point of
34: the Ising universality class using unitary maps that have recently
35: been demonstrated in ion trap quantum gates. We reverse the idea with
36: which Feynman conceived quantum computing, and ask whether a realisable
37: simulation corresponds to a physical system. We proceed to show that
38: a specific simulation (a unitary map) is physically equivalent to
39: a Hamiltonian that belongs to the same universality class as the transverse
40: Ising Hamiltonian. We present experimental signatures, and numerical
41: simulation for these in the six-qubit case.
42:
43: \pacs{42.50.Vk,03.67.Lx,05.50.+q,73.43.Nq}
44: \end{abstract}
45: \maketitle
46:
47: \section{\label{sec1}Introduction}
48:
49: Feynman suggested that it is possible to simulate one quantum system
50: with another\cite{Feynman}. However, we will turn this thesis around
51: by posing the question of what sort of a system some unitary map on
52: a quantum computer might correspond to.
53:
54: In particular, we examine the ion-trap model of quantum computing,
55: and find that the unitary maps which have been realised on these correspond
56: to the time evolution of Hamiltonians which are linked closely to
57: the Ising model. Finally, we consider the considerable theoretical
58: body of work concerned with quantum phase transitions and renormalization
59: group theory. This will later be the key to the problem of identifying
60: a quantum phase transition in a unitary map.
61:
62:
63: \subsection{Simulating Quantum Systems}
64:
65: Feynman's first conception of quantum computing\cite{Feynman} held
66: the simulation of quantum systems as a key goal. Entanglement has
67: been described as the quintessential feature of quantum mechanics\cite{Schrodinger}.
68: In general, the arbitrary time evolution of a system is considered
69: an \emph{NP-hard\cite{NP}} problem, as memory and processing resources
70: increase exponentially in the size of the problem, $n$, on a classical
71: computer. It is only for extraordinarily simple systems, or ones for
72: which there are strong symmetries, that such calculations are tractable.
73:
74: Feynman suggested that the problem could be reduced to one in polynomial
75: time on a computer based on quantum principles. These include the
76: ability of a quantum system to perform unitary operations on a set
77: of quantum bits (\emph{qubits}), and to exist in entangled states.
78: Feynman showed that, in principle, it was possible to perform, in
79: polynomial time, algorithms which were only possible in non-polynomial
80: time on a classical computer.
81:
82: Since the original formulation of the problem, the application of
83: quantum computing to classical problems has become more common. Several
84: algorithms have been suggested, including the Deutch-Jozsa algorithm\cite{DJ},
85: Shor's factorization algorithm\cite{Shor}, and Grover's searching
86: algorithm\cite{Grover}. However, all of these systems are widely
87: considered far removed from current experimental abilities.
88:
89: Recently, Lloyd\cite{Lloyd} revisited Feynman's original problem,
90: and showed that it was possible to implement the time evolution of
91: an arbitrary spin Hamiltonian to a particular precision, $\varepsilon$,
92: in polynomial time. The procedure essentially involves the decomposition
93: of a Hamiltonian into realizable (local) unitary operations, and the
94: time-wise stepping through a Hamiltonian to some arbitrary accuracy.
95:
96: It will be our desire to avoid such an abstracted simulation of a
97: quantum system, and rather consider the possibility of finding a quantum
98: phase transition in a quantum algorithm naturally realizable with
99: current quantum computing experimental hardware. In this way, we will
100: essentially reverse the Feynman thesis, and conclude that quantum
101: algorithms (or unitary maps) will correspond to the observables of
102: some physical system.
103:
104:
105: \subsection{Ion Trap Quantum Computers and the Ising Model}
106:
107: DiVincenzo\cite{Universal} and Barenco \emph{et al.}\cite{Universal2}
108: has shown that single-site rotations and two-site controlled NOTs
109: are universal for quantum computation. Further, the Sorensen-Molmer\cite{smgate},
110: phase gate\cite{Wineland}, and indeed almost any two-site entangling
111: gate\cite{Nielsen} are universal. Hence, they will be able to affect
112: any unitary transformation.
113:
114: Cirac and Zoller's paper\cite{Ion} on cold ion-trap quantum computers
115: introduces the use of a spatially confined ion spin as a qubit, and
116: the excitation of vibrational modes as a means of coupling qubits.
117: Further, it has been shown that high fidelity state-preparation\cite{hifiprep}
118: and readout\cite{hifireadout} are feasible.
119:
120: Milburn has suggested a robust phase space scheme to use ion traps
121: to simulate nonlinear interactions in spin systems\cite{milburn}.
122: A significant advantage of this scheme is that it does not require
123: the cooling of vibrational states. The method involves the application
124: of Raman pulses faster than the vibrational heating time, effectively
125: decoupling the effect of vibrational modes. In particular, Milburn
126: shows that the evolution of a Hamiltonian of the form
127:
128: \begin{equation}
129: H_{int}=\hbar\chi\sigma_{z}^{(1)}\sigma_{z}^{(2)}\label{eq:}\end{equation}
130: may be achieved by a pulse sequence
131:
132: \begin{eqnarray}
133: U_{int}=e^{-iH_{int}} & = & e^{i\kappa_{x}\hat{X}\sigma_{z}^{(1)}}e^{i\kappa_{p}\hat{P}\sigma_{z}^{(2)}}\label{eq:}\\
134: & & e^{-i\kappa_{x}\hat{X}\sigma_{z}^{(1)}}e^{-i\kappa_{p}\hat{P}\sigma_{z}^{(2)}}\nonumber \end{eqnarray}
135: where $\hat{X}=\frac{a+a^{\dagger}}{\sqrt{2}}$ and $\hat{P}=\frac{a-a^{\dagger}}{i\sqrt{2}}$,
136: and expressions for $\kappa_{x}$ and $\kappa_{p}$ given in Ref.
137: \cite{warm_milburn}.
138:
139: Further, Wineland's research group have recently demonstrated\cite{Wineland}
140: considerable success in achieving few-qubit interactions with this
141: scheme. In particular, they present a two qubit phase gate, which
142: has the form
143:
144: \begin{eqnarray*}
145: \left|\downarrow\downarrow\right\rangle \rightarrow\left|\downarrow\downarrow\right\rangle & , & \left|\uparrow\uparrow\right\rangle \rightarrow\left|\uparrow\uparrow\right\rangle ,\\
146: \left|\downarrow\uparrow\right\rangle \rightarrow e^{i\phi}\left|\downarrow\uparrow\right\rangle & . & \left|\uparrow\downarrow\right\rangle \rightarrow e^{i\phi}\left|\uparrow\downarrow\right\rangle \end{eqnarray*}
147: which can be recast as $\left|\Psi\right\rangle \rightarrow e^{-i\chi\sigma_{x}^{(1)}\sigma_{x}^{(2)}}\left|\Psi\right\rangle $.
148: Apart from an uninteresting global additive phase, this may be considered
149: to model the time evolution of a Hamiltonian of the form $\sigma_{x}^{(n)}\sigma_{x}^{(n+1)}$.
150:
151: Further, it is well known that single rotations in any basis, which
152: correspond to the evolution of a spin operator, are easily implementable
153: on such an architecture\cite{Ion}. They result in unitary transformations
154: of the form
155:
156: \begin{equation}
157: U_{single}=e^{-i\hbar\theta\sigma_{x}^{(1)}}\label{eq:}\end{equation}
158: which can be implemented trivially through a single Raman pulse.
159:
160: Following Feynman's original intentions for quantum computing, one
161: may consider the mapping of Hamiltonian with such terms onto an ion-trap
162: quantum computer. Turning this problem around, we will consider the
163: properties of a unitary map composed of terms which can be experimentally
164: implemented, and investigate their relationship with the transverse
165: Ising spin chain.
166:
167:
168: \subsection{Quantum Phase Transitions and Universality Classes}
169:
170: The quantum phase transition in the one dimensional transverse Ising
171: model\cite{Sachdev} is very well understood. The Hamiltonian is given
172: by:
173:
174: \begin{equation}
175: H_{Ising}=\sum_{n=1}^{N}\mu B\sigma_{x}^{(n)}+J\sigma_{z}^{(n)}\sigma_{z}^{(n)}\label{ising}\end{equation}
176:
177:
178: It is known that for an external field with interaction strength \textbf{$\mu B$}
179: and local exchange interaction term with strength $J$, that a phase
180: transition occurs for $\mu B=\pm J$. One can intuitively consider
181: the phase transition as a result of the incongruent symmetries between
182: the two phases, which is reflected in the difference in behaviour
183: of the two terms in the Hamiltonian under the transformation $\sigma_{n}\rightarrow-\sigma_{n}$.
184: In the regime $J>\mu B$, the system is in a ferromagnetic phase,
185: with $\left\langle \sigma_{x}^{(n)}\right\rangle \neq0$, and the
186: system displays long range order. On the other hand, for $J<\mu B$,
187: the system is paramagnetic, with$\left\langle \sigma_{x}^{(n)}\right\rangle =0$,
188: and there is no broken symmetry.
189:
190: Using arguments from renormalization group theory, we may place a
191: great number of related problems into the same universality class\cite{Cardy},
192: and we may expect to see a similar phase transition occur in a number
193: of related systems.
194:
195:
196: \section{The Model}
197:
198: In the following, we will put together the components introduced in
199: Section \ref{sec1} in a intuitive way. We consider the composition
200: of the two unitary maps, similar to those demonstrated in Ref \cite{Wineland},
201: which corresponds to the composition of the time evolution of two
202: Hamiltonians. In form, it will look similar to the one-dimensional
203: transverse Ising chain Hamiltonian. We will then apply the Jordan-Wigner
204: transformation to this model to express the Hamiltonians in terms
205: of non-interacting fermions. We are then able to perform a composition
206: of operators in an $SU(2)$ representation to yield a single Hamiltonian.
207: We will find that this model is highly non-local. However, using renormalization
208: group theory concepts, it can be shown that the Hamiltonian belongs
209: in the same universality class as the transverse Ising chain. Hence,
210: we conclude that our separated model has the same quantum phase transition
211: as the transverse Ising chain, even though we have implemented the
212: map in a much simpler way.
213:
214:
215: \subsection{Model Unitary Transformation and Experimental Realization}
216:
217: It is natural to decompose the Ising Hamiltonian, $H_{Ising}$ into
218: two distinct parts:
219:
220: \begin{equation}
221: H_{\chi}=\chi\sum_{n=1}^{N}\sigma_{z}^{(n)}\sigma_{z}^{(n+1)}\label{eq:}\end{equation}
222:
223:
224: \begin{equation}
225: H_{\theta}=\theta\sum_{n=1}^{N}\sigma_{x}^{(n)}\label{eq:}\end{equation}
226:
227:
228: These parts are of even and odd symmetry under $\vec{\sigma_{n}}\rightarrow-\vec{\sigma_{n}}$,
229: respectively. Unitary maps of the form $\left|\Psi\right\rangle \rightarrow e^{iH_{\chi,\theta}}\left|\Psi\right\rangle $
230: have been realised experimentally. It is impossible to perform them
231: both at the same time with only single qubit rotations and two qubit
232: gates, because they do not commute - the evolution of the combined
233: Hamiltonian is not the composition of the evolutions of both Hamiltonians.
234: Note however that terms $\sigma_{z}^{(n)}\sigma_{z}^{(n+1)}$ and
235: $\sigma_{z}^{(m)}\sigma_{z}^{(m+1)}$ do commute, and so
236:
237: \begin{equation}
238: e^{iH_{\chi}}=\prod_{n=1}^{N}e^{i\chi\sigma_{z}^{(n)}\sigma_{z}^{(n+1)}}\label{eq:}\end{equation}
239: is realisable in principle with current technology.
240:
241: The combined Hamiltonian may be approximated by Lloyd's\cite{Lloyd}
242: methods, which involves applying terms such as $\frac{1}{m}H_{\chi}$
243: and $\frac{1}{m}H_{\theta}$ repeatedly, $m$ times. However this
244: requires a large overhead - instead we will consider the unitary map
245:
246: \begin{equation}
247: U(\chi,\theta)=e^{-iH_{\chi}}e^{-iH_{\theta}}=e^{-i\bar{H}}\neq e^{-i(H_{\chi}+H_{\theta})}\label{u_def}\end{equation}
248:
249:
250: This map has been proposed by Milburn \emph{et al.} \cite{warm_milburn}
251: as an easier unitary map to to simulate than the map which corresponds
252: to the time evolution of transverse Ising chain Hamiltonian. We are
253: interested in whether this mapping will have the same quantum phase
254: transition behaviour as the transverse Ising chain.
255:
256:
257: \subsection{Jordan-Wigner Transformation}
258:
259: We will follow Jordan and Wigner\cite{JW} in using the following
260: definitions to introduce a new set of operators, $a_{n}$, where
261:
262: \begin{eqnarray}
263: \sigma_{x}^{(n)} & = & 1-2a_{n}a_{n}^{\dagger}\label{eq:}\\
264: \sigma_{y}^{(n)} & = & -i(a_{n}-a_{n}^{\dagger})\label{eq:}\\
265: \sigma_{z}^{(n)} & = & a_{n}^{\dagger}+a_{n}\label{eq:}\end{eqnarray}
266: where $\sigma_{x}^{(n)}$, $\sigma_{y}^{(n)}$ and $\sigma_{z}^{(n)}$
267: take the form of the Pauli spin matrices in the $\left|0\right\rangle $,$a_{n}^{\dagger}\left|0\right\rangle $
268: basis. From these definitions, the operators $a_{n}$ and $a_{n}^{\dagger}$
269: can be shown to obey the following relations:
270:
271: \begin{eqnarray*}
272: \{ a_{n}^{\dagger},a_{n}\}=1, & a_{n}^{2}=0, & a_{n}^{\dagger^{2}}=0,\\
273: {}[a_{m}^{\dagger},a_{n}]=0, & [a_{m}^{\dagger},a_{n}^{\dagger}]=0, & [a_{m},a_{n}]=0,m\ne n\end{eqnarray*}
274:
275:
276: With these definitions, our unitary map becomes
277:
278: \begin{eqnarray}
279: U(\chi,\theta) & = & e^{-i\chi\sum_{n=1}^{N}a_{n}^{\dagger}a_{n+1}^{\dagger}+a_{n}a_{n+1}+a_{n}a_{n+1}^{\dagger}+a_{n}^{\dagger}a_{n+1}}\label{u_def_c}\\
280: & & e^{-i\theta\sum_{n=1}^{N}1-2a_{n}a_{n}^{\dagger}}\nonumber \end{eqnarray}
281:
282:
283: We then introduce the following operators
284:
285: \begin{eqnarray}
286: c_{n} & = & e^{i\pi\sum_{j=1}^{n-1}a_{j}^{\dagger}a_{j}}a_{n}\label{eq:}\\
287: c_{n}^{\dagger} & = & a_{n}^{\dagger}e^{-i\pi\sum_{j=1}^{n-1}a_{j}^{\dagger}a_{j}}\label{eq:}\end{eqnarray}
288:
289:
290: It can be shown that they obey fermionic anti-commutation relations.
291:
292: We may understand these as an expression of domain wall creation and
293: destruction. We can re-express $U(\chi,\theta)$ with this new set
294: of operators as
295:
296: \begin{eqnarray}
297: U(\chi,\theta) & = & e^{-i\chi\sum_{n=1}^{N}c_{n}^{\dagger}c_{n+1}^{\dagger}-c_{n}c_{n+1}-c_{n}c_{n+1}^{\dagger}+c_{n}^{\dagger}c_{n+1}}\label{eq:}\\
298: & & e^{-i\theta\sum_{n=1}^{N}c_{n}^{\dagger}c_{n}-c_{n}c_{n}^{\dagger}}\nonumber \end{eqnarray}
299:
300:
301: Finally, we will define the Fourier transformed versions of the fermion
302: operators as
303:
304: \begin{eqnarray}
305: c_{n} & = & \frac{1}{\sqrt{N}}\sum_{k}C_{k}e^{ink}\label{eq:}\\
306: c_{n}^{\dagger} & = & \frac{1}{\sqrt{N}}\sum_{k}C_{k}^{\dagger}e^{-ink}\label{eq:}\end{eqnarray}
307:
308:
309: However, it is important to take note of the boundary terms. Strictly,
310: in order to have Eqs. (\ref{u_def}) and (\ref{u_def_c}) identical,
311: we must make the identification\cite{cyclic}
312:
313: \[
314: c_{N+1}=c_{1}(e^{i\sum_{j=1}^{N}c_{j}^{\dagger}c_{j}}+1)\]
315:
316:
317: It may be argued that in the thermodynamic limit, this term will be
318: irrelevant, and we may make the identification $c_{N+1}=c_{1}$.
319:
320: Due to cyclic boundary conditions, we will require $k$ to take the
321: discrete values
322:
323: \[
324: k=\frac{2\pi m}{L},m=-\frac{L}{2},...,-1,0,1,\frac{L-2}{2}\]
325:
326:
327: These operators satisfy fermion anti-commutation relations:
328:
329: \begin{eqnarray*}
330: \{ C_{k},C_{l}^{\dagger}\} & = & \delta_{kl}\\
331: \{ C_{k},C_{l}\} & = & \{ C_{k}^{\dagger},C_{l}^{\dagger}\}=0\end{eqnarray*}
332:
333:
334: Using the definitions of $c_{n}$ and $c_{n}^{\dagger}$, and the
335: thermodynamic limit we can re-write $U(\chi,\theta)$ as
336:
337: \begin{eqnarray}
338: U(\chi,\theta) & = & e^{-i\chi\sum_{k}2\cos kC_{k}^{\dagger}C_{k}-i\sin k(C_{k}^{\dagger}C_{-k}^{\dagger}+C_{k}C_{-k})}\label{eq:}\\
339: & & e^{-i\theta\sum_{k}(2C_{k}^{\dagger}C_{k}-1)}\nonumber \end{eqnarray}
340: where we require the thermodynamic limit so that the property $2\sum_{k}C_{k}^{\dagger}C_{k}=\sum_{k}(C_{k}^{\dagger}C_{k}+C_{-k}^{\dagger}C_{-k})$
341: holds.
342:
343: To simplify matters, let us further define
344:
345: \begin{eqnarray}
346: \hat{A}_{k} & = & \chi(2\cos kC_{k}^{\dagger}C_{k}-i\sin k(C_{k}^{\dagger}C_{-k}^{\dagger}+C_{k}C_{-k}))\label{eq:}\\
347: \hat{B}_{k} & = & \theta(2C_{k}^{\dagger}C_{k}-1)\label{eq:}\end{eqnarray}
348: such that we may write
349:
350: \begin{equation}
351: U(\chi,\theta)=e^{-i\sum_{k}\hat{A}_{k}}e^{-i\sum_{k}\hat{B}_{k}}=\Pi_{k}U_{k}(\chi,\theta)\label{eq:}\end{equation}
352: where $U_{k}(\chi,\theta)\equiv e^{-i\hat{A}_{k}}e^{-i\hat{B}_{k}}$.
353:
354: We have now completely decoupled the problem, and may express the
355: operators $\hat{A_{k}}$ and $\hat{B_{k}}$ in the basis $\left|0\right\rangle ,C_{k}^{\dagger}\left|0\right\rangle ,C_{-k}^{\dagger}\left|0\right\rangle ,C_{k}^{\dagger}C_{-k}^{\dagger}\left|0\right\rangle $.
356: It is be possible to find eigenstates of $U(\chi,\theta)$ in closed
357: form in this basis.
358:
359:
360: \subsection{Combining }
361:
362: However, it would be nice to be able to express $U(\chi,\theta)$
363: as a single exponential. While $\hat{A_{k}}$ and $\hat{B_{k}}$ do
364: not commute, it turns out that there is a faithful representation
365: in $SU(2)$, if we make the following definitions :
366:
367: \begin{eqnarray*}
368: \nu_{1}^{(k)} & = & -i(C_{k}^{\dagger}C_{-k}^{\dagger}+C_{k}C_{-k})\\
369: \nu_{2}^{(k)} & = & (-C_{k}^{\dagger}C_{-k}^{\dagger}+C_{k}C_{-k})\\
370: \nu_{3}^{(k)} & = & C_{k}^{\dagger}C_{k}+C_{-k}^{\dagger}C_{-k}-I\end{eqnarray*}
371:
372:
373: Hence, we can express $\hat{A}_{k}=\chi(\cos k+\vec{\alpha}.\vec{\nu_{k}})$
374: and $\hat{B}_{k}=\theta\vec{\beta}_{k}.\vec{\nu}_{k}$, where $\vec{\alpha}_{k}=\chi(\sin k,0,\cos k)$
375: and $\vec{\beta}_{k}=\theta(0,0,1)$. We have that $[\nu_{l}^{(k)},\nu_{m}^{(k')}]=-2i\epsilon_{l,m,n}\nu_{n}^{(k)}\delta_{k}^{k'}$
376: where $\epsilon_{l,m,n}$ is the Levi-Civita symbol, so that $\{\nu_{1}^{(k)},\nu_{2}^{(k)},\nu_{3}^{(k)}\}$
377: have the same properties as the $SU(2)$ matrices $\{\sigma_{1},\sigma_{2},\sigma_{3}\}$.
378: Relating the fermionic operators to $SU(2)$ in this way was inspired
379: by a similar approach in the theory of superconductors\cite{Schreiffer}.
380:
381: $SU(2)$ is closed under composition with a well understood composition
382: relation, which we can now apply to our system\cite{closed2}
383:
384: \begin{equation}
385: U_{k}(\chi,\theta)=e^{-i\chi\vec{\alpha}_{k}.\vec{\nu_{k}}}e^{-i\theta\vec{\beta}_{k}.\vec{\nu_{k}}}=e^{-i\cos k}e^{-i\kappa_{k}\vec{\gamma}_{k}(\chi,\theta).\vec{\nu_{k}}}\label{u_final}\end{equation}
386: where
387:
388: \begin{eqnarray}
389: \vec{\gamma}_{k}(\chi,\theta) & = & (\sin k\cos\theta\sin\chi,-\sin k\sin\theta\sin\chi,\label{gamma_def}\\
390: & & (\sin\theta\cos\chi+\cos k\cos\theta\sin\chi))\nonumber \\
391: \kappa_{k} & = & {\frac{\cos^{-1}\eta_{k}}{\sqrt{1-\eta_{k}^{2}}}}\label{kappa_def}\\
392: \eta_{k} & = & \cos\theta\cos\chi-\cos k\sin\theta\sin\chi\label{eta_def}\\
393: & = & \cos^{2}\frac{k}{2}\cos(\theta+\chi)+\sin^{2}\frac{k}{2}\cos(\theta-\chi)\nonumber \end{eqnarray}
394:
395:
396: This composition has the simple physical interpretation of two rotations
397: being composed, and the result can be derived using quaternion composition\cite{closed}.
398: However, when using quaternions, special care has to be given to the
399: double cover of $SO(3)$ under $SU(2)$. The second equality of Eq.
400: (\ref{u_final}) defines an effective Hamiltonian, $\bar{H}_{k}$,
401: and we stress that $\bar{H}_{k}\neq\hat{A}_{k}+\hat{B}_{k}$ because
402: $\hat{A}_{k}$and $\hat{B}_{k}$ do not commute.
403:
404: Hence, we have the final form of the decoupled, and combined transformation
405:
406: \begin{equation}
407: U(\chi,\theta)=\Pi_{k}U_{k}(\chi,\theta)=e^{-i\sum_{k}\kappa_{k}\vec{\gamma}_{k}(\chi,\theta).\vec{\nu_{k}}}\label{decoupled}\end{equation}
408:
409:
410: Hence, we have found that the effective Hamiltonian , $\bar{H}$,
411: defined in Eq (\ref{u_def}) is given by $\bar{H}=\sum_{k}\kappa_{k}\vec{\gamma}_{k}(\chi,\theta).\vec{\nu_{k}}$
412:
413: We now check the limit $\chi\rightarrow0$, which implies $\kappa\rightarrow{\frac{\theta}{\sin\theta}}$,
414: and $\vec{\gamma}_{k}\rightarrow\{0,0,\cos k\sin\theta\}$. Hence
415: $U(\chi,\theta)=e^{-i\sum_{k}{\frac{\theta}{\sin\theta}}\sin\theta\cos k\nu_{3}}=e^{-i\theta\sum_{k}{\frac{-i}{2}}\sigma_{x}\sigma_{y}-\sigma_{y}\sigma_{x}}=e^{-i\theta\sum_{k}{\frac{-i}{2}}2i\sigma_{x}\sigma_{y}}=e^{-i\theta\sum_{k}\sigma_{z}}$.
416: On the other hand, in the limit $\theta\rightarrow0$, $\kappa={\frac{\chi}{\sin\chi}}$,
417: and $\vec{\gamma}_{k}=\{\sin\chi\sin k,0,\sin\chi\cos k\}$. Hence
418: $U(\chi,\theta)=e^{-i\chi\sum_{k}\sin k\nu_{1}+\cos_{k}\nu_{3}}$.
419: Thus, we retrieve the expected behaviour in the limit as we turn off
420: either the exchange or external field terms.
421:
422: Having expressed $U(\chi,\theta)$ in this form, it is now possible
423: to show that it directly corresponds to some physical Hamiltonian.
424: We may perform a Bogoliubov transformation by defining some fermion
425: creation operator
426:
427: \begin{equation}
428: \gamma_{k}\gamma_{k}^{\dagger}=\vec{\gamma}_{k}(\chi,\theta).\vec{\nu_{k}}\label{eq:}\end{equation}
429: with associated energy, $\epsilon_{k}=\kappa_{k}$. Hence, we may
430: consider our ground state as a vacuum state $|0>$, and excitations
431: as $\gamma_{k}^{\dagger}\left|0\right\rangle $. It is important to
432: note here that the excitations of lowest energy will occur at an extremum
433: of $\epsilon_{k}$. We can show that this occurs at $k=0,\pi$ by
434: noting that
435:
436: \begin{equation}
437: \frac{\partial\epsilon_{k}}{\partial k}=\frac{\partial\kappa_{k}}{\partial k}=\frac{\partial\kappa_{k}}{\partial\eta_{k}}\frac{\partial\eta_{k}}{\partial k}\label{eq:}\end{equation}
438: from which it follows that
439:
440: \begin{equation}
441: \left.\frac{\partial\eta_{k}}{\partial k}\right|_{k=0,\pi}=\left.\sin k\sin\theta\sin\chi\right|_{k=0,\pi}=0\label{eq:}\end{equation}
442:
443:
444: Hence, the elementary excitations will be for $k=0$ or $k=\pi$,
445: whichever corresponds to a lower energy.
446:
447:
448: \subsection{Closed-form Hamiltonian}
449:
450: We can now work backwards from our expression for $U(\chi,\theta)$
451: to a single combined Hamiltonian. The results here will only be valid
452: in the thermodynamic limit, which we have assumed in the previous
453: section. Before doing so, we should present a list of identities which
454: will prove to be useful.
455:
456: \begin{eqnarray*}
457: \sum_{k}e^{iak}\nu_{1}^{(k)} & = & i\sum_{n}c_{n}^{\dagger}c_{n+a}^{\dagger}-c_{n}c_{n+a}\\
458: \sum_{k}e^{iak}\nu_{2}^{(k)} & = & -\sum_{n}c_{n}^{\dagger}c_{n+a}^{\dagger}+c_{n}c_{n+a}\\
459: \sum_{k}e^{iak}\nu_{3}^{(k)} & = & \sum_{n}2c_{n}^{\dagger}c_{n}-I\\
460: \sum_{k}\nu_{1}^{(k)}=0 & , & \sum_{k}\cos(ak)\nu_{1}^{(k)}=0\\
461: \sum_{k}\sin(ak)\nu_{1}^{(k)} & = & \sum_{n}c_{n}^{\dagger}c_{n+a}^{\dagger}-c_{n}c_{n+a}\\
462: \sum_{k}\nu_{2}^{(k)}=0 & , & \sum_{k}\cos(ak)\nu_{2}^{(k)}=0\\
463: \sum_{k}\sin(ak)\nu_{2}^{(k)} & = & i\sum_{n}(c_{n}^{\dagger}c_{n+a}^{\dagger}+c_{n}c_{n+a})\end{eqnarray*}
464:
465:
466: Now, recall that
467:
468: \begin{equation}
469: U(\chi,\theta)=e^{-i\sum_{k}\kappa_{k}\vec{\gamma}_{k}(\chi,\theta).\vec{\nu}_{k}}\label{eq:}\end{equation}
470:
471:
472: However, $\kappa_{k}$ is an even function of $k$, and so can be
473: expanded in terms of a Fourier series involving $\cos k$. Let us
474: write
475:
476: \begin{equation}
477: \kappa_{k}=\sum_{l=0}^{\infty}a_{l}\cos(lk)\label{eq:}\end{equation}
478:
479:
480: \begin{equation}
481: U(\chi,\theta)=e^{-i\sum_{k,l}a_{l}\cos(lk)\vec{\gamma}_{k}(\chi,\theta).\vec{\nu}_{k}}\label{a_series}\end{equation}
482:
483:
484: We will now substitute our expression for $\vec{\gamma}_{k}$ and
485: expand.
486:
487: \begin{eqnarray}
488: U(\chi,\theta) & = & e^{-i\sum_{k,l}a_{l}\cos lk\vec{\gamma_{k}}(\chi,\theta).\vec{\nu}_{k}}\label{eq:}\\
489: & = & e^{-i(\Lambda_{1}+\Lambda_{2}+\Lambda_{3})}\label{eq:}\\
490: & = & e^{-i\bar{H}}\label{eq:}\end{eqnarray}
491:
492:
493: where
494:
495: \begin{eqnarray}
496: \Lambda_{1} & = & \cos\theta\sin\chi\sum_{k,l}a_{l}\sin k\cos(lk)\nu_{1}^{(k)}\label{eq:}\\
497: \Lambda_{2} & = & -\sin\theta\sin\chi\sum_{k,l}a_{l}\sin k\cos(lk)\nu_{2}^{(k)}\label{eq:}\\
498: \Lambda_{3} & = & \sin\theta\cos\chi\sum_{k,l}a_{l}\cos(lk)\nu_{3}^{(k)}\label{eq:}\\
499: & & +\cos\theta\sin\chi\sum_{k,l}a_{l}\cos k\cos(lk)\nu_{3}^{(k)}\nonumber \end{eqnarray}
500:
501:
502: and the sum over $l$ ranges $1,2,3...$
503:
504: We may rewrite this as
505:
506: \begin{eqnarray}
507: \Lambda_{1} & = & \cos\theta\sin\chi\sum_{k,l}a_{l}[\sin(l+1)k-\sin(l-1)k]\nu_{1}^{(k)}\label{eq:}\\
508: & = & \cos\theta\sin\chi[a_{0}(c_{n}^{\dagger}c_{n+1}^{\dagger}-c_{n}c_{n+1})\label{eq:}\\
509: & & +\sum_{n,l}\frac{(a_{l+1}-a_{l-1})}{2}(c_{n}^{\dagger}c_{n+l}^{\dagger}-c_{n}c_{n+l})]\nonumber \\
510: \Lambda_{2} & = & -\sin\theta\sin\chi\sum_{k,l}a_{l}[\sin(l+1)k-\sin(l-1)k]\nu_{2}^{(k)}\label{eq:}\\
511: & = & -i\sin\theta\sin\chi\sum_{n,l}[a_{o}(c_{n}^{\dagger}c_{n+1}^{\dagger}+c_{n}c_{n+1})\label{eq:}\\
512: & & +\sum_{l}\frac{(a_{l+1}-a_{l-1})}{2}(c_{n}^{\dagger}c_{n+l}^{\dagger}+c_{n}c_{n+l})]\nonumber \\
513: \Lambda_{3} & = & \sin\theta\cos\chi\sum_{k,l}a_{l}\cos(lk)\nu_{3}^{(k)}\label{eq:}\\
514: & & +\cos\theta\sin\chi\sum_{k,l}a_{l}[\cos(l+1)k+\cos(l-1)k]\nu_{3}^{(k)}\nonumber \\
515: & = & \sin\theta\cos\chi\sum_{l}[a_{l}(c_{n}^{\dagger}c_{n+l}-c_{n}c_{n+l}^{\dagger})]+\label{eq:}\\
516: & & +\cos\theta\sin\chi\sum_{n,l}[a_{0}(c_{n}^{\dagger}c_{n+1}-c_{n}c_{n+1}^{\dagger})\nonumber \\
517: & & +\frac{(a_{l+1}-a_{l-1}))}{2}(c_{n}^{\dagger}c_{n+l}-c_{n}c_{n+l}^{\dagger})]\nonumber \end{eqnarray}
518:
519:
520: Therefore, the quantum spin chain Hamiltonian $\bar{H}$ which represents
521: a physical system corresponding to the separated unitary map (\ref{u_def})
522: is highly non-local. Terms such as $c_{n}c_{n+a}^{\dagger}=a_{n}e^{-i\pi\sum_{j=n}^{n+a-1}a_{j}a_{j}^{\dagger}}a_{n+a}^{\dagger}$
523: for $a>1$ will not only involve $a_{n},a_{n+a}$, but also $c_{m}$
524: and $c_{m}^{\dagger}$ $\forall n<m<n+a$.
525:
526: If we define
527:
528: \begin{eqnarray*}
529: \sigma_{+}^{(n)} & = & \frac{\sigma_{z}^{(n)}+i\sigma_{y}^{(n)}}{2}\\
530: \sigma_{-}^{(n)} & = & \frac{\sigma_{z}^{(n)}-i\sigma_{y}^{(n)}}{2}\end{eqnarray*}
531: we can write $c_{n}c_{n+a}^{\dagger}$ as $\sigma_{+}^{(n)}e^{-i\pi\sum_{j=n}^{n+a-1}\frac{1-\sigma_{x}^{(j)}}{2}}\sigma_{-}^{(n+a)}$,
532: explicitly showing the dependence on non-neighbouring spins.
533:
534:
535: \subsection{Range of the Interactions}
536:
537: To be in the universality class of the Ising model, we would expect
538: the non-local terms, $a_{l}$ to decrease exponentially with separation
539: $l$. Thus, when viewed at larger length scales, the non-local terms
540: would become irrelevant. The behaviour of $a_{n}$ for a variety of
541: $\theta=\chi$ is calculated numerically and presented in Fig. \ref{a_coeff}.
542: The case for $\theta\neq\chi$ is similar, and displays the same exponential
543: decrease in $a_{l}$ with $l$. Hence, we may naturally expect nearest-neighbour
544: interactions to be the most important interactions in this model -
545: an idea which we will make concrete in the following section.
546:
547: %
548: \begin{figure}
549: \includegraphics[%
550: width=3.04414in]{a_coefficients.ps}
551:
552:
553: \caption{\label{a_coeff}The behaviour of Fourier coefficients $a_{n}$, as
554: defined in Eq. \ref{a_series}, for $n=1,2...10$ for a variety of
555: $\theta$ and $\chi$. Note that the larger $\left|\chi-\frac{\pi}{2}\right|$
556: and $\left|\theta-\frac{\pi}{2}\right|$, the larger the decay. The
557: exponential decay in these coefficients implies that the interactions
558: in the Hamiltonian are short-ranged, and suggests that renormalization
559: techniques should be highly effective in this model.}
560: \end{figure}
561:
562:
563: Deriving an analytic expression for $a_{n}$ seems to be very difficult.
564: However it is possible to show a general dependence on $\theta$ and
565: $\chi$ of the form $\sin^{n}\theta\sin^{n}\chi$ for any particular
566: $n$. We can expand $\kappa_{k}$ in terms of $\eta_{k}$, where $\eta_{k}$
567: is defined in Eq. (\ref{eta_def}), as
568:
569: \begin{equation}
570: \kappa_{k}=\sum_{p=0}^{\infty}\frac{2^{p-q}\Gamma^{2}(\frac{p+1}{2})}{\Gamma(p+1)}\eta_{k}^{p}=\sum_{p=0}^{\infty}c_{p}\eta_{k}^{p}\label{kappa_series}\end{equation}
571: where $\Gamma$ is the Euler gamma function. In turn, $\eta^{p}$
572: can be expressed as a series in terms of $\cos^{q}k$ as
573:
574: \begin{eqnarray}
575: \eta_{k}^{p} & = & \sum_{q=0}^{p}\left(\begin{array}{c}
576: p\\
577: q\end{array}\right)(\cos\theta\cos\chi)^{p-q}(-\cos k\sin\theta\sin\chi)^{q}\label{eta_series}\\
578: & = & \sum_{q=0}^{p}d_{p,q}\cos^{q}k\nonumber \end{eqnarray}
579:
580:
581: Finally, we can express $\cos^{q}k$ in terms of $\cos rk$. For the
582: case of $q$ even, this is
583:
584: \begin{eqnarray}
585: \cos^{q}k & = & \sum_{r=0,r\in evens}^{q}\left(\begin{array}{c}
586: q\\
587: \frac{q-r}{2}\end{array}\right)\frac{\cos rk}{2^{q-1}}+c\label{cos_series}\\
588: & = & \sum_{r=0}^{q}e_{q,r}\cos rk\nonumber \end{eqnarray}
589: where $c$ is an unenlightening constant, and a similar expression
590: holds for $p-q$ odd.
591:
592: We can combine Equations (\ref{kappa_series}),(\ref{eta_series})
593: and (\ref{cos_series}) to yield an expression for $\kappa_{k}$
594:
595: \begin{equation}
596: \kappa_{k}=\sum_{p=0}^{\infty}c_{p}\eta^{p}=\sum_{p=0}^{\infty}\sum_{q=0}^{p}\sum_{r=0}^{q}c_{p}d_{p,q}e_{q,r}\cos rk\label{eq:}\end{equation}
597: from which we can read the coefficient $a_{l}$ of $\cos lk$ as
598:
599: \begin{equation}
600: a_{l}=\sum_{p=0}^{\infty}c_{p}\eta^{p}=\sum_{p=0}^{\infty}\sum_{q=0}^{p}c_{p}d_{p,q}e_{q,l}\label{eq:}\end{equation}
601:
602:
603: However, $e_{q,l}$ is only non-zero for $q>l$, and so we can replace
604: $\sum_{q=0}^{p}$ with $\sum_{q=l}^{p}$ to yield
605:
606: \begin{eqnarray}
607: a_{l} & = & \sum_{p=0}^{\infty}c_{p}\eta^{p}=\sum_{p=0}^{\infty}c_{p}\sum_{q=l}^{p}d_{p,q}e_{q,l}\label{eq:}\end{eqnarray}
608: which involves terms in $d_{p,q}\propto(\sin\theta\sin\chi)^{q}(\cos\theta\cos\chi)^{s}$
609: for $q>l$. Hence, for any given $a_{l}$, there is a behaviour proportional
610: to $(\sin\theta\sin\chi)^{l}$.
611:
612: Of physical interest is the behaviour of $a_{l}$ with respect to
613: $l$ for a given $\theta$ and $\chi$. We will attempt to factor
614: out any behaviour in $l$ by noting that since $d_{pq},e_{q,l}<1$,
615: we can write
616:
617: \begin{eqnarray}
618: a_{l} & = & \sum_{p=0}^{\infty}c_{p}\sum_{q=l}^{p}d_{p,q}e_{q,l}\label{eq:}\\
619: & \leq & \sum_{p=0}^{\infty}c_{p}\sum_{q=l}^{p}d_{p,q}\sum_{q=l}^{p}e_{q,l}\nonumber \end{eqnarray}
620:
621:
622: However, if we turn our attention to the sum over $d_{p,q}$, we can
623: construct a further limit on $a_{l}$
624:
625: \begin{eqnarray}
626: \sum_{q=l}^{p}d_{p,q} & = & \sum_{q=l}^{p}\left(\begin{array}{c}
627: p\\
628: q\end{array}\right)(\cos\theta\cos\chi)^{p-q}(-\cos k\sin\theta\sin\chi)^{q}\label{eq:}\\
629: & \leq & \sum_{q'=0}^{p-l}\left(\begin{array}{c}
630: p\\
631: q'+l\end{array}\right)(\cos\theta\cos\chi)^{p-q'-l}(\sin\theta\sin\chi)^{q'+l}\label{eq:}\\
632: & \leq & \sum_{q'=0}^{p-l}\left(\begin{array}{c}
633: p-l\\
634: q'\end{array}\right)(\cos\theta\cos\chi)^{p-l-q'}(\sin\theta\sin\chi)^{q'+l}\label{eq:}\\
635: & \leq & (\sin\theta\sin\chi)^{l}(\cos\theta\cos\chi+\sin\theta\sin\chi)^{p-l}\label{eq:}\\
636: & \leq & (\sin\theta\sin\chi)^{l}\label{eq:}\end{eqnarray}
637:
638:
639: Leading to our strictest inequality for $a_{l}$, showing an exponential
640: decay in $l$
641:
642: \begin{eqnarray}
643: a_{l} & \leq & (\sin\theta\sin\chi)^{l}\sum_{p=0}^{\infty}c_{p}\sum_{q=l}^{p}e_{q,l}\label{eq:}\\
644: & \leq & (\sin\theta\sin\chi)^{l}(\sum_{p=0}^{\infty}c_{p}p\max_{q}e_{q,l})\label{eq:}\end{eqnarray}
645:
646:
647: The case $\theta,\chi\rightarrow\frac{\pi}{2}$ has asymptotically
648: constant $a_{l}$. However, for $\theta,\chi\neq\frac{\pi}{2}$, we
649: can say that
650:
651: \begin{eqnarray}
652: a_{l}\leq ke^{-\mu l}\rightarrow0 & \mathrm{\mathrm{as}} & l\rightarrow\infty\label{eq:}\end{eqnarray}
653: where $\mu=\ln(\sin\theta\sin\chi)$. Thus the terms in our model
654: $\bar{H}$ has only short range interactions.
655:
656:
657: \subsection{Renormalizing the Hamiltonian}
658:
659: We have seen that the Hamiltonian, $\bar{H}$, associated with the
660: complete unitary $U$ is very complicated and involves non-local interactions.
661: Hence, we are forced to use renormalization group methods to extract
662: the interesting physics from this case.
663:
664: Consider the continuum limit of the Hamiltonian, $\bar{H}$, which
665: is applicable in the thermodynamic limit. Near criticality, the physics
666: will be driven by long-wavelength effects, which suggest that the
667: wavevector $k$ will be small. The relevant excitations at low temperature
668: happen at the extremum of $\epsilon_{k}$, which we have shown occurs
669: at $k=0$. Hence, in the continuum limit, we can consider only low
670: lying states, near $k=0$. Under these approximations, our Hamiltonian
671: becomes
672:
673: \begin{eqnarray}
674: \bar{H}\simeq\bar{H}' & = & \sum_{k}\bar{\kappa}\vec{\bar{\gamma}}_{k}(\chi,\theta).\vec{\nu}_{k}\label{eq:}\\
675: \vec{\bar{\gamma}}_{k} & = & (k\cos\theta\sin\chi,-k\sin\theta\sin\chi,\sin(\theta+\chi))\label{eq:}\\
676: \bar{\kappa} & = & \frac{\theta+\chi}{\sin(\theta+\chi)}\label{eq:}\end{eqnarray}
677: which yields in terms if the fermion operators
678:
679: \begin{eqnarray}
680: \bar{H}' & = & \sum_{k}ik\frac{(\theta+\chi)\sin\chi}{\sin(\theta+\chi)}(e^{i\theta}C_{k}^{\dagger}C_{-k}^{\dagger}+e^{-i\theta}C_{k}C_{-k})\label{eq:}\\
681: & & +2(\theta+\chi)C_{k}^{\dagger}C_{k}\nonumber \end{eqnarray}
682:
683:
684: Defining the continuum Fermi field\cite{Fields} as
685:
686: \begin{equation}
687: \Psi(x_{i})=\frac{1}{\sqrt{a}}c_{i}\label{eq:}\end{equation}
688: where $a$ is the lattice spacing. $\Psi(x)$ satisfies the usual
689: anti-commutation relation $\{\Psi(x),\Psi^{\dagger}(x')\}=\delta(x-x')$.
690: Note that we can replace the sum over $k$ with an integral, by making
691: the substitution
692:
693: \[
694: \sum_{k}a\rightarrow\int dx\]
695:
696:
697: Further, we expand the terms $C_{k}^{\dagger}C_{-k}^{\dagger}$ and
698: $C_{k}C_{-k}$ into terms of first order gradients $\frac{\partial\Psi^{\dagger}}{\partial x}$
699: and $\frac{\partial\Psi}{\partial x}$ through the use of identities
700: (Chapter 4 of Ref. \cite{Sachdev}). This yields
701:
702: \begin{eqnarray}
703: \bar{H}'\simeq\tilde{H} & = & E_{0}+\int dx[\frac{(\theta+\chi)\sin\chi}{\sin(\theta+\chi)}(e^{i\theta}\Psi^{\dagger}\frac{\partial\Psi^{\dagger}}{\partial x}-e^{-i\theta}\Psi\frac{\partial\Psi}{\partial x})\label{eq:}\\
704: & & -2(\theta+\chi)\Psi^{\dagger}\Psi]\nonumber \end{eqnarray}
705: and $E_{0}$ is some constant.
706:
707: Applying the transformation $\Psi\rightarrow e^{i\frac{\theta}{2}}\Psi$,
708: we have that
709:
710: \begin{eqnarray}
711: \tilde{H} & = & E_{0}+\int dx[\frac{(\theta+\chi)\sin\chi}{\sin(\theta+\chi)}(\Psi^{\dagger}\frac{\partial\Psi^{\dagger}}{\partial x}-\Psi\frac{\partial\Psi}{\partial x})\label{eq:}\\
712: & & -2(\theta+\chi)\Psi^{\dagger}\Psi]\nonumber \end{eqnarray}
713:
714:
715: If we do not perform this transformation, we will have terms of the
716: form $\int dx\frac{\partial\Psi^{\dagger}}{\partial x}+\Psi\frac{\partial\Psi}{\partial x}$
717: in the Hamiltonian. One can show that these terms correspond to interactions
718: of the form $\sigma_{z}^{(n)}\sigma_{y}^{(n+1)}+\sigma_{y}^{(n)}\sigma_{z}^{(n+1)}$.
719: Chapter 4 of Ref. \cite{Fields} has further details regarding these
720: sorts of chiral symmetries in systems.
721:
722: One can show that the Lagrangian corresponding to this Hamiltonian
723: will then be
724:
725: \begin{eqnarray}
726: \mathcal{\tilde{L}} & = & \Psi^{\dagger}\frac{\partial\Psi}{\partial\tau}+\frac{(\theta+\chi)\sin\chi}{\sin(\theta+\chi)}(\Psi^{\dagger}\frac{\partial\Psi^{\dagger}}{\partial x}-\Psi\frac{\partial\Psi}{\partial x})\label{eq:}\\
727: & & -2(\theta+\chi)\Psi^{\dagger}\Psi\nonumber \end{eqnarray}
728: where $\tau$ is imaginary time.
729:
730: Now we introduce the crucial step where we considering the effect
731: of scaling the problem. If we consider the effect of viewing the problem
732: at a scale $\delta^{l}$ more coarse in space, and $\delta^{zl}$
733: more coarse in time, we can introduce the new variables
734:
735: \begin{eqnarray}
736: x' & = & x\delta^{-l}\label{eq:}\\
737: \tau' & = & \tau\delta^{-zl}\label{eq:}\\
738: \Psi' & = & \Psi\delta^{l/2}\label{eq:}\end{eqnarray}
739:
740:
741: We choose the value of the dynamic critical exponent, $z$, to be
742: identically equal to $1$, corresponding to an isotropy between space
743: and time, in order to leave the velocity-like coefficients of $\Psi^{\dagger}\frac{\partial\Psi^{\dagger}}{\partial x}$
744: and $\Psi\frac{\partial\Psi}{\partial x}$ unchanged.
745:
746: For criticality to hold, these scaling conditions must leave the Lagrangian
747: unchanged. This occurs only when the quantity $\theta+\chi$ is identically
748: zero. This happens for $\theta=-\chi$, exactly as in the transverse
749: Ising model.
750:
751: Formally, if we write
752:
753: \begin{equation}
754: \tilde{\mathcal{L}}=\Psi^{\dagger}\frac{\partial\Psi}{\partial\tau}+u(\Psi^{\dagger}\frac{\partial\Psi^{\dagger}}{\partial x}-\Psi\frac{\partial\Psi}{\partial x})+\Delta\Psi^{\dagger}\Psi\label{eq:}\end{equation}
755: we will require that
756:
757: \begin{eqnarray}
758: \Delta' & = & \Delta\delta^{l}\label{eq:}\\
759: u' & = & u\label{eq:}\end{eqnarray}
760: implying that the scaling dimension of the term $\Delta\Psi^{\dagger}\Psi$,$\dim(\Delta)=1$.
761: Hence, $u$, and $\Delta$ are relevant parameters, having non-negative
762: scaling factors.
763:
764: If we include second (or higher) order effects in $k$, we will include
765: terms of the form $\Delta'\Psi^{\dagger}\frac{\partial^{2}\Psi}{\partial x^{2}}$
766: or $\Delta''\Psi^{\dagger}\frac{\partial\Psi^{\dagger}}{\partial x}\frac{\partial\Psi}{\partial x}\Psi$
767: (or higher derivatives). From a simple analysis, one can show that
768: the parameters $\Delta'$ and $\Delta''$ are irrelevant, as the scaling
769: dimensions are\cite{Sachdev}
770:
771: \begin{eqnarray*}
772: \dim(\Delta')=-1 & , & \dim(\Delta'')=-2\end{eqnarray*}
773:
774:
775: Recall the Ising chain in a transverse field, Eq. (\ref{ising}),
776: given by
777:
778: \begin{eqnarray}
779: H_{Ising} & = & \sum_{n=1}^{N}\mu B\sigma_{x}^{(n)}+J\sigma_{z}^{(n)}\sigma_{z}^{(n+1)}\label{eq:}\\
780: & = & \sum_{k}2BC_{k}^{\dagger}C_{k}+2J\cos kC_{k}^{\dagger}C_{k}\label{eq:}\\
781: & & +iJ\sin k(C_{k}^{\dagger}C_{-k}^{\dagger}+C_{k}C_{-k})\nonumber \end{eqnarray}
782: The Lagrangian for this model has the form
783:
784: \begin{eqnarray}
785: \mathcal{L}_{Ising} & = & \Psi^{\dagger}\frac{\partial\Psi}{\partial\tau}+2(B+J)\Psi^{\dagger}\Psi-J(\Psi^{\dagger}\frac{\partial\Psi^{\dagger}}{\partial x}-\Psi\frac{\partial\Psi}{\partial x})\label{eq:}\end{eqnarray}
786:
787:
788: Hence, we may make an association between the two models through the
789: mapping
790:
791: \begin{eqnarray}
792: J & = & -\frac{(\theta+\chi)\sin\chi}{\sin(\theta+\chi)}\label{eq:}\\
793: B+J & = & \theta+\chi\label{eq:}\end{eqnarray}
794:
795:
796: Thus we conclude that our continuum Hamiltonian $\tilde{H}$ belongs
797: in the same universality class as the transverse Ising Hamiltonian,
798: $H_{Ising}$. Hence it is possible to access the physical properties
799: at criticality of this well known model in a very straight forward
800: manner. The crucial assumptions which have been made are the assumption
801: of operation in the thermodynamic limit, and the low temperature (and
802: hence small $k$ excitation) regime, both of which are necessary for
803: renormalization to work. We will show in the next section that for
804: moderate $N$, the signatures of quantum phase transitions are still
805: observable.
806:
807:
808: \section{Signatures of a Quantum Phase Transition}
809:
810: The first experimental realisations of a quantum simulations will
811: perhaps be seen on ion trap quantum computers. In this section we
812: review a few basic experimental signatures which may be seen in an
813: ion trap laboratory.
814:
815:
816: \subsection{Ground State Energy }
817:
818: Recall that we have the following unitary map which describes the
819: system in terms of non-interacting fermions:
820:
821: \begin{equation}
822: U(\chi,\theta)=e^{-i\sum_{k}\kappa_{k}\vec{\gamma}_{k}(\chi,\theta).\vec{\nu_{k}}}\label{eq:}\end{equation}
823:
824:
825: By inspection, the corresponding Hamiltonian is given by:
826:
827: \begin{equation}
828: \bar{H}=\sum_{k}\kappa_{k}\vec{\gamma}_{k}(\chi,\theta).\vec{\nu_{k}}\label{eq:}\end{equation}
829:
830:
831: We note that eigenstates of $\bar{H}$ are simply products of the
832: eigenstates of $\bar{H}_{k}$, where $\bar{H}_{k}=\kappa_{k}\vec{\gamma}_{k}(\chi,\theta).\vec{\nu_{k}}$.
833: In the basis $\left|0\right\rangle ,C_{k}^{\dagger}\left|0\right\rangle ,C_{-k}^{\dagger}\left|0\right\rangle ,C_{k}^{\dagger}C_{-k}^{\dagger}\left|0\right\rangle $,
834: there will be four complex eigenvalues $\{\lambda_{k}^{(i)},i=1,2,3,4\}$,
835: with arguments $\{\omega_{k}^{(i)},i=1,2,3,4\}$. Let us define the
836: argument of a total system state, $\Omega^{(i)}=\sum_{k}\omega_{k}^{(i)}$,
837: where we form an eigenstate of $\bar{H}$ from matching eigenstates
838: of $\bar{H}_{k}$. Physically, we associate the argument of this state
839: with energy.
840:
841: Now, we will consider the mapping
842:
843: \begin{eqnarray*}
844: \chi\rightarrow r\cos\phi & , & \theta\rightarrow r\sin\phi\end{eqnarray*}
845: where $\phi$ can be considered as a relative strength between exchange
846: and field coupling terms, and $r$ is an overall strength. The Ising
847: criticality condition $\theta=\pm\chi$ is now $\phi=\pm\frac{\pi}{4},\pm\frac{3\pi}{4}$.
848:
849: In Fig. \ref{6q_d2phase}, we observe an sharp peak in the second
850: derivative of the ground state energy, $\Omega^{(1)}$, with respect
851: to $\phi$. In the thermodynamic limit, this would become a singularity,
852: indicating a second order phase transition. Further, we observe that
853: this condition occurs for $\phi=\pm\frac{\pi}{4},\pm\frac{3\pi}{4}$,
854: which corresponds to the Ising transition.
855:
856: We will demonstrate the nature of this singularity explicitly, by
857: noting that the argument of the ground state energy is
858:
859: \begin{eqnarray}
860: \omega_{k}^{(1)} & = & -\kappa\sqrt{(\cos k\cos\theta\sin\chi+\sin\theta\cos\chi)^{2}+(\sin k\sin\chi)^{2}}\label{eq:}\\
861: & = & \cos^{-1}\eta_{k}\equiv E_{k}\label{eq:}\end{eqnarray}
862:
863:
864: The next two eigenvalues are equal to $1$, and hence $\omega_{k}^{(2)}=\omega_{k}^{(3)}=0$.
865: The highest excited eigenstate has $\omega_{k}^{(4)}=-\omega_{k}^{(1)}$
866: by symmetry.
867:
868: Substituting $\theta\rightarrow r\cos\phi$ and $\chi\rightarrow r\sin\phi$,
869: we evaluate
870:
871: \begin{eqnarray}
872: \left.\frac{\partial^{2}\Omega^{(1)}}{\partial\phi^{2}}\right|{}_{\phi=\pm\frac{\pi}{4},\pm\frac{3\pi}{4}} & \simeq & \frac{2N}{\pi}\int_{0}^{\pi-\frac{\pi}{N}}\frac{(\cos k-1)dk}{\sqrt{1-(\cos^{2}\frac{r}{\sqrt{2}}-\cos k\sin^{2}\frac{r}{\sqrt{2}})^{2}}}\label{eq:}\end{eqnarray}
873:
874:
875: We find that the residue of the integrand is $\frac{4}{\sin r/\sqrt{2}}$,
876: and hence has a $\frac{1}{k}$ singularity. We can now conclude that
877: in the limit as $N\rightarrow\infty$, the value of $\left.\frac{\partial^{2}|\Omega^{(1)}|}{\partial\phi^{2}}\right|{}_{\phi=\pm\frac{\pi}{4},\pm\frac{3\pi}{4}}$
878: will be infinite, with a logarithmic singularity.
879:
880: Alternatively, following Ref. \cite{bunder_mckenzie} and expressing
881: $E_{k}$ as\\
882: \begin{eqnarray}
883: E_{k} & = & \cos^{-1}\eta_{k}\label{omega_simple}\\
884: & = & \cos^{-1}(\cos^{2}\frac{k}{2}\cos(\theta+\chi)+\sin^{2}\frac{k}{2}\cos(\theta-\chi))\nonumber \end{eqnarray}
885:
886:
887: Bunder and McKenzie\cite{bunder_mckenzie} note that for some wave
888: vector $k$, $E_{k}=0$, corresponding to a vanishing energy gap in
889: the system. When $E_{k}$ is expressed as Eq (\ref{omega_simple}),
890: it is clear that there will be no energy gap for $k=0$ if $\theta=-\chi$
891: and for $k=\pi$ if $\theta=\chi$. Without loss of generality, we
892: can consider the $k=0$ case, as the other is symmetric. Since the
893: relevant excitations are at $k\simeq0$, we may use Eq (\ref{omega_simple})
894: to expand $E_{k}^{2}$ as a series in $k$
895:
896: \begin{eqnarray}
897: E_{k}^{2} & \simeq & (\theta+\chi)^{2}+k^{2}\frac{(\theta+\chi)(\cos\theta+\chi-\cos\theta-\chi)}{2\sin(\theta+\chi)}\label{eq:}\\
898: & \equiv & \xi^{2}+k^{2}\zeta^{2}\nonumber \end{eqnarray}
899: and the ground state will have energy
900:
901: \begin{eqnarray}
902: E_{0} & = & \frac{-1}{2\pi}\int_{-k_{c}}^{k_{c}}dk\sqrt{E_{k}^{2}}\label{eq:}\end{eqnarray}
903: where $k_{c}$ is a cutoff wavevector. While analytical solutions
904: are possible, it is unenlightening to solve this problem. Instead,
905: we can set out to determine the behaviour of the energy with respect
906: to a variable $\xi^{2}$ by using the indefinite integral
907:
908: \begin{eqnarray}
909: E_{0} & = & -\int d\xi^{2}\frac{\partial}{\partial\xi^{2}}E_{0}\label{eq:}\end{eqnarray}
910:
911:
912: Carrying this out we obtain
913:
914: \begin{eqnarray}
915: E_{0} & = & \int d\xi^{2}\frac{\partial}{\partial\xi^{2}}E_{0}\label{eq:}\\
916: & = & \frac{-1}{4\pi}\int d\xi^{2}\int_{-k_{c}}^{k_{c}}dk\frac{1}{\sqrt{\xi^{2}+k^{2}\zeta^{2}}}\nonumber \\
917: & = & \frac{-1}{4\pi}\int d\xi^{2}\frac{-2}{\zeta}(1-\ln\frac{\xi}{2\zeta k_{c}})\nonumber \\
918: & = & \frac{-\xi^{2}}{2\pi\zeta}(1-2\ln\frac{\xi}{2\zeta k_{c}})\nonumber \end{eqnarray}
919:
920:
921: Thus we confirm the logarithmic nature of this singularity, and find
922: that $E_{0}\sim-\xi^{2-\alpha}$ where we have the value of the critical
923: exponent $\alpha=0^{+}$. This is the same behaviour as that found
924: in the transverse Ising model\cite{bunder_mckenzie}, which we expect
925: by their inclusion in the same universality class.
926:
927: One can see the behaviour of ${\frac{\partial^{2}\Omega^{(1)}}{\partial\phi^{2}}}$
928: with respect to $\phi$ in Fig. \ref{comparing} for $N=200$ and
929: $r=1.9$. There exists a quantum phase transition at $\phi=\pm{\frac{\pi}{4}},\pm{\frac{3\pi}{4}}$
930: as evidenced by the singularity in ${\frac{\partial^{2}\Omega^{(1)}}{\partial\phi^{2}}}$.
931: This numerical modeling corresponds to our theoretical expectation
932: for the positions of the phase transitions. For a finite set of qubits,
933: one can clearly see the peak in the second derivative of the energy
934: with respect to $\phi$ in Fig. \ref{6q_d2phase}.
935:
936: %
937: \begin{figure}
938: \includegraphics[%
939: width=3.04414in]{comparing-mm-ti.ps}
940:
941:
942: \caption{\label{comparing}The second derivative of the phase of the eigenvalues
943: for the model Hamiltonian $\bar{H}$ (solid), based on Eq. (\ref{u_final}),
944: and for the transverse Ising Hamiltonian, $H_{Ising}$ (dashed), as
945: a function of $\phi=\tan^{-1}\frac{\chi}{\theta}$. Note that both
946: show singularities at $\theta=\pm\chi$. $\left|\theta^{2}+\chi^{2}\right|=r$
947: was chosen to be $1.9$ so as to highlight the differences between
948: the plots. For smaller $\theta$ and $\chi$, the commutator between
949: $H_{\theta}$ and $H_{\chi}$ becomes small, and the model becomes
950: asymptotically closer to the Ising model. Fig. \ref{a_coeff} shows
951: that the smaller $\theta$ and $\chi$, the faster the model's non-neighbour
952: parameters decay.}
953: \end{figure}
954:
955:
956: %
957: \begin{figure}
958: \includegraphics[%
959: width=3.04414in]{d2edphi_6.ps}
960:
961:
962: \caption{\label{6q_d2phase}The second derivative of the ground state energy
963: of a 6 qubit model as a function of $\phi=\tan^{-1}\frac{\chi}{\theta}$,
964: for $r=1.9$. The maximum value is attained at $\phi=\frac{\pi}{4}$,
965: the value at which a quantum phase transition occurs in the thermodynamic
966: limit. While we see a strong maximum, in the thermodynamic limit,
967: we expect to see a singularity.}
968: \end{figure}
969:
970:
971:
972: \subsection{Entanglement}
973:
974: It has recently been shown that entanglement scales near a quantum
975: critical point \cite{Osborne,Osterloh}. Quantum phase transitions
976: are driven by quantum fluctuations\cite{Sachdev}, and entanglement
977: is a natural manner for non-local effects to manifest themselves.
978: As entanglement is a physical resource, it may be directly measured,
979: by a number of schemes\cite{measuring1,measuring2}.
980:
981: We denote the nearest neighbour entanglement in the ground state by
982: ${\cal E}$. In the transverse Ising model, we see that the derivative
983: of entanglement with respect to $\phi$, $\frac{\partial{\cal E}}{\partial\phi}$,
984: near criticality to scale as a function of $\left|\phi-\phi_{c}\right|$.
985: Since we are in the same universality class, we expect to see identical
986: behaviour in this model. However, experimentally, isolating the ground
987: state of the system to observe this may be very difficult.
988:
989: %
990: \begin{figure}
991: \includegraphics[%
992: width=3.04414in]{entanglement_6.ps}
993:
994:
995: \caption{\label{6q_entanglement}The derivative of the nearest neighbour entanglement,
996: with respect to $\phi$ for a 6 qubit model. Note that this has a
997: maximum very close to the critical point $\phi=\frac{\pi}{4}$. We
998: expect to see the nearest neighbour concurrence vary as $\log\left|\phi-\phi_{c}\right|$
999: in the thermodynamic limit.}
1000: \end{figure}
1001:
1002:
1003:
1004: \subsection{Spectroscopic Measurement of Eigenvalues}
1005:
1006: Experimentally, it is possible that an ion trap may be used to implement
1007: the unitary map of the form
1008:
1009: \[
1010: \left|\Psi\right\rangle \rightarrow e^{-i\chi\sum_{n=1}^{N}\sigma_{z}^{(n)}\sigma_{z}^{(n+1)}}\left|\Psi\right\rangle \equiv\left|\Psi'\right\rangle \]
1011: followed by another map of the form
1012:
1013: \[
1014: \left|\Psi'\right\rangle \rightarrow e^{-i\theta\sum_{n=1}^{N}\sigma_{x}^{(n)}}\left|\Psi'\right\rangle \equiv\left|\Psi''\right\rangle \]
1015:
1016:
1017: Let us introduce the notation
1018:
1019: \[
1020: \left|\Psi^{(m)}\right\rangle \equiv U^{m}\left|\Psi\right\rangle \]
1021: where $\left|\Psi^{(m)}\right\rangle $ is the state after we repeat
1022: this unitary map, $U$, $m$ times.
1023:
1024: The state $\left|\Psi\right\rangle $ can be decomposed as
1025:
1026: \begin{eqnarray}
1027: U^{m}\left|\Psi\right\rangle & = & \sum_{n}\ket{\phi_{n}}\bra{\phi_{n}}U^{m}\ket{\Psi}\label{eq:}\\
1028: & = & \sum_{n}\ket{\phi_{n}}\braket{\phi_{n}}{\Psi}e^{imE_{n}}\label{eq:}\end{eqnarray}
1029: where $\left|\phi_{n}\right\rangle $ are the eigenstates of $U$,
1030: with {}``energy'' $E_{n}$. Without loss of generality, we assume
1031: these are ordered with
1032:
1033: \[
1034: E_{0}<E_{1}<\ldots<E_{M-1}\]
1035: where $M=2^{N}$.
1036:
1037: We can proceed to measure $U^{m}\left|\Psi\right\rangle $ in some
1038: set of basis states, $\left|i\right\rangle $, which will typically
1039: be binary computational basis states. Hence, we can measure
1040:
1041: \begin{eqnarray}
1042: \left|\bra{i}U^{m}\ket{\Psi}\right|^{2} & = & \left|\sum_{n}\braket{i}{\phi_{n}}\braket{\phi_{n}}{\Psi}e^{imE_{n}}\right|^{2}\label{eq:}\end{eqnarray}
1043:
1044:
1045: \begin{equation}
1046: =\sum_{n,n'}\braket{i}{\phi_{n}}\braket{i}\phi_{n'}^{*}\braket{\phi_{n}}{\Psi}\braket{\phi_{n'}}\Psi^{*}e^{im(E_{n}-E_{n'})}\label{eq:}\end{equation}
1047:
1048:
1049: If we perform a Fourier transform of $\left|\bra{i}U^{m}\ket{\Psi}\right|^{2}$
1050: over $m$, we expect to see peaks around the allowable transition
1051: energies $E_{n}-E_{n'}$. A numerical simulation of this is shown
1052: in Fig. \ref{q4_fourier} for 4 qubits. Let us define
1053:
1054: \begin{equation}
1055: F_{n}=\sum_{m=0}^{n-1}e^{\frac{i2\pi m}{n}}\left|\bra{i}U^{m}\ket{\Psi}\right|^{2}\label{eq:}\end{equation}
1056: to be these Fourier components.
1057:
1058: Since we are considering Unitary maps, and not Hamiltonians, we can
1059: only determine the eigenvalues of $U$ to within an additive constant
1060: of $2\pi$. Hence, in order that the Fourier components are not aliased
1061: (that the energy levels do not {}``wrap around'' on themselves),
1062: we require the ground state to have an energy $E_{0}>-\pi$, and the
1063: highest excited state to have an energy $E_{2^{N}}<\pi$. Since the
1064: energy is a function which scales with $O(N,\theta,\chi)$, we require
1065: the condition $\max(\left|\theta\right|,\left|\chi\right|)<\frac{k_{int}}{N}$
1066: where $k_{int}$ is $O(1)$. Keeping $\theta$ and $\chi$ small in
1067: this manner will ensure that the energies will be resolvable uniquely
1068: by the Fourier transform.
1069:
1070: If we have a given set of energy eigenvalues, $\left\{ \bar{E}_{0},\bar{E}_{1}\ldots\bar{E}_{M-1}\right\} $,
1071: we can form the set of energy differences, $\left\{ \bar{E}_{i,j}\equiv\bar{E}_{i}-\bar{E}_{j}\right\} $.
1072: We can then calculate the Fourier transform of these differences,
1073: and compare our measured spectrum with the calculated spectrum. If
1074: we have $n\gg M$, then the problem is over determined, and we can
1075: apply a least-squares method to reconstruct the original energy spectrum
1076: (to within an additive constant, and global sign change). We may apply
1077: the Levenberg Marquardt algorithm\cite{leven,MARQUARDT,LM} to perform
1078: this reconstruction in polynomial time with an initial guess at the
1079: set of energy eigenvalues.
1080:
1081: %
1082: \begin{figure}
1083: \includegraphics[%
1084: width=3.04414in]{eigen6_e.ps}
1085:
1086:
1087: \caption{\label{q6_eigens}Several of the lowest energy eigenvalues for the
1088: 6 qubit model. One can clearly see the excitation gap closing as $\phi$
1089: approaches $\frac{\pi}{4}$. In the thermodynamic limit, we expect
1090: the gap to be identically zero at $\phi=\frac{\pi}{4}$. However,
1091: we can still observe the energy gap, $\Delta$, behaving as $\Delta\sim\left|\phi-\phi_{c}\right|^{\gamma}$
1092: with some higher order corrections.}
1093: \end{figure}
1094:
1095:
1096: %
1097: \begin{figure}
1098: \includegraphics[%
1099: width=3.04414in]{4bit-fourier.ps}
1100:
1101:
1102: \caption{\label{q4_fourier}The Fourier transform of $\left|\bra{i}U^{m}\ket{\Psi}\right|^{2}$,$F_{m,2048}$,
1103: is shown as a vertical density, as function of the horizontal co-ordinate
1104: $\theta$, for a fixed $\chi=0.2$, and 4 qubits. 4 qubits are chosen
1105: so as to provide a complex, but not confusing diagram. The white bands
1106: indicate a large Fourier component. The superimposed grey lines show
1107: all energy differences - note that some of these are disallowed. In
1108: this case, $2048$ samples are used in the Fourier series, and simulations
1109: are taken in steps of $0.01$ in $\theta$. From this diagram, we
1110: can see the energy gap between the ground state and first excited
1111: state approaching zero. We can also see a level crossing, where one
1112: of the grey lines is reflected through the origin at $\theta\simeq0.25$. }
1113: \end{figure}
1114:
1115:
1116: One could change the value of $\frac{\theta}{\chi}$ over many experiments
1117: to tune the system through the critical coupling. In the thermodynamic
1118: limit, the energy gap to the first excited state would vanish at criticality,
1119: but we see in Fig. \ref{q6_eigens} that the condition is not strictly
1120: met for a finite number of qubits.
1121:
1122: One can use this to show that the energy gap, $\Delta$, for the excitation
1123: from the ground to first excited state obeys the relation
1124:
1125: \begin{equation}
1126: \Delta\sim\left|\phi-\phi_{c}\right|^{\gamma}\label{eq:}\end{equation}
1127: with $\gamma=1$, as we expect from a standard treatment of the transverse
1128: Ising problem\cite{Sachdev}.
1129:
1130:
1131: \subsubsection{Controlled-U Spectroscopy}
1132:
1133: The above method requires knowledge of the approximate values of $\left|\bra{i}U^{m}\ket{\Psi}\right|^{2}$,
1134: which means that a measurement with result $\ket{i}$ must be achieved
1135: multiple times. In work by Miquel \emph{et al.}\cite{controlled_u_spec},
1136: it has been shown that spectroscopy can be achieved much more easily
1137: by implementing a controlled-U gate, and measuring only a single qubit\cite{one_bit}.
1138: We can achieve this by using an ancillary qubit, and express the controlled-U
1139: operation as
1140:
1141: \[
1142: CU:\bra{i}\otimes\bra{\Psi}\rightarrow\bra{i}\otimes U^{i}\bra{\Psi}\]
1143: where $\bra{i}$ can be either $\bra{0}$, which takes $\bra{\Psi}$
1144: to $\bra{\Psi}$, or $\bra{1}$, which takes $\bra{\Psi}$ to \textbf{$U\bra{\Psi}$.}
1145:
1146: If we do a weak measurement on the control bit, we yield the result
1147:
1148: \begin{eqnarray*}
1149: \left\langle \sigma_{z}\right\rangle =\Re\left[Tr(U\rho)\right] & , & \left\langle \sigma_{y}\right\rangle =\Im\left[Tr(U\rho)\right]\end{eqnarray*}
1150: where $\rho$ is the density matrix corresponding to the state has
1151: been prepared in. If we prepare it in the mixed state given by $\rho=I/2^{n}$
1152: where $I$ is the identity operator, we yield $\left\langle \sigma_{z}\right\rangle =\Re\left[Tr(U)\right]/N$,
1153: which is proportional to the sum of the eigenvalues of $U$. If we
1154: repeat this for $U^{m}$ for a variety of $m$, we can use the method
1155: above to reconstruct the energy level diagram.
1156:
1157: Further, Miquel \emph{et al.}\cite{controlled_u_spec} propose a scheme
1158: using the quantum Fourier transform to probe specific regions of the
1159: spectrum of the eigenvalues of $U$. This is achieved by introducing
1160: an effective time scale into $U$, and exploiting the conjugacy of
1161: energy and time.
1162:
1163:
1164: \subsection{Phase Estimation Algorithm}
1165:
1166: In work done by Abrams and Lloyd\cite{abramslloyd}, and further explored
1167: in an ion trap context by Travaglione and Milburn\cite{eigenest},
1168: it has been shown that it is possible to estimate the eigenvalues
1169: associated with any unitary transformation, $U$. These correspond
1170: directly to the energy eigenvalues of the equivalent Hamiltonian,
1171: which we are interested in. The scheme also yields an approximate
1172: eigenvector with high probability.
1173:
1174: Starting with a mixed index state $\ket j_{I}$, and the state of
1175: the target system $\ket{\Psi}$, we perform the transformation
1176:
1177: \[
1178: \Lambda(U):\ket j_{I}\ket\Psi_{T}\rightarrow\ket j_{I}\otimes U^{j}\ket\Psi_{T}\]
1179: followed by a Fourier transformation on the index register. Measuring
1180: the index register will then yield, with high probability, an approximate
1181: eigenvector of $U$ in the target state, and information about the
1182: phase of the eigenvalue of $U$ in the index register.
1183:
1184: Note, however, that use is made of an index register, which is at
1185: least the same size as the system of interest. This makes it a much
1186: more difficult problem to conquer experimentally, as a system twice
1187: as big will be much more prone to decoherence. In ion trap implementations,
1188: trapping twice as many ions will also be more difficult. While this
1189: technique is superior to the spectroscopic measurements suggested
1190: in the previous section, scalability issues may keep it from being
1191: experimentally feasible for some time.
1192:
1193:
1194: \section{Conclusion}
1195:
1196: We have presented a number of key ideas which will drive our search
1197: for a quantum phase transition in a system which is implementable
1198: on an ion-trap quantum computer in a natural way.
1199:
1200: We have taken the Feynman thesis and turned it around, to ask what
1201: might happen if we have some implementable unitary transformation.
1202: The two fields which will help to answer this question have been introduced
1203: - namely, the ion-trap quantum architecture which will provide our
1204: unitary transformation, and the tools of renormalization group theory.
1205: In this paper, we have shown that the Hamiltonian corresponding to
1206: a separated Ising map belongs to the same universality class as the
1207: transverse Ising model. Further, the map presented here is realisable
1208: in a very natural way on an ion-trap quantum computing architecture.
1209:
1210: We have also suggested some experimental signatures , including ground
1211: state energy and entanglement, and spectroscopic information which
1212: may lead to the reconstruction of the energy spectrum.
1213:
1214: After completion of this work, we became aware of some other work
1215: on simulating quantum phase transitions in ion traps\cite{similar}
1216:
1217: \begin{acknowledgments}
1218: JPB would like to thank Ben Powell for useful discussions. This work
1219: was supported by the Australian Research Council.
1220: \end{acknowledgments}
1221: \bibliographystyle{apsrev}
1222: \bibliography{article1}
1223:
1224: \end{document}
1225: