quant-ph0512255/ccc.tex
1: \chapter{Coherent classical communication}\label{chap:ccc}
2: \section{Introduction and definition}
3: 
4: One of the main differences between classical and quantum Shannon
5: theory is the number of irreversible, but optimal, resource
6: transformations that exist in quantum Shannon theory.  The highest
7: rate that ebits or cbits can be created from qubits is one-for-one:
8: $\qtq\geq \qq$ and $\qtq\geq \ctc$.  But the best way to
9: create qubits from cbits and ebits is teleportation: $2\ctc+\qq\geq
10: \qtq$.  These protocols are all asymptotically 
11: optimal---for example, the classical communication requirement of
12: teleportation cannot be decreased even if entanglement is free---but
13: composing them is extremely wasteful: $3\qtq \geq 2\ctc + \qq
14: \geq \qtq$.  This sort of irreversibility represents one of the
15: main challenges of quantum information theory: resources may be
16: qualitatively equivalent but quantitatively incomparable.
17: 
18: In this chapter we will introduce a new primitive resource: the {\em
19: coherent bit} or {\em cobit}.  To emphasize its connection with
20: classical communication, we denote the asymptotic resource (defined
21: below) by $\cof$.\footnote{Other work\cite{DHW05,Devetak05a} 
22: uses $[q\ra qq]$ to denote cobits, in order to emphasize their central
23: place among isometries from $A$ to $AB$.}
24: Coherent classical communication will
25: simplify and improve a number of topics in quantum Shannon theory:
26: \bi
27: \item We will find that coherently decoupled cbits can be described
28: more simply and naturally as cobits.
29: \item Replacing coherently decoupled cbits with cobits will make many
30: resource 
31: transformations reversible.  In particular, teleportation and
32: super-dense coding become each other's inverses, a result previously
33: only known when unlimited entanglement is allowed.
34: \item More generally, we find that many forms of irreversibility in
35: quantum Shannon theory are equivalent to the simple map $\cof\geq
36: \ctc$. 
37: \item We will expand upon \prop{U-CgE} to 
38: precisely explain how cbits are more powerful when they are sent
39: through unitary means.  This has a number of consequences for unitary
40: gate capacities.
41: \item In the next chapter, coherent classical communication will be
42: used to relate many of the different protocols in quantum
43: Shannon theory, give simple proofs of some existing protocols and
44: create some entirely new protocols.  These will allow us to determine
45: two-dimensional tradeoff curves for the capacities of channels and
46: states to create or consume cbits, ebits and qubits.
47: \ei
48: 
49: Coherent classical communication  can be defined in
50: two ways, which we later show to be equivalent.  
51: 
52: \bi
53: \item {\em Explicit definition in terms of finite resources:} 
54: 
55: \sloppypar{Fix a basis for $\bbC^d$: $\{\ket{x}\}_{x=0}^{d-1}$.
56: First, we recall from 
57: \sect{distantlabs} the definitions of quantum and classical
58: communication: $\id_d = \sum_x
59: \ket{x}^{B} \bra{x}^{A'}$ (a perfect quantum
60: channel), $\bar{\id}_d = \sum_x \ket{x}^{B}\ket{x}^{E} \bra{x}^{A'}$
61: (a perfect classical channel in the QP formalism) and
62: $\bar{\Delta}_d = \sum_x \ket{x}^A\ket{x}^{B}\ket{x}^{E} \bra{x}^{A'}$
63: (the classical copying operation in the QP formalism).  Then we define
64: a perfect coherent channel as
65: \be \Delta_d = \sum_{x=0}^{d-1}
66: \ket{x}^A\ket{x}^{B} \bra{x}^{A'}.\ee
67: It can be thought of as a
68: purification of a cbit in which Alice controls the environment, as a
69: sort of quantum analogue to a feedback channel.
70: The asymptotic resource is then given by $\cof := \<\Delta_2\>$.}
71: 
72: \item {\em Operational definition as an asymptotic resource:}
73: We can also define a cobit as a cbit sent through unitary, or more
74: generally isometric, means.  The approximate version of this statement
75: is that whenever a protocol creates coherently decoupled cbits
76: (cf.~\defn{coh-decoupling-output}), then a modified
77: version of the protocol will create cobits.  Later we will prove a precise
78: form of this statement,  known as ``Rule
79: O,'' because it describes how output cbits should be made coherent.
80: 
81: When $C$ {\em input cbits} are coherently decoupled
82: (cf.~\defn{coh-decoupling-input}) we instead find that replacing them
83: with $C$ cobits results in $C$ extra ebits being generated in the
84: output.  This input rule is known as ``Rule I.''  Both rules are
85: proved in \sect{ccc-proofs}.
86: 
87: The canonical example of coherent decoupling is when cbits are sent
88: using a unitary gate.  In \thm{bidi-ccc}, we show that cbits sent
89: through unitary means can indeed be coherently decoupled, and thereby
90: turned into cobits.
91: \ei
92: 
93: The rest of the chapter is organized as follows.
94: \begin{description}
95: \item[\sect{ccc-source}] will give some simple examples of how cobits
96: can be obtained.
97: \item[\sect{ccc-use}] will then describe how to use coherent
98: classical communication to make quantum protocols reversible and more
99: efficient.  It will conclude with a precise statement of Rules I and O.
100: \item[\sect{ccc-apps}] will apply these general principles to remote
101: state preparation\cite{BHLSW03}, which leads to new 
102: protocols for super-dense coding of quantum state\cite{HHL03} as well
103: as many new results for unitary gate capacities.
104: \item[\sect{ccc-proofs}] collects some of the longer proofs from the
105: chapter, in order to avoid interrupting the exposition of the rest of
106: the chapter. 
107: \item[\sect{ccc-discuss}] concludes with a brief discussion.
108: \end{description}
109: 
110: {\em Bibliographical note:} Most of this chapter is based on
111: \cite{Har03}, though in \sect{ccc-proofs} the proofs of Rules I and O
112: are from 
113: \cite{DHW05} (joint work with Igor Devetak and Andreas Winter).
114: and the proof of \thm{bidi-ccc} is from
115: \cite{HL04} (joint work with Debbie Leung).
116: 
117: \section{Sources of coherent classical communication}
118: \label{sec:ccc-source}
119: 
120: Qubits and cbits arise naturally from noiseless and dephasing channels
121: respectively, and can be obtained from any noisy channel by
122: appropriate coding \cite{Holevo98,SW97,Lloyd96,Shor02,Devetak03}.
123: Similarly, we will show both a 
124: natural primitive yielding coherent bits and a coding theorem that
125: can generate coherent bits from a broad class of unitary operations.
126: 
127: The simplest way to send a coherent message is by modifying
128: super-dense coding (SD).  In SD, Alice and Bob begin
129: with $\ket{\Phi_2}$ and want to use $\id_2$ to send a two bit
130: message $a_1a_2$ from Alice to Bob.  Alice encodes her message by
131: applying $Z^{a_1}X^{a_2}$ to her half of 
132: $\ket{\Phi_2}$ and then sending it to Bob, who decodes by applying
133: $(H\otimes I)\textsc{cnot}$ to the state, obtaining
134: %
135: $$(H\otimes I)\textsc{cnot}(Z^{a_1}X^{a_2}\otimes I)\ket{\Phi_2}
136: =\ket{a_1}\ket{a_2}$$
137: %
138: Now modify this protocol so that Alice starts with a quantum state
139: $\ket{a_1a_2}$ and applies $Z^{a_1}X^{a_2}$ to her half of
140: $\ket{\Phi_2}$ conditioned on her quantum input.  After she sends her
141: qubit and Bob decodes, they will be left with the state
142: $\ket{a_1a_2}^A\ket{a_1a_2}^B$.  Thus, 
143: \be \qtq + \qq \reduction 2 \cof \label{eq:SDC-cc}\ee
144: 
145: In fact, any unitary operation capable of classical communication is
146: also capable of an equal amount of coherent classical communication,
147: though in general this only holds asymptotically.  The following
148: theorem gives a general prescription for obtaining coherent
149: communication and proves part of the equivalence of the two
150: definitions of cobits given in the introduction.
151: 
152: \begin{theorem}\label{thm:bidi-ccc}
153: For any bipartite unitary or isometry $U$, if
154: \be
155: 	\<U\>  \geqslant  C_1 \ctc + C_2 \cbc + E \qq
156: \label{eq:cbit-toff} 
157: \ee
158: for $C_1,C_2\geq 0$ and $E\in\bbR$ then
159: \be
160: 	\<U\>  \geqslant  C_1 \cof + C_2 \cob + E \qq
161: \label{eq:cobit-toff} 
162: \ee
163: \end{theorem}
164: If we define $\CoCoE(U)=\{(C_1,C_2,E) : \<U\>  \geqslant  C_1 \cof + C_2
165: \cob + E \qq\}$, then this theorem states that $\CCE(U)$ and
166: $\CoCoE(U)$ coincide on the quadrant $C_1,C_2\geq 0$.
167: 
168: Here we will prove only the case where $C_2=0$, deferring the full
169: bidirectional proof to \sect{ccc-proofs}.  By appropriate coding (as
170: in \cite{BS03b}), we can reduce the one-way case of \thm{bidi-ccc} to
171: the following coherent analogue of HSW coding.
172: 
173: \begin{lemma}[Coherent HSW]\label{lemma:HSW}
174: Given a PP ensemble of bipartite pure states
175: \be \ket{\cE} = \sum_{x\in\cX}
176: \sqrt{p_x}\ket{x}^R\ket{x}^{X_A}\ket{\psi_x}^{AB}\ee
177:  and an isometry 
178: \be U_\cE = \sum_x \oprod{x}^{X_A} \ot \ket{\psi_x}^{AB}\ee
179: then 
180: \be \<U_\cE : \cE^{X_A}\> \geq I(X_A;B)_\cE\cof + H(B|X_A)\qq.\ee
181: 
182: \end{lemma}
183: 
184: \begin{proof}
185: A slightly modified form of HSW coding (e.g. \cite{Devetak03}) holds
186: that for any $\delta>0,\epsilon>0$ and 
187: every $n$ sufficiently large there exists a code $\cC\subset\cS^n$
188: with $|\cC|=\exp(n(I(X_A;B)_\cE-\delta))$, a decoding POVM
189: $\{D_{c^n}\}_{c^n\in\cC}$ with error $<\epsilon$ and a type $q$ with
190: $\|p-q\|_1\leqslant |\cX|/n$ such that every codeword $c^n:=c_1\ldots
191: c_n\in\cC$ (corresponding to the state $\ket{\psi_{c^n}}^{AB}:=
192: \ket{\psi_{c_1}}^{A_1B_1}\cdots \ket{\psi_{c_n}}^{A_nB_n}$)
193: has type $q$ (i.e. $\forall x, |\{c_j=x\}|=nq_x$).  By
194: error $<\epsilon$, we mean that for any $c^n\in\cC$,
195: $\bra{\psi_{c^n}}(I\otimes D_{c})\ket{\psi_{c^n}} > 1-\epsilon$.
196: 
197: Using Neumark's Theorem\cite{Peres93}, Bob can make his decoding POVM into
198: a unitary operation $U_D$ defined by $U_D\ket{0}\ket{\phi} = \sum_{c^n}
199: \ket{c^n} \sqrt{D_{c^n}}\ket{\phi}$.  Applying this to his half of a
200: codeword $\ket{\psi_{c^n}}$ will
201: yield a state within $\epsilon$ of $\ket{c^n}\ket{\psi_{c^n}}$, since
202: measurements with nearly certain outcomes cause almost no
203: disturbance\cite{Winter99}.
204: 
205: The communication strategy begins by applying $U_\cE$ to
206: $\ket{c^n}_{X_A}$ to obtain $\ket{c^n}^{X_A}\ket{\psi_{c^n}}^{AB}$.
207: Bob then decodes unitarily with $U_D$ to yield a state within
208: $\epsilon$ of $\ket{c^n}^{X_A}\ket{c^n}^{X_B}\ket{\psi_{c^n}}^{AB}$.
209: Since $c^n$ is of type $q$, Alice and Bob can coherently permute the
210: states of $\ket{\psi_{c^n}}$ to obtain a state within $\epsilon$ of
211: $\ket{c^n}_{X_A}\ket{c^n}_{X_B}\ket{\psi_1}^{\otimes nq_1}\cdots
212: \ket{\psi_{|\cX|}}^{\otimes nq_{|\cX|}}$.  Then they can apply
213: entanglement concentration\cite{BBPS96} to $\ket{\psi_1}^{\otimes
214: nq_1}\cdots \ket{\psi_{|\cX|}}^{\otimes nq_{|\cX|}}$ to obtain
215: $\approx nH(B|X_A)_\cE$ ebits without disturbing the coherent message
216: $\ket{c^n}_{X_A}\ket{c}_{X_B}$.
217: \end{proof}
218: This will be partially superseded by the full proof of \thm{bidi-ccc}.
219: However, it is worth appreciating the 
220: key ideas of the proof---making measurements coherent via Neumark's
221: Theorem and finding a way to decouple ancillas shared by Alice and
222: Bob---as they will appear again in the later proofs, but surrounded by
223: more mathematical details.
224: 
225: There are many cases in which no ancillas are produced, so we do not
226: need the assumptions of Lemma~\ref{lemma:HSW} that communication 
227:  occurs in large blocks.  For example, a {\sc cnot} can
228: transmit one coherent bit from Alice to Bob or one coherent bit from
229: Bob to Alice.  Recall the protocol given in \eq{e-asstd-cnot} for
230: $\cnot + \qq\reduction \ctc+\cbc$:
231: $ (Z^aH\otimes I)\textsc{cnot}(X^a\otimes Z^b)\ket{\Phi_2}^{AB} =
232: \ket{b}^A\ket{a}^B$. 
233: This can be made coherent by conditioning
234: the encoding on a quantum register $\ket{a}^{A'}\ket{b}^{B'}$, so that
235: %
236: \be\textsc{cnot} + \qq \reduction \cof + \cob
237: \label{eq:CNOT-power1}\ee
238: 
239: \section{Rules for using coherent classical communication}
240: \label{sec:ccc-use}
241: 
242: By discarding her state after sending it, Alice can convert coherent
243: communication into classical communication, so
244: $\cof\geqslant \ctc$.  Alice can also generate entanglement by
245: inputting a superposition of messages (as in \prop{U-CgE}), so
246: $\cof\geqslant \qq$.  The true power of coherent
247: communication comes from performing both tasks---classical communication
248: and entanglement generation---simultaneously.  This is possible
249: whenever the classical message sent is coherently decoupled,
250: i.e. random and nearly independent 
251: of the other states at the end of the protocol.
252: 
253: Teleportation \cite{BBCJPW98} satisfies these conditions, and indeed a
254: coherent version has already been proposed in \cite{BBC98}.  Given an
255: unknown quantum state $\ket{\psi}^{A}$ and an EPR pair
256: $\ket{\Phi_2}^{AB}$, Alice begins coherent teleportation not by a Bell
257: measurement on her two qubits but by unitarily rotating the Bell basis
258: into the computational basis via a {\sc CNOT} and Hadamard gate.  This
259: yields the state $\frac{1}{2}\sum_{ij}\ket{ij}^AX^iZ^j\ket{\psi}^B$.
260: Using two coherent bits, Alice can send Bob a copy of her register to
261: obtain $\frac{1}{2}\sum_{ij}\ket{ij}^A\ket{ij}^BX^iZ^j\ket{\psi}^B$.
262: Bob's decoding step can now be made unitary, leaving the state
263: $(\ket{\Phi_2}^{AB})^{\otimes 2}\ket{\psi}^B$.  In terms of resources,
264: this can be summarized as: $2\cof + \qq \reduction \qtq + 2\qq$.
265: Canceling the ebits on both sides (possible since $\cof\geq \qq$) gives
266: $2\cof \geq \qtq + \qq$.
267: Combining this relation with \eq{SDC-cc} yields the
268: equality\footnote{Our use of the Cancellation Lemma means that this
269: equality is only asymptotically valid.  \mscite{vanEnk05} proves a
270: single-shot version of this equality, but it requires that the two
271: cobits be applied in series, with local unitary operations in between.}
272: %
273: \be 2\cof = \qtq + \qq\label{eq:cc-equality}.\ee
274: %
275: This has two important implications.  First, teleportation and
276: super-dense coding are reversible so long as all of the classical
277: communication is left coherent.  Second, cobits are equivalent, as
278: resources, to the existing resources of qubits and ebits.  This means
279: that we don't need to calculate quantities such as the cobit capacity
280: of a quantum channel; coherent communication introduces a new tool for
281: solving old problems in quantum Shannon theory, and is not directly a
282: source of new problems.
283: 
284: Another protocol that can be made coherent is Gottesman's
285: method\cite{Got98} for simulating a distributed CNOT using one
286: ebit and one cbit in either direction.  At first
287: glance, this appears completely irreversible, since a CNOT can be used
288: to send one cbit forward or backwards, or to create one ebit, but no
289: more than one of these at a time.
290: 
291: Using coherent bits as inputs, though, allows the recovery of 2 ebits
292: at the end of the protocol, so $\cof + \cob + \qq \reduction
293: \<\cnot\> + 2\qq$, or using entanglement catalytically,
294: $\cof + \cob \geq \<\cnot\> + \qq$.  Combined with
295: \eq{e-asstd-cnot}, this yields another equality:
296: $$\<\cnot\>+\qq = \cof + \cob.$$
297: Another useful bipartite unitary gate is \textsc{swap}, which we
298: recall is equivalent to $\qtq + \qbq$.  Applying
299: \eq{cc-equality} then yields
300: $$2\<\cnot\>=1\<\swap\>$$
301: which explains the similar communication and entanglement capacities
302: for these gates found in the last chapter.  Previously, the most efficient
303: methods known to transform between these gates gave $3\<\cnot\>\geq
304: 1\<\swap\> \geq 1\<\cnot\>$.
305: 
306: A similar argument can be applied to \dcnot.  Since $\<\dcnot\> + 2\qq
307: \geq 2\ctc + 2\cbc$, it follows (from \thm{bidi-ccc} or direct
308: examination) that $\<\dcnot\> + 2\qq \geq 2\cof + 2\cob$ and
309: that $\<\dcnot\> \geq \qtq + \qbq = \<\swap\>$.  Combining
310: this with \prop{U-TP-bound}, we find that $\<\dcnot\>=\<\swap\>$,
311: a surprising fact in light of the observation of \cite{HVC02} that
312: \dcnot\ is easier for some nonlocal Hamiltonians to simulate than
313: \swap.  In fact, by the same argument, any gate in $\cU_{d\times d}$
314: with $C_+^E(U)=4\log d$ must be equivalent to the $d\times d$ \swap\
315: gate.
316: 
317: The above examples give the flavor of when classical communication can
318: be replaced by coherent communication (i.e. ``made coherent.'')  In
319: general, we require that the classical message be (almost) uniformly
320: random and (almost) coherently decoupled from all other systems,
321: including the environment.  This leads us to two general rules regarding making
322: classical communication coherent.  When coherently-decoupled cbits are
323: in the input to a protocol, Rule I (``input'') says that replacing them with
324: cobits not only performs the protocol, but also has the side 
325: effect of generating entanglement.  Rule O (``output'') is simpler; it
326: says that if a protocol outputs coherently-decoupled cbits, then it
327: can be modified to instead output cobits.  
328: Once coherently decoupled cbits are replaced
329: with cobits we can then use \eq{cc-equality} to in turn replace
330: cobits with qubits and ebits.  Thus, while cobits are conceptually
331: useful, we generally start and finish with protocols involving the
332: standard resources of cbits, ebits and qubits.
333: 
334: Below we give formal statements of rules I and O, deferring their
335: proofs till the end of the chapter.
336: %We shall be working in the CP picture.
337: 
338: \begin{theorem}[Rule I]
339: If, for some quantum resources $\alpha, \beta \in {\cR}$,
340: $$
341: \alpha + R \, [c \rightarrow c : \tau] \geq \beta 
342: $$
343: and the classical resource $ R \, \ctctau$ is 
344: coherently decoupled  then
345: $$
346: \alpha + \frac{R}{2} \,[q \rightarrow q] \geq \beta + \frac{R}{2} \,[q \, q].
347: $$
348: \end{theorem}
349: 
350: {\bf Remark:} This can be thought of as a coherent version of
351: \lem{rcr}.
352: 
353: The idea behind the proof is that replacing $R [c\ra
354: c:\tau]$ with $R \coftau$ then gives an extra output of $R\qq$,
355: implying that $\alpha + R\coftau \geq \beta + R\qq$.  Then $\coftau$
356: can be replaced by $\half(\qtq + \qq)$ using \eq{cc-equality} and
357: \lem{noo}.  To prove this rigorously will require carefully accounting
358: for the errors, which we will do in \sect{ccc-proofs}.
359: 
360: \begin{theorem}[Rule O]
361: If, for some quantum resources $\alpha, \beta \in {\cR}$, 
362: $$
363: \alpha  \geq \beta + R \,[c \rightarrow c]
364: $$
365: and the classical resource is 
366: decoupled with respect to the RI then 
367: $$
368: \alpha  \geq \beta + \frac{R}{2} \, [q \, q] +  
369: \frac{R}{2} \, [q \rightarrow q].
370: $$
371: \end{theorem}
372: 
373: Here the proof is even simpler: $R\ctc$ in the output is replaced with
374: $R\cof$, which is equivalent to $\frac{R}{2}(\qtq+\qq)$.  Again, the
375: details are given in \sect{ccc-proofs}.
376: 
377: In the next chapter, we will show how Rules I and O can be used to
378: obtain a family of optimal protocols (and trade-off curves) for
379: generating cbits, ebits and qubits from noisy channels and states.
380: First, we show a simpler example of how a protocol can be made
381: coherent in the next section.
382: 
383: \section{Applications to remote state preparation and unitary gate
384: capacities}
385: %\sect{Coherent remote state preparation}
386: \label{sec:ccc-apps}
387: \subsection{Remote state preparation}
388: Remote state preparation (RSP) is the task of simulating a $\{c\ra
389: q\}$ channel, usually using cbits and ebits.  In this section, we show
390: how RSP can be made coherent, not only by applying Rule I to the input
391: cbits, but also by replacing the $\{c\ra q\}$ channel by a coherent
392: version that will preserve superpositions of inputs.  Finally, we will
393: use this coherent version of RSP to derive the capacity of a unitary
394: gate to send a classical message from Alice to Bob while
395: using/creating an arbitrary amount of entanglement.
396: 
397: Begin by recalling from \sect{known} our definition of RSP.  Let
398: $\cE=\sum_i p_i \oprod{i}^{X_A} \ot \oprod{\psi_i}^{AB}$ be an 
399: ensemble of bipartite states and  $\cN_\cE:\oprod{i}^{X_A}
400: \ra \oprod{i}^{X_A} \ot \oprod{\psi_i}^{AB}$ the $\{c\ra q\}$ channel
401: such that $\cN(\cE^{X_A}) = \cE$.  The main coding theorem of
402: RSP\cite{BHLSW03} states that
403: \be
404: I(X_A ; B)_\cE [c\ra c] + H(B)_\cE \qq \geq \<\cN_\cE : \cE^{X_A}\>.
405: \label{eq:RSP2}\ee
406: 
407: We will show that the input cbits in \eq{RSP2} are coherently
408: decoupled, so that according to Rule I, replacing them with cobits
409: will perform the protocol and return some entanglement at the same
410: time.  This reduces the entanglement cost to $H(B) - I(X_A;B) =
411: H(B|X_A)$, so that
412: \be
413: I({X_A}:B)_\cE \cof + H(B|X_A)_\cE \qq \geq \<\cN_\cE : \cE^{X_A}\>.
414: \label{eq:cobit-rsp}\ee
415: In fact, we can prove an even stronger statement, in which not only is
416: the input coherently 
417: decoupled, but there is a sense in which the output is as well.
418: Define a coherent analogue of $\cN_\cE$, which we call $U_\cE$, by
419: \be U_\cE = \sum_i \oprod{i}^{X_A} \ot \ket{\psi_i}^{AB}.\ee
420: We also replace the QP ensemble $\cE$ with the (PP formalism) pure
421: state $\ket{\cE}$ given by
422: \be \ket{\cE} = \sum_i \sqrt{p_i} \ket{i}^R \ket{i}^{X_A}
423: \ket{\psi_i}^{AB}.\ee
424: We will prove that
425: \be
426: I({X_A}:B)_\cE \cof + H(B|X_A)_\cE \qq \geq \<U_\cE : \cE^{X_A}\>.
427: \label{eq:coherent-rsp}\ee
428: Since $\<U_\cE : \cE^{X_A}\> \geq \<\cN_\cE : \cE^{X_A}\>$, this
429: RI implies \eq{cobit-rsp}; in particular, the presence of
430: the reference system $R$ ensures that $\cE^{X_A}$ is the same in both
431: cases, even if the $\ket{\psi_i}$ are not all orthogonal.
432: Proving \eq{coherent-rsp} will require careful examination of the
433: protocol from \cite{BHLSW03}, so we defer the details until
434: \sect{ccc-proofs}.
435: 
436: {\bf Remark:} An interesting special case is when $H(A)_\cE=0$, so
437: that Alice is preparing pure states in Bob's lab rather than entangled
438: states.  In this case, $H(B|X_A)_\cE=0$ and \eq{cobit-rsp} becomes
439: simply
440: \be H(B)_\cE \cof \geq \<U_\cE : \cE^{X_A}\>.\ee
441: Thus, if we say (following \cite{BHLSW03}) that \eq{RSP2} means that
442: ``1 cbit + 1 ebit $\geq$ 1 
443: remote qubit,'' then \eq{cobit-rsp} means that ``1 cobit $\geq$ 1
444: remote qubit.''  Here ``$n$ remote qubits'' mean the ability of Alice
445: to prepare an $n$-qubit state of her choice in Bob's lab, though we
446: cannot readily define an asymptotic resource corresponding to this
447: ability, since it would violate the quasi-i.i.d. condition
448: (\eq{asy-quasi-iid}).  Despite not being
449: formally defined as a resource, we can 
450: think of remote qubits as intermediate in strength between qubits and
451: cbits, just as cobits are; i.e. 1 qubit $\geq$ 1 remote qubit $\geq$ 1
452: cbit.  As resources intermediate between qubits and cbits, remote
453: qubits and cobits have complementary 
454: attributes: remote qubits share with qubits the ability to transmit
455: arbitrary pure states, though they cannot create entanglement, while
456: cobits can generate entanglement, but at first glance appear to only
457: be able to faithfully transmit the computational basis states to Bob.
458: Thus it is interesting that in fact 1 cobit $\geq$ 1 remote
459: qubit, and that (due to \cite{BHLSW03}) this map is optimal.
460: 
461: \eq{cobit-rsp} yields two other useful corollaries, which we state
462: in the informal language of remote qubits.
463: 
464: \begin{corollary}[RSP capacity of unitary gates]
465: If $U$ is a unitary gate or isometry with
466:  $\<U\>\geq C \ctc$  then $\<U\>\geq C \rqbsf$.
467: \end{corollary}
468: 
469: \begin{corollary}
470: {\bf (Super-dense coding of quantum states)}
471: \label{cor:sddc}
472: $\qtq + \qq \geq 2 \rqbsf$
473: \end{corollary}
474: More formally, we could say that if
475:  $H(B)_\cE \leq C$  for an ensemble $\cE$, then $\<U\> \geq \<U_\cE
476:  : \cE^{X_A}\>$, and similarly for \cor{sddc}.  We can also express
477:  \cor{sddc} entirely in terms of standard resources as
478: \be \half I(X_A ; B)_\cE \qtq  + \l(H(B)_\cE - \half I(X_A ; B)_\cE\r)
479:  \qq \geq \<U_\cE : \cE^{X_A}\>.\ee
480: Though this last expression is not particularly attractive, it turns
481: out to be optimal, and in fact to give rise to optimal trade-offs for
482: performing RSP with the three resources of cbits, ebits and
483: qubits\cite{AH03} (see also \cite{AHSW04} for a single-shot version of
484: the coding theorem).  We will find this pattern repeated many times in
485: the next chapter; by making existing protocols coherent and using
486: basic information-theoretic inequalities, we obtain a series of
487: optimal tradeoff curves.
488: 
489: Corollary~\ref{cor:sddc} was first proven directly in \cite{HHL03}
490: (see also \cite{AHSW04}) and
491: in fact, finding an alternate proof was the original motivation for
492: the idea of coherent classical communication.
493: 
494: {\em Coherent RSP:} Now, we explore the consequences of the stronger
495: version of coherent RSP in \eq{coherent-rsp}.  Just as RSP and HSW
496: coding reverse one another given free entanglement, coherent RSP
497: (\eq{coherent-rsp}) and coherent HSW coding (\lem{HSW}) reverse each
498: other, even taking entanglement into account.  Combining them gives
499: the powerful equality
500: \be
501: I({X_A}:B)_\cE \cof + H(B|X_A) \qq = \<U_\cE : \cE^{X_A}\>,
502: \label{eq:coh-RSP-HSW}\ee
503: which improves the original RSP-HSW duality in \eq{RSP-HSW-eq} by
504: eliminating the need for free 
505: entanglement.   This remarkable statement simultaneously implies
506: entanglement concentration, entanglement dilution, the HSW theorem and
507: remote state preparation and super-dense coding of entangled
508: states.\footnote{On the other hand, we had to use almost all of these
509: statements in order to prove the result!  Still it is nice to see them
510: all unified in a single powerful equation.  Also, recent work by
511: Devetak\cite{Devetak05a} further generalizes the equalities that can
512: be stated about isometries from $A$ to $AB$.}  
513: 
514: \subsection{One-way classical capacities of unitary gates}
515: \label{sec:U-Ce-cap}
516: Here we will use \eq{coh-RSP-HSW} to determine the 
517: capacity of a unitary gate $V$ to simultaneously send a classical message
518: and generate or consume entanglement at any finite rate.  The proof idea is
519: similar to one in \thm{U-CE-cap}; we will use the equivalence between
520: (coherent) ensembles and standard resources (cobits and ebits) to turn
521: a one-shot improvement in mutual information and expected entanglement
522: into an asymptotically efficient protocol.  Now that we have an
523: improved version of the duality between RSP and HSW coding, we obtain
524: a precise accounting of the amount of entanglement generated/consumed.
525: 
526: \begin{theorem}\label{thm:Ce-cap}
527: Define $\CE(V) := \{(C,E) : (C,0,E)\in\CCE(V)\}$ and 
528: \be \Delta_{I,E}(V) := \l\{ (C,E) : \exists \cE \st
529: I(X_A ; B)_{V(\cE)} - I(X_A ; B)_{\cE} \geq C
530: \text{ and }
531: H(B|X_A)_{V(\cE)} - H(B|X_A)_{\cE} \geq E\r\},\ee
532: where $\cE$ is an ensemble of bipartite pure states in $AB$ conditioned on a
533: classical register $X_A$.
534: 
535: Then $\CE(V)$ is equal to the closure of $\Delta_{I,E}(V)$.
536: \end{theorem}
537: Thus the asymptotic capacity using $-E$ ebits of assistance per use of
538: $V$ (or simultaneously outputting $E$ ebits) equals the largest
539: increase in mutual information possible with one use of $V$ if the
540: average entanglement decreases by no more than $-E$.  \thm{U-CE-cap}
541: proved this for $E=-\infty$ and our proof here is quite similar.  Note that
542: the statement of the theorem is the same whether we consider QP
543: ensembles $\cE$ or PP ensembles $\ket{\cE}$, though the proof
544: will use the coherent version of RSP in \eq{coherent-rsp}.
545: 
546: \begin{proof}
547: {\em  Coding theorem:}
548: Suppose there exists an ensemble $\cE$ with 
549: $C = I(X_A ; B)_{V(\cE)} - I(X_A ; B)_{\cE}$ and 
550: $E = H(B|X_A)_{V(\cE)} - H(B|X_A)_{\cE}$.
551: Then
552: %
553: \ben
554: \<V\> + \<U_\cE\> &\geq& \<U_{V(\cE)}\> \non\\
555: &\geq & I(X_A ; B)_{V(\cE)} \cof + H(B|X_A)_{V(\cE)} \qq
556: %\label{eq:CeU-HSW}
557: \\ &\geq & 
558:  \l(I(X_A ; B)_{V(\cE)} - I(X_A ; B)_{\cE}\r)\cof +
559: \l(H(B|X_A)_{V(\cE)} - H(B|X_A)_{\cE}\r)\qq + \,\<U_\cE\>
560: %\label{eq:CeU-RSP}
561: \een
562: Here the second RI used coherent HSW coding (Lemma~\ref{lemma:HSW}) and
563: the third RI used
564: coherent RSP (\eq{coherent-rsp}).  We now use the Cancellation Lemma
565: to show that 
566: $\<V\> \geq C\cof + E\qq$, implying that $(C,0,E)\in\CCE(V)$.
567: 
568: {\em Converse:}
569: We will actually prove a stronger result, in which Bob is allowed
570: unlimited classical communication to Alice.  Thus, we will show that
571: if $\<V\> + \infty \cbc \geq C \ctc + E \qq$, then there is a sequence
572: of ensembles $\{\tilde{\cE}_n\}$ with $(I(X_A ; B)_{V(\tilde{\cE}_n)}
573: - I(X_A ; B)_{\tilde{\cE}_n}, H(B|X_A)_{V(\tilde{\cE}_n)} -
574: H(B|X_A)_{\tilde{\cE}_n})$ converging to $(C,E)$ as $n\ra\infty$.
575: This will imply that $(C,E)$ is in the closure of $\Delta_{I,E}(V)$.
576: 
577: Let $Y:=Y_AY_B$ the cumulative record of all of Bob's classical
578: messages to Alice.  Using the QP formalism, we 
579: assume without loss of generality that Bob always transmits his full
580: measurement outcome (cf.~Section III of \cite{HL02}) so that Alice
581: and Bob always hold a pure state 
582: conditioned on $X_AY$; i.e. $H(AB|X_AY)=0$ and $H(A|X_AY)=H(B|X_AY) =
583: I(A\> BX_AY) = I(B \> AX_AY)$.  
584: 
585: First consider the case when $E>0$.  Fix a protocol which uses $V$ $n$
586: times to communicate $\geq n(C-\delta')$ bits and create $\geq
587: n(E-\delta')$ ebits with error $\leq \epsilon$.  They start with a
588: product state $\cE_0$ for which $I(X_A ; B)_{\cE_0}=0$ and $H(B |
589: X_A)_{\cE_0}=0$.  Denote their state immediately after $j$ uses of $V$
590: by ${\cE_j}$. (Without loss of generality, we assume that the $n$ uses
591: of $V$ are applied serially.)  Then by \lem{fano}, $I(X_A;B)_{\cE_n}
592: \geq n(C-\delta)$ and $H(B|X_AY)_{\cE_n} \geq n(E-\delta)$ where
593: $\delta=O(\delta+\epsilon)\ra 0$ as $n\ra \infty$.
594: 
595: Now define the ensemble $\tilde{\cE}_n=\smfrac{1}{n}
596: \sum_{j=1}^{n} \oprod{jj}^{Z_AZ_B} \ot V^\dag(\cE_j^{ABX_AY_AY_B})$.  We
597: think of $\hat{X}:=X_AY_AZ_A$ as the message variable and $\hat{B}:=
598: BY_BZ_B$ as representing Bob's system.  We will prove that
599: $I(\hat{X} ; \hat{B})_{V(\tilde{\cE}_n)}
600: - I(\hat{X} ; \hat{B})_{\tilde{\cE}_n} \geq C-\delta$ and that
601: $H(\hat{B}|\hat{X})_{V(\tilde{\cE}_n)} -
602: H(\hat{B}|\hat{X})_{\tilde{\cE}_n} \geq E-\delta$.
603: 
604: First consider the change in mutual information.  Since $Y_A=Y_B$ and
605: $Z_A=Z_B$ (as random variables), $I(\hat{X} ; \hat{B})_{\tilde{\cE}_n}
606: = I(X_AY_AZ_A ; BY_BZ_B)_{\tilde{\cE}_n} = I(X_A ; B |
607: YZ)_{\tilde{\cE}_n} + H(YZ)_{\tilde{\cE}_n}$ and similarly when we
608: replace $\tilde{\cE}_n$ with $V(\tilde{\cE}_n)$.  Since $V$ doesn't
609: act on $Y$ or $Z$, we have 
610: $H(YZ)_{\tilde{\cE}_n}=H(YZ)_{V(\tilde{\cE}_n)}$ and thus
611: \be\begin{split} I(\hat{X} ; \hat{B})_{V(\tilde{\cE}_n)}
612: &- I(\hat{X} ; \hat{B})_{\tilde{\cE}_n}
613: = I(X_A ; B | YZ)_{V(\tilde{\cE}_n)}
614: - I(X_A ; B | YZ)_{\tilde{\cE}_n}
615: \\&= \frac{1}{n}\sum_{j=1}^n I(X_A ; B | Y)_{\cE_{j}} - 
616: I(X_A ; B | Y)_{V^\dag(\cE_j)}
617: \\&= \frac{1}{n}\l(I(X_A ; B | Y)_{\cE_n} - I(X_A ; B | Y)_{\cE_0}\r)
618: + \frac{1}{n}\sum_{j=1}^n\l(I(X_A ; B | Y)_{\cE_{j-1}} - 
619: I(X_A ; B | Y)_{V^\dag(\cE_j)}\r)
620: \\&\geq C - \delta
621: + \frac{1}{n}\sum_{j=1}^n\l(I(X_A ; B | Y)_{\cE_{j-1}} - 
622: I(X_A ; B | Y)_{V^\dag(\cE_j)}\r)
623: \end{split}\ee
624: Recall that going from $\cE_{j-1}$ to $V^\dag(\cE_j)$ involves local
625: unitaries, a measurement by Bob and classical communication of the
626: outcome from Bob to Alice.  We claim that $I(X_A;B|Y)$ does not
627: increase under this process, meaning that the 
628: expression inside the sum on the last line is always nonnegative and
629: that $I(\hat{X} ; \hat{B})_{V(\tilde{\cE}_n)}
630: - I(\hat{X} ; \hat{B})_{\tilde{\cE}_n}\geq C-\delta$, implying our
631: desired conclusion.  To prove this,
632: write $I(X_A;B|Y)$ as $I(X_A;BY) - I(X_A;Y)$.  The $I(X_A;BY)$ term is
633: nonincreasing due to the data-processing inequality\cite{SN96}, while
634: $I(X_A;Y)$ can only increase since each round of communication only
635: causes $Y$ to grow.
636: 
637: Now we examine the change in entanglement.
638: \be\begin{split} H(\hat{B}|\hat{X})_{V(\tilde{\cE}_n)}
639: -& H(\hat{B}|\hat{X})_{\tilde{\cE}_n} = 
640: H(BY_BZ_B | X_AY_AZ_A)_{V(\tilde{\cE}_n)}
641: - H(BY_BZ_B | X_AY_AZ_A)_{\tilde{\cE}_n}
642: \\&=H(B | X_AYZ)_{V(\tilde{\cE}_n)}
643: - H(B | X_AYZ)_{\tilde{\cE}_n}
644: \\&= \frac{1}{n}\l(H(B|X_AY)_{\cE_n} - H(B|X_AY)_{\cE_0}\r)
645:  + \frac{1}{n}\sum_{j=1}^n
646: \l(H(B|X_AY)_{\cE_{j-1}} - H(B|X_AY)_{V^\dag(\cE_j)}\r)
647: \\ &\geq E - \delta
648:  + \frac{1}{n}\sum_{j=1}^n
649: \l(H(B|X_AY)_{\cE_{j-1}} - H(B|X_AY)_{V^\dag(\cE_j)}\r)
650: \end{split}\ee
651:  We would like to show that this last term is positive, or
652: equivalently that $H(B|X_AY)$ is at least as large for $\cE_{j-1}$ as
653: it is for $V^\dag(\cE_j)$.  This change from $\cE_{j-1}$ to
654: $V^\dag(\cE_j)$ involves local unitaries, a measurement by Bob and
655: another classical message from Bob to Alice, 
656: which we call $Y_j$.  Also, call the first $j-1$ messages $Y^{j-1}$.
657: Thus, we would like to show that $H(B|X_AY^{j-1})_{\cE_{j-1}} -
658: H(B|X_AY^{j-1}Y_j)_{V^\dag(\cE_j)} \geq 0$.  This can be expressed as an
659: average over $H(B)_{\cE_{j-1|x,y^{j-1}}} -
660: H(B|Y_j)_{V^\dag(\cE_{j|x,y^{j-1}})}$, where $\cE_{j-1|x,y^{j-1}}$
661: indicates that we have conditioned $\cE_{j-1}$ on $X_A=x$ and
662: $Y^{j-1}=y^{j-1}$.  This last quantity is positive 
663: because of principle that the average entropy of states output from a
664: projective measurement is no greater than the entropy of the
665: original state\cite{Nielsen99a}.  Thus $H(\hat{B}|\hat{X})_{V(\tilde{\cE}_n)}
666: - H(\hat{B}|\hat{X})_{\tilde{\cE}_n} \geq E-\delta$.
667: 
668: As $n\ra \infty$, $\delta\ra 0$, proving the theorem.
669: 
670: The case when $E\leq 0$ is similar.  We now begin with
671: $H(B|X_AY_A)_{\cE_0} \leq n(-E+\delta) = -n(E-\delta)$ and since $X_A$
672: and $Y$ are classical registers, finish with
673: $H(B|X_AY_A)_{\cE_n} \geq 0$.  Thus $H(B|X_AY_A)_{\cE_n} -
674: H(B|X_AY_A)_{\cE_0} \geq n(E-\delta)$.  The rest of the proof is the
675: same as the $E>0$ case.
676: \end{proof}
677: 
678: \subsection{Two-way cbit, cobit, qubit and ebit capacities of unitary
679: gates}
680: \label{sec:u-bidi-caps}
681: So far we have two powerful results about unitary gate capacity
682: regions: \thm{bidi-ccc} relates $\CCE$ and $\CoCoE$ in the
683: $C_1,C_2\geq 0$ quadrant and \thm{Ce-cap} gives an expression for
684: $\CE(U)$ in terms of a single use of $U$.  Moreover, the proof of
685: \thm{Ce-cap} also showed that backwards classical communication cannot
686: improve the forward capacity of a unitary gate.  This allows us to
687: extend \thm{bidi-ccc} to $C_1 \leq 0$ or $C_2 \leq 0$ as follows:
688: 
689: \begin{theorem}\label{thm:bidi-nnn}
690: For arbitrary real numbers $C_1,C_2,E$, 
691: \be
692:   (C_1,C_2,E) \in \CCE  \Longleftrightarrow
693: 		(C_1,C_2,E{-}\min(C_1,0){-}\min(C_2,0)) \in 
694: 	{\CoCoE} \,. 
695: \label{eq:thm12}\ee
696: \end{theorem}
697: 
698: This theorem is a direct consequence of the following Lemma, which
699: enumerates the relevant quadrants of the $(C_1,C_2)$ plane.
700: 
701: \begin{lemma}
702: For any bipartite unitary or isometry $U$ and $C_1,C_2 \geq 0$, 
703: \bea 
704: 	C_2 \cbc + \<U\> & \geqslant & C_1 \ctc + E \qq
705: 	\quad \quad {\rm iff}
706: \label{eq:1locc} 
707: \\
708: 	\<U\> & \geqslant & C_1 \ctc + E \qq
709: 	\quad \quad {\rm iff}
710: \label{eq:celo}
711: \\
712: 	\<U\> & \geqslant & C_1 \cof + E \qq 
713: 	\quad \quad {\rm iff}
714: \label{eq:ccelo}
715: \\
716: 	C_2 \cob + \<U\> & \geqslant & C_1 \cof + (E{+}C_2) \qq
717: \label{eq:1loccc} 
718: \eea
719: %
720: and
721: %
722: \bea 
723: 	C_1 \ctc + C_2 \cbc + \<U\> & \geqslant & E \qq 
724: 	\quad {\rm iff}
725: \label{eq:2locc} 
726: \\
727: 	\<U\> & \geqslant & E \qq 
728: 	\quad {\rm iff}
729: \label{eq:elo}
730: \\
731: 	C_1 \cof + C_2 \cob + \<U\> & \geqslant & (E{+}C_1{+}C_2) \qq 
732: \label{eq:2loccc} 
733: \eea
734: \end{lemma}
735: 
736: Basically, the rate at which Alice can send Bob cbits while
737: consuming/generating ebits is not increased by (coherent) classical
738: communication from Bob to Alice, except for a trivial gain of
739: entanglement when the assisting classical communication is coherent.
740: 
741: \begin{proof}
742: Combining (TP) and coherent SD (\eq{SDC-cc}) yields $2\ctc + \qq +
743: \qtq + \qq \reduction \qtq + 2\cof$.  Canceling the $\qtq$ from both sides
744: and dividing by two gives us
745: % 
746: \be
747: 	\ctc +  \qq \geqslant  \cof \,.
748: \label{eq:tp-sd}
749: \ee
750: 
751: For the first part of the lemma, recall from the proof of
752: \thm{Ce-cap} that free backcommunication does not improve the forward
753: capacity of a gate.  This means that \eq{1locc} $\Rightarrow$
754: \eq{celo}.  We obtain \eq{celo} $\Leftrightarrow$ \eq{ccelo} from
755: \thm{bidi-ccc} and \eq{ccelo} $\Rightarrow$ \eq{1loccc} follows from
756: $\cof\geq\qq$ and composability (\thm{composability}).  Finally,
757: \eq{1loccc} $\Rightarrow$ \eq{1locc} because of \eq{tp-sd}.
758: 
759: For the second part of the theorem, \eq{2locc} $\Rightarrow$ \eq{elo}
760: follows from \thm{U-E-cap}, \eq{elo} $\Rightarrow$ \eq{2loccc} is
761: trivial and \eq{2loccc} $\Rightarrow$ \eq{2locc} is a consequence of
762: \eq{tp-sd}.
763: \end{proof}
764: 
765: {\em Quantum capacities of unitary gates:} These techniques also allow
766: us to determine the quantum capacities of unitary gates.  Define $\QQE$
767: to be the region $\{(Q_1,Q_2,E): U \geqslant Q_1 \qtq
768:  + Q_2 \qbq +E \qq\}$, corresponding to two-way quantum
769: communication.  We can also consider coherent classical communication
770: in one direction and quantum communication in the other; let $\QCoE$ be
771: the region $\{(Q_1,C_2,E) : U \geqslant Q_1 \qtq + C_2\cob
772: +E \qq\}$ and define $\CoQE$ similarly.
773: 
774: As a warmup, we can use the equality $2\cof = \qtq + \qq$ to relate
775: % 
776: C$\!_{\rm o}\!$E and QE, defined as 
777: % 
778: ${\rm C}\!_{\rm o}\!{\rm E}=\{(C,E) : (C,0,E)\in {\CoCoE} \}$ and 
779: % 
780: $\QE=\{(Q,E) : (Q,0,E)\in \QQE\}$.  We claim that
781: % 
782: \be
783:  (Q,E)\in\QE \Leftrightarrow (2Q, E-Q)\in {\rm C}\!_{\rm o}\!{\rm E} \,.
784: \label{eq:one-way-toff}
785: \ee
786: % 
787: To prove \eq{one-way-toff}, choose any $(Q,E)\in\QE$.  
788: % 
789: Then $\<U\> \geq Q \qtq + E\qq = 2Q \cof + (E-Q)\qq$, so 
790: % 
791: $(2Q, E-Q)\in {\rm C}\!_{\rm o}\!{\rm E}$.  Conversely, if
792: $(2Q,E-Q)\in {\rm C}\!_{\rm o}\!{\rm E}$, then 
793: $U\geq 2Q \cof + (E-Q)\qq = Q \qtq + E \qq$, so $(Q,E)\in \QE$.
794: 
795: Note that the above argument still works if we add the same resource,
796: such as $Q_2\qbq$, to the right hand side of each resource inequality.
797: Therefore, the same argument that proved \eq{one-way-toff} also
798: establishes the following equivalences for bidirectional rate regions:
799: % 
800: \be
801: \begin{array}{ccc}
802: (Q_1,Q_2,E)\in\QQE & \Longleftrightarrow & (2Q_1, Q_2, E-Q_1) \in{\CoQE} 
803: \\[2ex] \Updownarrow & & \Updownarrow
804: \\[2ex] (Q_1,2Q_2, E-Q_2) \in {\QCoE} & \Longleftrightarrow & 
805: 	(2Q_1,2Q_2,E-Q_1-Q_2)\in {\CoCoE}
806: \end{array} .
807: \ee
808: % 
809: Finally, \eq{thm12} further relates QQE, QCE, CQE, CCE, where QCE and
810: CQE are defined similarly to $\QCoE$ and $\CoQE$ but with incoherent
811: classical communication instead.
812: 
813: Thus once one of the capacity regions (say $\CoCoE$) is determined, all 
814: other capacity regions discussed above are determined.  The main open
815: problem that remains is to find an efficiently computable expression
816: for part of this capacity region.  \thm{Ce-cap} gives a formula for
817: the one-way cbit/ebit tradeoff that involves only a single use of the
818: unitary gate, but we still need upper bounds on the optimal ensemble
819: size and ancilla dimension for it to be practical.
820: 
821: \section{Collected proofs}
822: \label{sec:ccc-proofs}
823: In this section we give proofs that various protocols can be made
824: coherent.  We start with Rules I and O (from \sect{ccc-use}), which
825: gave conditions for when coherently decoupled cbits could be replaced
826: by cobits in asymptotic protocols.  Then we show specifically how
827: remote state preparation can be made coherent, proving
828: \eq{coherent-rsp}.  Finally, we show how two-way classical
829: communication from unitary operations can be made coherent, and prove
830:  \thm{bidi-ccc}. 
831: 
832: \subsection{Proof of Rule I}
833: 
834: In what follows we shall fix $\epsilon$ and consider a sufficiently
835: large $n$ so that the protocol is $\epsilon$-valid,
836: $\epsilon^2$-decoupled and accurate to within $\epsilon$.
837: %We also drop the index $n$ for readability.
838:   
839: Whenever the resource inequality features
840: $[c \rightarrow c]$ in the input this means that Alice performs
841: a von Neumann measurement on some subsystem $A_1$ of dimension $D
842: \approx \exp(n(R+\delta))$, 
843: the outcome of which she sends to Bob,
844: who then performs an unitary operation depending on the received information.
845: 
846: Before Alice's von Neumann measurement, the joint state of  $A_1$ and 
847: the remaining quantum system $Q$ is
848: $$
849: \sum_x \sqrt{p_x}\ket{x}^{A_1} \ket{\phi_x}^{Q},
850: $$
851: where by $\epsilon$-validity
852: $$
853: \sum_x |p_x - D^{-1}| \leq \epsilon. 
854: $$
855: Upon learning the measurement outcome $x$, Bob
856: performs some unitary $U_x$ on his part of $Q$, almost decoupling
857: it from $x$:
858: $$
859: \left\|\sum_x p_x \proj{x} \otimes \theta_x' -
860: \sum_x p_x \proj{x} \otimes \bar{\theta}' \right\|_1 
861: = \sum_x p_x  \| \theta_x' - \bar{\theta}'  \|_1 
862: \leq \epsilon^2,
863: $$
864: where $\ket{\theta_x'} = U_x \ket{\phi_x}$ and
865: $\bar{\theta}' = \sum_x p_x \theta_x$.  To simplify the analysis,
866: extend $Q$ to a larger Hilbert space on which there exist
867: purifications $\oprod{\bar{\theta}} \ext \bar{\theta}'$ and 
868: $\oprod{\theta_x} \ext \oprod{\theta_x'}$ such that (according to
869: \lem{purify-dist}) $\|\theta_x-\bar{\theta}\|_1 \leq 2
870: \sqrt{\|\theta_x'-\bar{\theta}'\|_1}$.  Then
871: \be
872: \sum_x p_x  \| \theta_x - \bar{\theta}  \|_1 \leq
873: \sum_x p_x  2\sqrt{\| \theta_x' - \bar{\theta}' \|_1} \leq
874: 2 \sqrt{\sum_x p_x  \| \theta_x' - \bar{\theta}' \|_1} \leq
875: 2\epsilon,\ee
876: where the second inequality uses the concavity of the square root.
877: 
878: If Alice refrains from the measurement and instead sends $A_1$ through
879: a \emph{coherent} channel, using $n(R+\delta)$ cobits, the resulting state is
880: $$
881: \sum_x \sqrt{p_x} \ket{x}^{A_1} \ket{x}^{B_1} \ket{\phi_x}^{Q}.
882: $$
883: Bob now performs the \emph{controlled} unitary 
884: $\sum_x \proj{x}^{B_1} \otimes  U_x$, giving rise to
885: $$
886: \ket{\Upsilon}^{A_1B_1Q} = 
887: \sum_x \sqrt{p_x} \ket{x}^{A_1} \ket{x}^{B_1} \ket{\theta_x}^Q.
888: $$
889: We may assume, w.l.o.g., that $\braket{\bar{\theta}}{\theta_x}$ is real
890: and positive for 
891: all $x$, as this can be accomplished by either Alice or Bob via
892: an $x$-dependent global phase rotation. 
893: 
894: We now claim that $\ket{\Upsilon}^{A_1B_1Q}$ is close to
895: $\ket{\Phi_D}^{A_1B_1} \ket{\bar{\theta}}^Q$.
896: Indeed
897: \be 
898: \bra{\Upsilon} \Gamma\rangle \ket{\bar{\theta}} =
899: \sum_x \sqrt{\frac{p_x}{D}} \braket{\theta_x}{\bar{\theta}}
900: \geq \sum_x \sqrt{\frac{p_x}{D}} 
901: \l(1 - \frac{1}{2}\| \theta_x - \theta \|_1\r), \ee
902: according to \eq{fid-trace}.
903: To bound this, we split the sum into two.
904: For the first term, we apply \eq{fid-trace} to the diagonal density
905: matrices $\sum_x p_x\proj{x}$ and $\sum_x D^{-1}\proj{x}$ to obtain
906: \be \sum_x \sqrt{\frac{p_x}{D}} \geq 1 - \frac{1}{2} \sum_x |p_x -
907: D^{-1}| \geq 1-\frac{\epsilon}{2} \ee
908: The second term is
909: \ben \sum_x \sqrt{\frac{p_x}{D}} \frac{1}{2}\| \theta_x - \theta \|_1
910:  &=& \sum_x \frac{1}{2}\l[p_x + \frac{1}{D} -
911: \bigl(\sqrt{p_x}-\sqrt{1/D}\bigr)^2\r]
912:  \frac{1}{2}\| \theta_x - \theta \|_1 \\
913: &\leq & \sum_x \frac{1}{2}\l(p_x + \frac{1}{D}\r)
914:  \frac{1}{2}\| \theta_x - \theta \|_1
915: \leq \sum_x \frac{1}{2}\l(2p_x + \l|p_x - \frac{1}{D}\r|\r)
916:  \frac{1}{2}\| \theta_x - \theta \|_1
917: \\ &\leq &\sum_x p_x  \frac{1}{2}\| \theta_x - \theta \|_1
918: + \sum_x \l|p_x - \frac{1}{D}\r| \leq 2\epsilon. \een
919: Putting this together, we find that 
920: $$ \bra{\Upsilon} \Gamma\rangle \ket{\bar{\theta}} \geq 1-3\epsilon $$
921: and  by \eq{fid-trace},
922: $$
923: \|\Upsilon - \Phi_D \otimes \bar{\theta} \|_1 \leq \sqrt{6\epsilon}
924: $$ 
925: Finally, since tracing out subsystems cannot increase trace distance,
926: $$
927: \|\Upsilon^{A_1 B_1} - \Phi_D \|_1 \leq \sqrt{6\epsilon}
928: $$ 
929: Thus, the total effect of replacing cbits
930: cobits is the generation of a state close to
931: $\Phi_D$.  This analysis ignores the fact that the cobits are only
932: given up to an error $\epsilon$.  However, due to the triangle
933: inequality, this only enters in as an
934: additive factor, and the overall error of $\epsilon +
935: \sqrt{6\epsilon}$ is still asymptotically vanishing.  Furthermore,
936: this mapping preserves the $\epsilon$-validity of the original
937: protocol (with respect to the inputs of $\alpha$) since all we have
938: done to Alice's states is to add purifying systems and add phases,
939: which w.l.o.g.~we can assume are applied to these purifying systems.
940: 
941: We have thus shown
942: $$
943: \alpha + R \,\cof \geq \beta + R \,[q \, q].
944: $$
945: \eq{cc-equality} and Lemmas \ref{lemma:noo} and \ref{lemma:cancel}
946: give the desired result
947: $$
948: \alpha + \frac{R}{2} \,[q \rightarrow q] \geq \beta + \frac{R}{2} \,[q \, q].
949: $$
950: \qed\bigskip
951: 
952: %{\tt We should show that the uniformity condition on $p$ may
953: %be relaxed, requiring only  
954: %Using an easy kind of ``homophonic substitution'' \cite{guenther}, we may
955: %relax the uniformity condition  for $p$: it suffices that 
956: %$n^{-1} \log p_x \approx {\rm const.}$ for all $x$.}
957: 
958: \subsection{Proof of Rule O}
959: Again fix $\epsilon$ and consider a sufficiently
960: large $n$ so that the protocol is $\epsilon$-valid,
961: $\epsilon^2$-decoupled and accurate to within $\epsilon$.
962: Now the roles of Alice and Bob are somewhat interchanged.
963: Alice performs a unitary operation depending on the classical
964: message $x$ to be sent and Bob performs a von Neumann measurement on
965: some subsystem $B_1$ which almost always succeeds in reproducing the message.
966: Namely, if  we denote by $p_{x'|x}$ the probability of outcome $x'$ given 
967: Alice's message was $x$ then, for sufficiently large $n$,
968: $$
969: \frac{1}{D} \sum_x p_{x|x} \geq 1 - \epsilon.
970: $$
971: Again $D=\exp(n(R+\delta))$.
972: Before Bob's measurement, the state of $B_1$ and 
973: the remaining quantum system $Q$ is
974: $$
975: \sum_{x'} \sqrt{p_{x'|x}} \ket{x'}^{B_1} \ket{\phi_{xx'}}^{Q}.
976: $$
977: Based on the outcome $x'$ of his measurement, Bob
978: performs some unitary $U_{x'}$ on $Q$, 
979: leaving the state of $Q$ almost decoupled from $xx'$:
980: $$
981: \left\|\sum_{xx'} D^{-1} p_{x'|x} \proj{x} \otimes \proj{x'} \otimes
982: \theta_{xx'}' - \sum_{xx'} D^{-1} p_{x'|x} \proj{x} \otimes \proj{x'} 
983: \otimes \bar{\theta}'\right\|_1 \leq \epsilon^2,
984: $$
985: where $\ket{\theta_{xx'}'} = U_{x'} \ket{\phi_{xx'}}$ and
986: $\bar{\theta}' = D^{-1} \sum_{xx'} p_{x'|x} \theta_{xx'}'$. 
987: Observe, as before, that we can use \lem{purify-dist} to extend $Q$ so
988: that $\bar{\theta}\ext\bar{\theta}'$ and
989: $\theta_{xx'}\ext\theta_{xx'}'$ are pure states,
990: $\braket{\bar{\theta}}{\theta_{xx}}$ is real and positive and
991: $\|\theta_{xx'}-\bar{\theta}\|_1 \leq 
992: 2\sqrt{\|\theta_{xx'}'-\bar{\theta}'\|_1}$.  Again we use the
993: concavity of $x\ra\sqrt{x}$ to bound
994: $$
995: D^{-1}\sum_{x} p_{x|x} \| \theta_{xx} - \bar{\theta}  \|_1 \leq
996: D^{-1}\sum_{xx'} p_{x|x'} \| \theta_{xx'} - \bar{\theta}  \|_1 \leq 2\epsilon.
997: $$
998: 
999: We now modify the protocol so that instead Alice performs
1000: \emph{coherent} communication. Given a subsystem $A_1$ in the
1001: state $\ket{x}^{A_1}$ she  encodes via 
1002: \emph{controlled} unitary operations, yielding 
1003: $$
1004: \ket{x}^{A_1} \sum_{x'} \sqrt{p_{x'|x}} \ket{x'}^{B_1} \ket{\phi_{xx'}}^{Q}.
1005: $$
1006: 
1007: Bob refrains from measuring $B_1$ and instead 
1008: performs the \emph{controlled} unitary 
1009: $\sum_{x'} \proj{x'}^{B_1} \otimes  U_{x'}$, giving rise to
1010: $$
1011: \ket{x}^{A_1}\ket{\Upsilon_x}^{B_1Q} = 
1012: \ket{x}^{A_1} \left( \sum_{x'} \sqrt{p_{x'|x}} \ket{x'}^{B_1} 
1013: \otimes \ket{\theta_{xx'}}^{Q} \right).
1014: $$
1015: We claim that this is a good approximation for $R\coftau +
1016: \<\bar{\theta}\>$, and according to the correctness of 
1017: the original protocol, $\bar{\theta}$ is close to the output of
1018: $\beta_n$.
1019: To check this, suppose Alice inputs $\ket{\Phi_D}^{RA_1}$ into the
1020: communication protocol.  We will compare the actual state
1021: $$\ket{\Upsilon}^{RA_1B_1Q} :=
1022: D^{-\smfrac{1}{2}}\sum_x
1023: \ket{x}^R\ket{x}^{A_1}\ket{\Upsilon_x}^{B_1Q}$$
1024:   with the ideal state
1025: $$\ket{\Phi_{\text{GHZ}}}^{RA_1B_1}\ot\ket{\bar{\theta}}^Q = 
1026: D^{-\smfrac{1}{2}}\sum_x
1027: \ket{x}^R\ket{x}^{A_1}\ket{x}^{B_1}\ket{\bar{\theta}}^Q.$$
1028:   Their inner
1029: product is 
1030: \ben \bra{\Upsilon}\Phi_{\text{GHZ}}\rangle\ket{\bar{\theta}}&=&
1031: \frac{1}{D}\sum_x
1032: \sqrt{p_{x|x}}\braket{\theta_{xx}}{\bar{\theta}}
1033: \geq
1034: \frac{1}{D}\sum_x
1035: p_{x|x}\braket{\theta_{xx}}{\bar{\theta}}
1036: \geq
1037: \frac{1}{D}\sum_x p_{x|x} 
1038: \l(1 - \half\l\|\theta_{xx}-\bar{\theta}\r\|_1\r)
1039: \\&\geq&
1040: \frac{1}{D}\sum_x p_{x|x} 
1041:  - \frac{1}{D}\sum_x \half\l\|\theta_{xx}-\bar{\theta}\r\|_1
1042: \geq (1 - \epsilon)-\epsilon = 1-2\epsilon
1043: \een
1044: Thus, we can apply \eq{fid-trace} to show that
1045: $$
1046: \|\Upsilon - \Phi_{GHZ} \otimes \theta \|_1 \leq 2 \sqrt{\epsilon}.
1047: $$ 
1048: 
1049: We have thus shown that
1050: $$
1051: \alpha \geq \beta + R \,\qtqtau.
1052: $$
1053: Using \thm{absolutize} and \eq{cc-equality}
1054: gives the desired result
1055: $$
1056: \alpha  \geq \beta + \frac{R}{2} \, \qq +  \frac{R}{2} \, \qtq. 
1057: $$
1058: \qed\bigskip
1059: 
1060: \subsection{Proof of Coherent RSP \peq{coherent-rsp}}
1061: To prove that RSP can be made coherent, we review the proof of
1062: \eq{RSP2} from \cite{BHLSW03} and show how it needs to be modified.
1063: We will assume knowledge of typical and conditionally typical
1064: projectors; for background on them, as well as the operator Chernoff
1065: bound used in the proof, see \cite{Winter99}.
1066: 
1067: The (slightly modified) proof from \cite{BHLSW03} is as follows.  Let
1068: $\cE =\sum_i p_i 
1069: \oprod{i}^{X_A} \ot \psi_i^{AB}$ be an ensemble of bipartite states,
1070: for which we would like to simulate $\cN_\cE$ or $U_\cE$.  Alice is
1071: given a string $i^n=(i_1,\ldots, i_n)$ and wants to prepare the joint
1072: state $\ket{\psi_{i^n}}^{AB} :=
1073: \ket{\psi_{i_1}}^{AB}\cdots\ket{\psi_{i_n}}^{AB}$.  
1074: Let $Q_{i^n}$ be the empirical distribution of $i^n$, i.e. the
1075: probability distribution on $i$ obtained by sampling from $i^n$.  We
1076: assume that  
1077: $\|p-Q_{i^n}\|_1\leq \delta$, and since our
1078: simulation of $U_\cE$ will be used in some $\eta$-valid protocol, we
1079: can do so with error $\leq \eta + \exp(-O(n\delta^2))$.  (Here
1080: $\eta,\delta\ra 0$ as $n\ra\infty$.)  Thus, the protocol begins by
1081: Alice projecting onto the set of $i^n$ with $\|p-Q_{i^n}\|_1\leq
1082: \delta$, in contrast with the protocol in \cite{BHLSW03}, which
1083: begins by having Alice measure $Q_{i^n}$ and send the result to Bob
1084: classically. 
1085: 
1086: Define $\Pi_{\cE^B|i^n,\delta}^n$ to be the
1087: conditionally typical projector for Bob's half of
1088: $\ket{\psi_{i^n}}^{AB}$, and let $\Pi_{\cE^B\!,\delta}^n$ be the typical
1089: projector for $n$ copies of $\cE^B$.  These projectors are defined in
1090: \cite{Winter99}, which also proves that the subnormalized state
1091: \be \ket{\psi'_{i^n}} = (\one \ot \Pi_{\cE^B\!,\delta}^n
1092: \Pi_{\cE^B|i^n,\delta}^n) \ket{\psi_{i^n}},
1093: \label{eq:rsp-typ-fidelity}\ee
1094: satisfies $\dblbraket{\psi'_{i^n}} \geq 1-2\epsilon$, where
1095: $\delta,\epsilon\ra 0$ as $n\ra\infty$.  This implies that
1096: $\|\psi_{i^n} - \psi''_{i^n}\|_1 \leq 2\sqrt{\epsilon}$, where we
1097: define the normalized state
1098: $\ket{\psi''_{i^n}} := \ket{\psi'_{i^n}} / \sqrt{\dblbraket{\psi'_{i^n}}}$.
1099: We will now write $\ket{\psi'_{i^n}}$ in a way which suggests
1100: how to construct it.  Let $\ket{\Phi_D}^{AB}$ be a maximally entangled
1101: state with $D := \rank \Pi_{\cE^B\!,\delta}^n$ and $\Phi_D^B = 
1102: \Pi_{\cE^B\!,\delta}^n / D$.  (By contrast, \cite{BHLSW03} chooses
1103: $\Phi$ to be a purification of $\Pi_{\sigma,\delta}^n$ with $\sigma :=
1104: \sum_x Q_{i^n}(x) \psi_x^B$.)
1105: Then $\ket{\psi'_{i^n}}$ can be
1106: written as $(M_{i^n} \ot \one)\ket{\Phi_D}$ where $\tr M_{i^n}^\dag
1107: M_{i^n} = D^{-1} 
1108: \dblbraket{\psi'_{i^n}}$.  Thus, Alice will apply a POVM
1109: composed of rescaled and rotated versions of $M_{i^n}$ to her half of
1110: $\ket{\Phi_D}$, and after transmitting the measurement outcome $k$ to Bob,
1111: he can undo the rotation and obtain his half of the correct state.
1112: The cost of this procedure is $\log D$ ebits and $\log K$ ebits, where
1113: we will later specify the number of POVM outcomes $K$.
1114: 
1115: We now sketch the proof that this is efficient.  From \cite{Winter99},
1116: we find the bounds
1117: \begin{align}
1118: D = \rank \Pi_{\cE^B\!,\delta}^n & \leq \exp\l(n(H(B)_\cE+\delta)\r)\\
1119: \tr_A \oprod{\psi'_{i^n}} &\leq
1120: \exp\l(-n(H(B|X_A)_\cE + \delta)\r) \Pi_{\cE^B\!,\delta}^n
1121: \end{align}
1122: Combining these last two equations and \eq{rsp-typ-fidelity} with the
1123: operator Chernoff 
1124: bound\cite{Winter99} means that there exist a set of unitaries
1125: $U_1,\ldots,U_K$ such that 
1126: $\log K \leq n(I(X_A ; B)_\cE + 3\delta + o(1))$ and
1127: whenever $\|Q_{i^n}-p\|_1\leq \delta$ we have
1128: \be
1129: (1-\epsilon)\frac{\Pi_{\cE^B\!,\delta}^n}{D} \leq
1130: \frac{1}{K}\sum_{k=1}^K \frac{U_k^\dag M_{i^n}^\dag M_{i^n} U_k}
1131: {\tr M_{i^n}^\dag M_{i^n}} \leq 
1132: (1+\epsilon)\frac{\Pi_{\cE^B\!,\delta}^n}{D}.
1133: \label{eq:rsp-randomization}\ee
1134: These conditions mean that Alice can construct a POVM
1135: $\{A_1^{(i^n)},\ldots,A_K^{(i^n)},A_{\text{fail}}^{(i^n)}\}$ with
1136: \be\begin{split}
1137: A_k^{(i^n)} &:= \frac{D}{\sqrt{K(1+\epsilon)\tr M_{i^n}^\dag M_{i^n}}}
1138: M_{i^n} U_k^*\\
1139: A_{\text{fail}}^{(i^n)} & := \sqrt{\Pi_{\cE^B\!,\delta}^n-\sum_k
1140: A_k^\dag A_k} 
1141: \end{split}\ee
1142: According to \eq{rsp-randomization}, the ``fail'' outcome has
1143: probability $\leq 2\epsilon$ of occurring when Alice applies this POVM
1144: to half of $\ket{\Phi_D}$.  And since $(U_k^* \ot
1145: \one)\ket{\Phi_D} = (\one \ot U_k^\dag)\ket{\Phi_D}$, if Alice sends
1146: Bob the outcome $k$ and Bob applies $U_k$ then the residual state
1147: will be $\ket{\psi''_{i^n}}$.
1148: 
1149: We now explain how to make the above procedure coherent.  First
1150: observe that conditioned on not observing the ``fail'' outcome, the
1151: residual state is completely independent of the classical message $k$.
1152: Thus, we can apply Rule I.  However, a variant of Rule O is also
1153: applicable, in that there is no need to assume the input $\ket{i^n}$
1154: is a classical register.  Again conditioning on success,  the
1155: only record of $i^n$ in the final state is the output state
1156: $\ket{\psi''_{i^n}}$.  Thus, if Alice performs the POVM
1157: \be A_k := \sum_{i^n} \oprod{i^n} \ot A_k^{(i^n)},\ee
1158: (with $A_{\text{fail}}$ defined similarly) and Bob decodes using
1159: \be \sum_k \oprod{k} \ot U_k\ee
1160: then (conditioned on a successful measurement outcome) $\sum_{i^n}
1161: \sqrt{p_{i^n}}\ket{i^n}^R\ket{i^n}^{X_A}$ will be 
1162: coherently mapped to 
1163: $\sum_{i^n}
1164: \sqrt{p_{i^n}}\ket{i^n}^{R}\ket{i^n}^{X_A}\ket{\psi''_{i^n}}^{AB}$.
1165: This achieves a simulation of $\<U_\cE : \cE^{X_A}\>$ using $I(X_A;B)$
1166: cobits and $H(B)$ ebits.  According to Rule I, the coherent
1167: communication returns $I(X_A;B)$ ebits at the end of the protocol,
1168: bringing the net entanglement cost down to $H(B|X_A)$.  Thus we have
1169: proven \eq{coherent-rsp}.
1170: \qed\bigskip
1171: 
1172: 
1173: \subsection{Proof of \thm{bidi-ccc}}
1174: For ease of notation, we first consider the $E=0$ case, so our
1175:  starting hypothesis is that $\<U\>\geq C_1\ctc + C_2\cbc$.
1176: At the end of the proof we will return to the $E \neq 0$ case.
1177: 
1178: \subsubsection{The definition of $\cP_n$}  
1179: 
1180: Formally, \eq{cbit-toff} indicates the existence of sequences of
1181: nonnegative real numbers $\{\epsilon_n\},\{\delta_n\}$ satisfying
1182: $\epsilon_n, \delta_n {\; \ra \; } 0$ as $n{\; \ra\; }\infty$; a
1183: sequence of protocols $\cP_n = (V_n \! \otimes \! W_n) \, U \, \cdots
1184: \, U \, (V_1 \!  \otimes \! W_1) \, U \, (V_0 \!  \otimes \! W_0)$,
1185: where $V_j,W_j$ are local isometries that may also act on extra local
1186: ancilla systems, and sequences of integers $C_1^{(n)},C_2^{(n)}$
1187: satisfying $nC_1 \geq C_1^{(n)} \geq n(C_1{-}\delta_n)$, $nC_2 \geq
1188: C_2^{(n)} \geq n(C_2{-}\delta_n)$, such that the following success
1189: criterion holds.
1190: 
1191: Let $a \in \{0,1\}^{C_1^{(n)}}$ and $b \in \{0,1\}^{C_2^{(n)}}$ be the
1192: respective messages of Alice and Bob.  Let $\ket{\varphi_{ab}}:=\cP_n
1193: (\ket{a}_{\A_1}\ket{b}_{\B_1})$.  Note that $\ket{\varphi_{ab}}$
1194: generally occupies a space of larger dimension than $\A_1 \ot \B_1$
1195: since $\cP_n$ may add local ancillas.
1196: % 
1197: To say that $\cP_n$ can transmit classical messages, we require that
1198: local measurements on 
1199: $\ket{\varphi_{ab}}$ can generate messages $b'$ for Alice and $a'$ for
1200: Bob according to a distribution $\Pr(a'b'|ab)$ such that
1201: % 
1202: \be 
1203: \forall_{a,b} \quad
1204: \sum_{a',b'} \smfrac{1}{2}\l| \, 
1205: \Pr(a'b'|ab) - \delta_{a,a'}\delta_{b,b'}\r| \leq \epsilon_n
1206: \label{eq:cc-condition}
1207: \ee
1208: %
1209: where $a', b'$ are summed over $\{0,1\}^{C_1^{(n)}}$ and
1210: $\{0,1\}^{C_2^{(n)}}$ respectively.  
1211: % 
1212: \eq{cc-condition} follows from applying our definition of a
1213: protocol to classical
1214: communication, taking the final state to be the distribution of
1215: the output classical messages.
1216: % 
1217: Since any measurement can be implemented as a joint unitary on the
1218: system and an added ancilla, up to a redefinition of $V_n, W_n$, we
1219: can assume
1220: % 
1221: \be
1222:  \ket{\varphi_{ab}} := \cP_n (\ket{a}_{\A_1}\ket{b}_{\B_1}) 
1223: = \sum_{a'\!,b'} |b'\>_{\A_1} |a'\>_{\B_1} 
1224: |\gamma_{a' \!, b'}^{a,b} \>_{\A_2 \B_2} 
1225: \label{eq:coh-comm} \, \ee
1226: % 
1227: where the dimensions of $\A_1$ and $\B_1$ are interchanged by $\cP_n$,
1228: and $|\gamma_{a'\!,b'}^{a,b}\>$ are subnormalized states with 
1229: $\Pr(a'b'|ab):=\<\gab|\gab\>$ satisfying \eq{cc-condition}.  Thus, for
1230: each $a,b$ most of the weight of $\ket{\varphi_{ab}}$ is contained in
1231: the $|\gamma_{a,b}^{a,b}\>$ term, corresponding to error-free
1232: transmission of the messages.  See Fig.\ I(a).
1233: 
1234: \subsubsection{The three main ideas for turning classical communication
1235: into coherent classical communication}
1236: 
1237: We first give an informal overview of the construction and the
1238: intuition behind it.  For simplicity, consider the error-free term 
1239: with $|\gamma_{a,b}^{a,b}\>$ in ${\A_2 \B_2}$.
1240: % 
1241: To see why classical communication via unitary means should be
1242: equivalent to coherent classical communication, consider the special
1243: case when $|\gamma_{a,b}^{a,b}\>_{\A_2 \B_2}$ is independent of $a,b$.  
1244: In this case, copying $a,b$ to local ancilla systems $\A_0,\B_0$
1245: before $\cP_n$ and discarding $\A_2 \B_2$ after $\cP_n$ leaves a state
1246: within trace distance $\eps_n$ of $\ket{b}_{\A_1} \ket{a}_{\A_0}
1247: \ket{a}_{\B_1}\ket{b}_{\B_0}$---the desired coherent classical
1248: communication. See Fig.\ I(b).
1249: % 
1250: In general $|\gamma_{a,b}^{a,b}\>_{\A_2 \B_2}$ will carry information
1251: about $a,b$, so tracing $\A_2 \B_2$ will break
1252: the coherence of the classical communication.
1253: % 
1254: Moreover, if the Schmidt coefficients of $|\gamma_{a,b}^{a,b}\>_{\A_2
1255: \B_2}$ depend on $a,b$, then knowing $a,b$ is not sufficient to
1256: coherently eliminate $|\gamma_{a,b}^{a,b}\>_{\A_2 \B_2}$ without
1257: some additional communication.  The remainder of our proof is built
1258: around the need to coherently eliminate this ancilla.
1259: 
1260: Our first strategy is to {\em encrypt} the classical messages $a,b$ by
1261: a shared key, in a manner that preserves coherence (similar to that in
1262: \mscite{Leung00}).  The coherent version of a shared key is a maximally
1263: entangled state.  Thus Alice and Bob (1) again copy their messages to
1264: $\A_0, \B_0$, then (2) encrypt, (3) apply $\cP_n$, and (4) decrypt.
1265: Encrypting the message makes it possible to (5) almost decouple the
1266: message from the combined ``key-and-ancilla'' system, which is
1267: approximately in a state $|\Gamma_{00}\>$ independent of $a,b$ (exact
1268: definitions will follow later).
1269: (6) Tracing out $|{\Gamma}_{00}\>$ gives the desired coherent
1270: communication.  Let $\cP_n'$ denote steps (1)-(5) (see Fig.\ I(c)).
1271: 
1272: % end of part 1
1273: 
1274: % figure 1 
1275: % \input qp-fig.tex
1276: 
1277: \begin{figure}[h]
1278: \centering \setlength{\unitlength}{0.56mm}
1279: \begin{picture}(300,55)
1280: % \put(0,-3){\framebox(220,60){}}
1281: 
1282: \put(137,50){\makebox{\bf (c)}}
1283: 
1284: \put(153,27.5){\line(1,5){4}}
1285: \put(153,27.5){\line(1,-5){4}}
1286: \put(157,47.5){\line(1,0){45}}
1287: \put(157,7.5){\line(1,0){45}}
1288: 
1289: \put(150,27.5){\line(1,5){4.5}}
1290: \put(150,27.5){\line(1,-5){4.5}}
1291: \put(154.5,50){\line(1,0){47.5}}
1292: \put(154.5,5){\line(1,0){47.5}}
1293: 
1294: \put(175,21.5){\line(1,0){10}}
1295: \put(175,33.5){\line(1,0){10}}
1296: 
1297: \put(178,40){\circle{3}}
1298: \put(178,33.5){\line(0,1){8}}
1299: \put(178,21.5){\line(0,-1){8}}
1300: \put(178,15){\circle{3}}
1301: 
1302: \put(182,33.5){\circle{3}}
1303: \put(182,47.5){\line(0,-1){15.5}}
1304: \put(182,5){\line(0,1){18}}
1305: \put(182,21.5){\circle{3}}
1306: 
1307: \put(198,33.5){\circle{3}}
1308: \put(198,50){\line(0,-1){18}}
1309: \put(198,7.5){\line(0,1){15.5}}
1310: \put(198,21.5){\circle{3}}
1311: 
1312: \multiput(202,50)(2,0){16}{{\line(1,0){1}}}
1313: \multiput(202,5)(2,0){16}{{\line(1,0){1}}}
1314: \multiput(202,47.5)(2,0){16}{{\line(1,0){1}}}
1315: \multiput(202,7.5)(2,0){16}{{\line(1,0){1}}}
1316: 
1317: \multiput(210,40)(2,0){6}{{\line(1,0){1}}}
1318: \multiput(210,15)(2,0){6}{{\line(1,0){1}}}
1319: \multiput(210,21.5)(2,0){6}{{\line(1,0){1}}}
1320: \multiput(210,33.5)(2,0){6}{{\line(1,0){1}}}
1321: 
1322: \multiput(228,25.5)(2,0){3}{{\line(1,0){1}}}
1323: \multiput(228,29.5)(2,0){3}{{\line(1,0){1}}}
1324: 
1325: \multiput(214,40)(0,2){5}{{\line(0,1){1}}}
1326: \multiput(218,33.5)(0,2){9}{{\line(0,1){1}}}
1327: 
1328: \multiput(214,15)(0,-2){6}{{\line(0,-1){1}}}
1329: \multiput(218,21.5)(0,-2){8}{{\line(0,-1){1}}}
1330: 
1331: \put(218,50){\circle{3}}
1332: \put(214,47.5){\circle{3}}
1333: \put(218,7.5){\circle{3}}
1334: \put(214,5){\circle{3}}
1335: 
1336: \put(195,21.5){\line(1,0){7}}
1337: \put(195,33.5){\line(1,0){7}}
1338: 
1339: \put(175,15){\line(1,0){27}}
1340: \put(175,40){\line(1,0){27}}
1341: 
1342: \put(195,25.5){\line(1,0){7}}
1343: \put(195,29.5){\line(1,0){7}}
1344: 
1345: \put(201,26.5){\makebox{
1346: $\} | \hs \gamma $}\raisebox{0.3ex}{\tiny 
1347: $
1348: ^{a \hs {\oplus} \hs x \hs, b \hs \oplus \hs y}
1349: _{\hs a \! {'} \! \oplus \hs x \hs, b \hs ' \! \hs \oplus \! y}
1350: $}{$\hs \>$}}
1351: 
1352: % \put(201,26.5){\makebox{
1353: % $\} | \hs \gamma $}{\tiny$_{a \hs {\oplus} \hs x \hs, b \hs \oplus \hs y}
1354: % 	  ^{a \! ' \! \oplus \hs x \hs, b \! ' \! \oplus \hs y}$}{$\hs \>$}} 
1355: 
1356: \put(185,20){\framebox(10,15){$\cP_n$}}
1357: 
1358: \put(158,52){\makebox{$\A_4$}}
1359: \put(158,43){\makebox{$\A_3$}}
1360: \put(161,38){\makebox{$\A_0$}}
1361: \put(161,32){\makebox{$\A_1$}}
1362: 
1363: \put(161,20){\makebox{$\B_1$}}
1364: \put(161,14){\makebox{$\B_0$}}
1365: \put(158,9){\makebox{$\B_3$}}
1366: \put(158,0){\makebox{$\B_4$}}
1367: 
1368: \put(168,38){\makebox{$|0\>$}}
1369: \put(168,32){\makebox{$|a\>$}}
1370: \put(168,20){\makebox{$|b\>$}}
1371: \put(168,14){\makebox{$|0\>$}}
1372: 
1373: \put(203,38.5){\makebox{\small $|a\>$}}
1374: \put(203,32.5){\makebox{\small $|b'\>$}}
1375: \put(203,20){\makebox{\small $|a'\>$}}
1376: \put(203,14){\makebox{\small $|b\>$}}
1377: 
1378: \put(188,2){\makebox{\footnotesize{$y$}}}
1379: \put(188,8){\makebox{\footnotesize{$x$}}}
1380: 
1381: \put(188,52){\makebox{\footnotesize{$y$}}}
1382: \put(188,45){\makebox{\footnotesize{$x$}}}
1383: 
1384: \put(230,2.5){\makebox(10,50){$\left. 
1385: \begin{array}{c} {}\\{~}\\{~}\\{~}\\{~}\\{~}\\{~}\\{~} \end{array} \right\} $}}
1386: 
1387: \put(240,26){\makebox{$|\Gamma_{\!a \hs \oplus \hs a' \hs, b \hs \oplus 
1388: \hs b' \hs }\>$}}
1389: 
1390: %-------------------------------
1391: \put(70,50){\makebox{\bf (b)}}
1392: 
1393: \put(85,21.5){\line(1,0){10}}
1394: \put(85,33.5){\line(1,0){10}}
1395: 
1396: \put(88,40){\circle{3}}
1397: \put(88,33.5){\line(0,1){8}}
1398: \put(88,21.5){\line(0,-1){8}}
1399: \put(88,15){\circle{3}}
1400: 
1401: \put(105,21.5){\line(1,0){8}}
1402: \put(105,33.5){\line(1,0){8}}
1403: 
1404: \put(85,15){\line(1,0){28}}
1405: \put(85,40){\line(1,0){28}}
1406: 
1407: \put(95,20){\framebox(10,15){$\cP_n$}}
1408: 
1409: \put(71,38){\makebox{$\A_0$}}
1410: \put(71,32){\makebox{$\A_1$}}
1411: 
1412: \put(71,20){\makebox{$\B_1$}}
1413: \put(71,14){\makebox{$\B_0$}}
1414: 
1415: \put(78,38){\makebox{$|0\>$}}
1416: \put(78,32){\makebox{$|a\>$}}
1417: \put(78,20){\makebox{$|b\>$}}
1418: \put(78,14){\makebox{$|0\>$}}
1419: 
1420: \put(115,39){\makebox{\small $|a\>$}}
1421: \put(115,33){\makebox{\small $|b'\>$}}
1422: \put(115,19){\makebox{\small $|a'\>$}}
1423: \put(115,13){\makebox{\small $|b\>$}}
1424: 
1425: \put(105,25.5){\line(1,0){5}}
1426: \put(105,29.5){\line(1,0){5}}
1427: \put(111,26.25){\makebox{\small $\} | \! \gab \>$}}
1428: 
1429: %-------------------------------
1430: \put(0,50){\makebox{\bf (a)}}
1431: 
1432: \put(15,21.5){\line(1,0){5}}
1433: \put(15,33.5){\line(1,0){5}}
1434: 
1435: \put(30,21.5){\line(1,0){7}}
1436: \put(30,33.5){\line(1,0){7}}
1437: 
1438: \put(30,25.5){\line(1,0){4}}
1439: \put(30,29.5){\line(1,0){4}}
1440: % \put(34,38){\small {$\sum_{a',b'}$}}
1441: \put(39,33){\small {$|b'\>$}}
1442: \put(39,19){\small {$|a'\>$}}
1443: \put(35,26){\small {$\} \, |\!\gab \>$}}
1444: 
1445: 
1446: \put(20,20){\framebox(10,15){$\cP_n$}}
1447: 
1448: \put(1,32){\makebox{$\A_1$}}
1449: \put(1,20){\makebox{$\B_1$}}
1450: 
1451: \put(8,32){\makebox{$|a\>$}}
1452: \put(8,20){\makebox{$|b\>$}}
1453: 
1454: \end{picture}
1455: \label{fig:3protocols}
1456: \caption{Schematic diagrams for $\cP_n$ and $\cP_n'$.
1457: % 
1458: (a) A given protocol $\cP_n$ for two-way classical communication.  
1459: % 
1460: The output is a superposition (over all $a',b'$) of the depicted states, 
1461: with most of the weight in the $(a',b')=(a,b)$ term.  
1462: The unlabeled output systems in the state $|\gab\>$ are $\A_2,\B_2$.
1463: %
1464: (b) The same protocol with the inputs copied to local ancillas $\A_0,
1465: \B_0$ before $\cP_n$.  If $|\g_{a,b}^{a,b}\>$ is independent of
1466: $a,b$, two-way coherent classical communication is achieved.
1467: %
1468: (c) The five steps of $\cP_n'$.  Steps (1)-(4) are shown in solid
1469: lines.  Again, the inputs are copied to local ancillas, but $\cP_n$ is
1470: used on messages encrypted by a coherent one-time-pad (the input
1471: $|a\>_{\A_1}$ is encrypted by the coherent version of the key
1472: $|x\>_{\A_3}$ and the output $|a' \hs \oplus x\>_{\B_1}$ is decrypted by
1473: $|x\>_{\B_3}$; similarly, $|b\>_{\B_1}$ is encrypted by $|y\>_{\B_4}$
1474: and $|b' \hs \oplus y\>_{\A_1}$ decrypted by $|y\>_{\A_4}$.  The
1475: intermediate state is shown in the diagram.  Step (5), shown in dotted
1476: lines, decouples the messages in $\A_{0,1},\B_{0,1}$ from
1477: $\A_{2,3,4},\B_{2,3,4}$, which is in the joint state very close to
1478: $|\Gamma_{00}\>$.
1479: % 
1480: }
1481: \end{figure}
1482: 
1483: % part 2 
1484: % \input part2.tex
1485: % beginning of part 2
1486: 
1487: If entanglement were free, then our proof of Theorem
1488: \ref{thm:bidi-ccc} would be finished.  However, we have borrowed
1489: $C_1^{(n)}{+}C_2^{(n)}$ ebits as the encryption key and replaced it
1490: with $|\Gamma_{00}\>$.  Though the entropy of entanglement has not
1491: decreased (by any significant amount), $|\Gamma_{00}\>$ is not
1492: directly usable in subsequent runs of $\cP_n'$.  To address this
1493: problem, we use a second strategy of running $k$ copies of $\cP_n'$ in
1494: parallel and performing entanglement concentration of
1495: $|\Gamma_{00}\>^{\otimes k}$.
1496: For sufficiently large $k$, with high probability, we recover most of
1497: the starting ebits.  The regenerated ebits can be used for more
1498: iterations of $\cP_n'^{\otimes k}$ to offset the cost of making the
1499: initial $k \lpm \! C_1^{(n)}{+}C_2^{(n)} \! \rpm$ ebits, without
1500: the need of borrowing from anywhere.
1501: 
1502: However, a technical problem arises with simple repetition of
1503: $\cP_n'$, which is that errors accumulate.  In particular, a na\"\i ve
1504: application of the triangle inequality gives an error $k \eps_n$ but
1505: $k$, $n$ are not independent.  In fact, the entanglement concentration
1506: procedure of \mscite{BBPS96} requires $k\gg
1507: \Sch(|\Gamma_{00}\>) = \exp(O(n))$ and we cannot guarantee that $k\epsilon_n
1508: \rightarrow 0$ as $k,n \ra\infty$.  Our third strategy is to treat the
1509: $k$ uses of $\cP_n'$ as $k$ uses of a slightly noisy channel, and
1510: encode only $l$ messages (each having $C_1^{(n)}, C_2^{(n)}$ bits in
1511: the two directions) using classical error correcting codes.  The error
1512: rate then vanishes with a negligible reduction in the communication
1513: rate and now making no assumption about how quickly $\epsilon_n$
1514: approaches zero.  We will see how related errors in decoupling and
1515: entanglement concentration are suppressed.   
1516: 
1517: We now describe the construction and analyze the error in detail.  
1518: 
1519: \subsubsection{The definition of $\cP_n'$}
1520: \vspace*{-2ex}
1521: % 
1522: \begin{enumerate}
1523: \item[0.]  
1524: % 
1525: Alice and Bob begin with inputs $\ket{a}_{\A_1}\ket{b}_{\B_1}$ and the
1526: entangled states $\ket{\Phi}^{\! \ot C_1^{(n)}}_{\A_3 \B_3}$ and
1527: $\ket{\Phi}^{\! \ot C_2^{(n)}}_{\A_4 \B_4}$.  (Systems $3$ and $4$
1528: hold the two separate keys for the two messages $a$ and $b$.)  The
1529: initial state can then be written as
1530: % 
1531: \be
1532: \frac{1}{\sqrt{N}}
1533: \sum_x \ket{xx}_{\A_3 \B_3}
1534: \sum_y \ket{yy}_{\A_4 \B_4}
1535: ~\ket{a}_{\A_1}\ket{b}_{\B_1} 
1536: \ee
1537: % 
1538: where $x$ and $y$ are summed over $\{0,1\}^{C_1^{(n)}}$ and
1539: $\{0,1\}^{C_2^{(n)}}$, and $N = \exp \lpm \! C_1^{(n)} {+} C_2^{(n)}
1540: \! \rpm$.
1541: 
1542: \item
1543: They coherently copy the messages to $\A_0, \B_0$.
1544: 
1545: \item 
1546: They encrypt the messages using the one-time-pad 
1547: % 
1548: $\ket{a}_{\A_1} \ket{x}_{\A_3} \ra \ket{a\oplus x}_{\A_1}\ket{x}_{\A_3}$
1549: % 
1550: and
1551: % 
1552: $\ket{b}_{\B_1} \ket{y}_{\B_4} \ra \ket{b\oplus y}_{\B_1}\ket{y}_{\B_4}$
1553: % 
1554: coherently to obtain
1555: % 
1556: \be
1557: \ket{a}_{\A_0} \ket{b}_{\B_0} \; 
1558: \frac{1}{\sqrt{N}}\sum_{xy}
1559: \ket{x}_{\A_3}\ket{y}_{\A_4} 
1560: \ket{x}_{\B_3}\ket{y}_{\B_4} 
1561: ~\ket{a \hs \oplus \hs x}_{\A_1}
1562: \ket{b \hs \oplus \hs y}_{\B_1}
1563: \,.
1564: \ee
1565: % 
1566: \item
1567: Using $U$ $n$ times, they apply $\cP_n$ to registers $\A_1$ and $\B_1$,
1568: obtaining an output state 
1569: \be
1570: \ket{a}_{\A_0} \ket{b}_{\B_0}
1571: \frac{1}{\sqrt{N}}\sum_{xy}
1572: \ket{x}_{\A_3} \ket{y}_{\A_4} 
1573: \ket{x}_{\B_3} \ket{y}_{\B_4} 
1574: \sum_{a',b'} 
1575: | b' \! \oplus y\>_{\A_1} |a' \! \oplus x\>_{\B_1} \gmany_{\A_2 \B_2}
1576: \,.
1577: \label{eq:ready-to-measure}
1578: \ee
1579: 
1580: \item Alice decrypts her message in $\A_1$ using her key $\A_4$ and Bob
1581: decrypts $\B_1$ using $\B_3$ coherently as 
1582: %
1583: $|b' \oplus y\>_{\A_1} |y\>_{\A_4} \ra |b'\>_{\A_1} |y\>_{\A_4}$
1584: % 
1585: and
1586: % 
1587: $|a' \oplus x\>_{\B_1} |x\>_{\B_3} \ra |a'\>_{\B_1} |x\>_{\B_3}$
1588: % 
1589: producing a state 
1590: % 
1591: \be
1592: \ket{a}_{\A_0} \ket{b}_{\B_0}
1593: \frac{1}{\sqrt{N}}\sum_{xy}
1594: \ket{x}_{\A_3} \ket{y}_{\A_4} 
1595:  \ket{x}_{\B_3} \ket{y}_{\B_4} 
1596: \sum_{a',b'}
1597: \ket{b'}_{\A_1} \ket{a'}_{\B_1} \gmany_{\A_2 \B_2}
1598: \,. \ee
1599: % 
1600: \item 
1601: Further {\sc cnot}s $\A_1 \ra \A_4$, $\A_0 \ra \A_3$, $\B_1 \ra \B_3$
1602: and $\B_0 \ra \B_4$ will leave $\A_{2,3,4}$ and $\B_{2,3,4}$ almost
1603: decoupled from the classical messages.  To see this, the state has become 
1604: %
1605: \bea
1606: % \begin{split}
1607: % 
1608: & & \ket{a}_{\A_0}  \ket{b}_{\B_0}
1609: \sum_{a',b'} \ket{b'}_{\A_1}\ket{a'}_{\B_1}
1610: \frac{1}{\sqrt{N}}\sum_{xy}
1611: \ket{a\oplus x}_{\A_3} 
1612: \ket{a'\oplus x}_{\B_3} 
1613: \ket{b'\oplus y}_{\A_4}
1614: \ket{b\oplus y}_{\B_4} 
1615: \gmany_{\A_2 \B_2}
1616: \non \\ & = & 
1617:  \ket{a}_{\A_0}  \ket{b}_{\B_0}
1618: \sum_{a',b'} \ket{b'}_{\A_1}\ket{a'}_{\B_1}
1619: \; \ket{\Gamma_{a\oplus a', b\oplus b'}}_{\A_{2,3,4} \B_{2,3,4}}
1620: \label{eq:phiab}
1621: \,, 
1622: % \end{split}
1623: \eea
1624: % 
1625: where 
1626: % 
1627: \bea 
1628: \ket{\Gamma_{\! a\oplus a' \!, b\oplus b' \hs}}_{\A_{2,3,4} \B_{2,3,4}} := 
1629: \frac{1}{\sqrt{N}} \sum_{xy}
1630: \ket{a\oplus x}_{\A_3} 
1631: \ket{a'\oplus x}_{\B_3} 
1632: \ket{b'\oplus y}_{\A_4}
1633: \ket{b\oplus y}_{\B_4} 
1634: \gmany_{\A_2 \B_2} \,.
1635: \eea
1636: % 
1637: The fact $\ket{\Gamma_{\! a\oplus a' \!, b\oplus b' \hs}}$ depends
1638: only on $a \oplus a'$ and $b \oplus b'$, without any other dependence
1639: on $a$ and $b$, can be easily seen by replacing $x,y$ with $a\oplus x,
1640: b\oplus y$ in $\sum_{xy}$ in the RHS of the above.
1641: % 
1642: Note that $\<\Gamma_{\! a\oplus a' \!, b\oplus b' \hs} | \Gamma_{\!
1643: a\oplus a' \!, b\oplus b' \hs} \> = \smfrac{1}{N}\sum_{xy}
1644: \Pr(a'\oplus x, b'\oplus y \, | \, a \oplus x , b \oplus y)$, so in
1645: particular for the state corresponding to the error-free term, we have
1646: $\<\Gamma_{00}|\Gamma_{00}\> = \smfrac{1}{N}\sum_{xy} \Pr(xy|xy) :=
1647: 1 - \bar{\eps}_n \geq 1-\epsilon_n$.\footnote{Thus it turns out that
1648: \eq{cc-condition} was more than we 
1649: needed; the {\em average} error (over all $a,b$) would have been
1650: sufficient.  In general, this argument shows that using shared
1651: entanglement (or randomness in the case of classical communication)
1652: can convert an average error condition into a maximum error condition,
1653: and will be further developed in \cite{DW05}.}
1654: 
1655: Suppose that Alice and Bob could project onto
1656: the space where $a'=a$ and $b'=b$, and tell each other they
1657: have succeeded (by using a little extra communication); then the
1658: resulting ancilla state $\smfrac{1}{\sqrt{1{-}\bar{\eps}_n}}
1659: |\Gamma_{00}\>$ has at least $C_1^{(n)} {+} \, C_2^{(n)} {+} \log
1660: (1{-}\eps_n)$ ebits, since its largest Schmidt coefficient is 
1661: $\leq \lbm \exp(C_1^{(n)}{+}C_2^{(n)} ) (1{-}\bar{\eps}_n) \rbm^{-1/2}$ and
1662: $\bar{\eps}_n \leq \eps_n$ (cf.~\prop{U-CgE}).
1663: % 
1664: Furthermore, $|\Gamma_{00}\>$ is manifestly independent of $a,b$.
1665: % 
1666: We will see how to improve the probability of successful projection
1667: onto the error free subspace by using block codes for error
1668: correction, and how correct copies of $|\Gamma_{00}\>$ can be
1669: identified if Alice and Bob can exchange a small amount of
1670: information.
1671: \end{enumerate}
1672: 
1673: \subsubsection{Main idea on how to perform error correction}
1674: 
1675: As discussed before, $|\Gamma_{00}\>$ cannot be used directly as an
1676: encryption key -- our use of entanglement in $\cP_n'$ is not
1677: catalytic.
1678: % 
1679: Entanglement concentration of many copies of $|\Gamma_{00}\>$ obtained
1680: from many runs of $\cP_n'$ will make the entanglement overhead for the
1681: one-time-pad negligible, but errors will accumulate.
1682: % 
1683: The idea is to suppress the errors in many uses of $\cP_n'$ by error
1684: correction.
1685: % 
1686: This has to be done with care, since we need to simultaneously ensure
1687: low enough error rates in both the classical message and the state to
1688: be concentrated, as well as sufficient decoupling of the classical
1689: messages from other systems.
1690: 
1691: Our error-corrected scheme will have $k$ parallel uses of $\cP_n'$,
1692: but the $k$ inputs are chosen to be a valid codeword of an error
1693: correcting code.  Furthermore, for each use of $\cP_n'$, the state in
1694: $\A_{2,3,4} \B_{2,3,4}$ will only be collected for entanglement
1695: concentration if the error syndrome is trivial for that use of
1696: $\cP_n'$.  We use the fact that errors occur rarely (at a rate of
1697: $\epsilon_n$, which goes to zero as $n\ra\infty$) to show that (1) most
1698: states are still used for concentration, and (2) communicating the
1699: indices of the states with non trivial error syndrome requires a
1700: negligible amount of communication.
1701: 
1702: % wcl: note changing notation: P_n'' -> P_{nk}''  
1703: 
1704: \subsubsection{Definition of $\cP_{nk}''$: error corrected version of
1705: $(\cP_n')^{\ot k}$ with entanglement concentration}
1706: 
1707: We construct two codes, one used by Alice to signal to Bob and one
1708: from Bob to Alice.  We consider high distance codes.  The distance of
1709: a code is the minimum Hamming distance between any two codewords,
1710: i.e.\ the number of positions in which they are different.
1711: 
1712: First consider the code used by Alice.  Let $N_1=2^{C_1^{(n)}}$.
1713: Alice is coding for a channel that takes input symbols from
1714: $[N_1]:=\{1,\ldots,N_1\}$ and has probability $\leq \epsilon_n$ of
1715: error on any input (the error rate depends on both $a$ and $b$).
1716: % 
1717: We would like to encode $[N_1]^l$ in $[N_1]^k$ using a code with
1718: distance $2k\alpha_n$, where $\alpha_n$ is a parameter that will be
1719: chosen later.  Such a code can correct up to any $\lfloor k\alpha_n
1720: {-} \smfrac{1}{2} \rfloor$ errors (without causing much problem, we
1721: just say that the code corrects $k\alpha_n$ errors).  Using standard
1722: arguments\footnote{We show the existence of a maximal code by
1723: repeatedly adding new 
1724: codewords that have distance $\geq 2k\alpha_n$ from all other chosen
1725: codewords.  This gives at least $N^k / \vol(N,2k\alpha_n,k)$
1726: codewords, where $\vol(N,k\delta,k)$ is the number of words in $[N]^k$
1727: within a distance $k\delta $ of a fixed codeword.
1728: % 
1729: But $\vol(N,k\delta,k) \leq \binom{k}{k\delta}N^{k\delta} \leq 2^{k
1730: H_2(\delta)} N^{k\delta}$.  (See \cite{CT91} or \eq{binom-bounds}
1731: later in this thesis for a derivation of $\binom{k}{k\delta} \leq 2^{k
1732: H_2(\delta)}$.)
1733: % 
1734: Altogether, the number of codewords $:= N^l \geq N^k / (2^{k H_2(2 \alpha_n)}
1735: N^{2 k \alpha_n})$, thus $l \geq k \lbm 1-2\alpha_n -
1736: \smfrac{H_2(2\alpha_n)}{\log N}\rbm$.}, we can construct such a code with $l \geq k
1737: \lbm 1 {-} 2\alpha_n {-} H_2(2\alpha_n)/C_1^{(n)} \rbm $, where
1738: $H_2(p)=-p\log p-(1{-}p)\log (1{-}p)$ is the binary entropy.
1739: % 
1740: The code used by Bob is chosen similarly, with $N_2=2^{C_2^{(n)}}$
1741: input symbols to each use of $\cP_n'$.  For simplicity, Alice's and
1742: Bob's codes share the same values of $l$, $k$ and $\alpha_n$.   
1743: % 
1744: We choose $\alpha_n \geq \max(1/C_1^{(n)}, 1/C_2^{(n)})$ so that $l
1745: \geq k(1{-}3\alpha_n)$.  
1746: 
1747: Furthermore, we want the probability of having $\geq k\alpha_n$ errors to be
1748: vanishingly small.  This probability is $\leq\exp(-k
1749: D(\alpha_n\|\epsilon_n)) \leq \exp(k + k\alpha_n\log\epsilon_n)$
1750: (using arguments from \cite{CT91}) $\leq \exp(-k)$ if
1751: $\alpha_n \geq -2/\log\epsilon_n$.  
1752: 
1753: Using these codes, Alice and Bob construct $\cP_{nk}''$ as follows
1754: (with steps 1-3 performed coherently).
1755: % 
1756: \vspace*{-2ex} 
1757: %
1758: \begin{enumerate} 
1759: % 
1760: \item[0.]  
1761: % 
1762: Let $(a^{\rm o}_1, \cdots, a^{\rm o}_l)$ be a vector of $l$ messages
1763: each of $C_1^{(n)}$ bits, and $(b^{\rm o}_1, \cdots, b^{\rm o}_l)$ be
1764: $l$ messages each of $C_2^{(n)}$ bits.
1765: % 
1766: \item 
1767: %
1768: Using her error correcting code, Alice encodes $(a^{\rm o}_1, \cdots,
1769: a^{\rm o}_l)$ in a valid codeword $\vec{a} = (a_1, \cdots, a_k)$ which
1770: is a $k$-vector.  Similarly, Bob generates a valid codeword $\vec{b} =
1771: (b_1, \cdots, b_k)$ using his code.  
1772: % 
1773: \item  
1774: Let $\vec \A_1 := \A_1^{\otimes k}$ denote a tensor product of $k$ input
1775: spaces each of $C_1^{(n)}$ qubits.  Similarly, $\vec \B_1:=
1776: \B_1^{\otimes k}$.  (We will also denote $k$ copies of $\A_{0,2,3,4}$,
1777: and $\B_{0,2,3,4}$ by adding the vector symbol.)
1778: % 
1779: Alice and Bob apply $(\cP_n')^{\otimes k}$ to $\ket{\vec{a}}_{\vec
1780: \A_1} |\vec{b}\>_{\vec \B_1}$; that is, in parallel, they apply
1781: $\cP_n'$ to each pair of inputs $(a_j, b_j)$.  The resulting state is
1782: a tensor product of states of the form given by \eq{phiab}:
1783: % 
1784: \be
1785: \bigotimes_{j=1}^k \lbL
1786: %
1787: \ket{a_j}_{\A_0}  \ket{b_j}_{\B_0}
1788: \sum_{a_j',b_j'} |b_j'\>_{\A_1} |a_j'\>_{\B_1}
1789: \; |\Gamma_{a_j \oplus a_j', b_j \oplus b_j'}\>_{\A_{2,3,4} \B_{2,3,4}} 
1790: \rbL .
1791: %
1792: \label{eq:Pn-parallel}
1793: \ee
1794: 
1795: Define $|\Gamma_{\va \oplus \va', \vb \oplus \vb'}\>_{\vec\A_{234}
1796: \vec\B_{234}} := \bigotimes_{j=1}^k |\Gamma_{a_j \oplus a_j', b_j
1797: \oplus b_j'}\>_{\A_{2,3,4} \B_{2,3,4}}$.  Then, \eq{Pn-parallel} can be
1798: written more succinctly as
1799: % 
1800: \be 
1801: % 
1802: \ket{\va}_{\vec \A_0} |\vb\>_{\vec \B_0} 
1803:  \sum_{\va',\vb'} |\vb'\>_{\vec \A_1} \ket{\va'}_{\vec \B_1}
1804: |\Gamma_{\va \oplus \va', \vb \oplus \vb'}\>_{\vec\A_{234} \vec\B_{234}} \,.
1805: % 
1806: \ee
1807: 
1808: \item
1809: Alice performs the error correction step on $\vec \A_1$ and Bob
1810: does the same on $\vec \B_1$.  According to our code
1811: constructions, this (joint) step fails with probability $\pfail\leq
1812: 2\cdot 2^{-k}$.  (We will see below why $\pfail$ is
1813: independent of $\va$ and $\vb$.)
1814: 
1815: In order to describe the residual state, we now introduce $\cG_{\hs
1816: \A} = \{\vx \, {\in} \, [N_1]^k : |\vx| \, {\leq} \, k\alpha_n\}$ and
1817: % 
1818: $\cG_{\B} = \{\vx \,{\in}\, [N_2]^k : |\vx| \, {\leq} \, k\alpha_n\}$,
1819: where $|\vx|:=|\{j : x_j \,{\neq}\, 0\}|$ denotes the Hamming weight
1820: of $\vx$.  Thus $\cG_{\hs \A,\B}$ are sets of correctable (good)
1821: errors, in the sense that there exist local decoding isometries
1822: $\cD_{\hs \A},\cD_{\B}$ such that for any code word $\va\in [N_1]^k$
1823: we have $\forall \va'\in \va\oplus\cG_{\hs \A}, \cD_{\hs \A}\ket{\va'}
1824: = \ket{\va}\ket{\va\oplus \va'}$ (and similarly, if $\vb\in[N_2]^k$ is
1825: a codeword, then $\forall \vb'\in \vb\oplus\cG_{\B}, \cD_{\B} |\vb'\>
1826: = |\vb\> |\vb\oplus \vb'\>$).  For concreteness, let the decoding maps
1827: take $\vec \A_1$ to $\vec \A_1 \vec \A_5$ and $\vec \B_1$ to $\vec
1828: \B_1 \vec \B_5$.
1829: 
1830: Conditioned on success, Alice and Bob are left with
1831: \bea
1832: % 
1833: & & \frac{1}{\sqrt{1{-}\pfail}} \, |\va,\vb\>_{\vec \A_{0,1}} 
1834: 		|\va,\vb\>_{\vec \B_{0,1}} 
1835: \sum_{\va' \hs \in \va \oplus \cG_{\hs \A}} 
1836: \sum_{~\vb' \hs \in \vb \oplus \cG_{\B}}
1837: |\vb \oplus \vb'\>_{\vec \A_5} |\va \oplus \va'\>_{\vec \B_5}
1838: |\Gamma_{\va\oplus \va', \vb\oplus \vb'}\>_{\vec \A_{234} \vec \B_{234}}
1839: \\ &:=& \frac{1}{\sqrt{1{-}\pfail}} \, |\va,\vb\>_{\vec \A_{0,1}} 
1840: 		|\va,\vb\>_{\vec \B_{0,1}} 
1841:  \sum_{\va'' \hs \in\cG_{\hs \A}} \sum_{~\vb'' \hs \in \cG_{\B}}
1842: |\vb''\>_{\vec \A_5} |\va''\>_{\vec \B_5}
1843: |\Gamma_{\va'',\vb''}\>_{\vec \A_{234}\vec \B_{234}},
1844: \eea
1845: %
1846: \sloppypar{where we have defined $\va'' := \va\oplus \va'$ and $\vb''
1847:  := \vb\oplus\vb'$.  Note that $2^{-k+1} \geq \pfail =
1848: \sum_{(\va'',\vb'')\not\in \cG_{\hs \A} \times \cG_{\B}}
1849: \dblbraket{\Gamma_{\va'',\vb''}}$, which is manifestly independent
1850: of $\vec a,\vec b$.}
1851: % 
1852: The ancilla is now {\em completely} decoupled from the message,
1853: resulting in coherent classical communication.  The only remaining
1854: issue is recovering entanglement from the ancilla, so for the
1855: remainder of the protocol we ignore the now decoupled states $|\vec a,
1856: \vec b\>_{\vec\A_{0,1}} |\vec a,\vec b\>_{\vec\B_{0,1}}$.
1857: 
1858: \item
1859: For any $\vx$, define $S(\vx):=\{j : x_j \,{\neq}\, 0\}$ to be set of
1860: positions where $\vx$ is nonzero.  If $\vx\in\cG_{\hs \A}$ (or
1861: $\cG_{\B}$), then $|S(\vx)| \leq k\alpha_n$.  Thus, $S(\vx)$ can be
1862: written using $\leq \log \sum_{j \leq k\alpha_n} \!\! \binom{k}{j}
1863: \leq \log \binom{k}{k \alpha_n} 
1864: + \log (k\alpha_n) \leq kH_2(\alpha_n) + \log (k \alpha_n)$ bits.
1865: 
1866: The next step is for Alice to compute $|S(\vb'')\>$ from $|\vb''\>$
1867: and communicate it to Bob using $\lpm \! kH_2(\alpha_n) + \log (k\alpha_n) 
1868: \! \rpm \ctc$.  Similarly, Bob sends $\ket{S(\va'')}$ to Alice using
1869: $\lpm \! kH_2(\alpha_n) + \log (k\alpha_n) \! \rpm \cbc$.  Here
1870: we need to assume that some (possibly inefficient) protocol to send
1871: $O(k)$ bits in either direction with error $\exp(-k-1)$ (chosen for
1872: convenience) and with $Rk$ uses of $U$ for some constant $R$.  Such a
1873: protocol was given by \prop{U-nonzero-CC} and the bound on the error can be
1874: obtained from the HSW theorem\cite{Holevo98,SW97,Hayashi:02f}.
1875: 
1876: Alice and Bob now have the state
1877: % 
1878: \be \frac{1}{\sqrt{1{-}\pfail}} \,
1879: %\ket{\va,\vb}_A \ket{\va,\vb}^B 
1880: \sum_{\va''\in\cG_{\hs \A}} \sum_{\vb''\in\cG_{\B}}
1881: |S(\va'') S(\vb'')\>_{\vec\A_6} \, |\vb''\>_{\vec\A_5} \, 
1882: |S(\va'') S(\vb'')\>_{\vec\B_6} \, |\va''\>_{\vec\B_5} \,  
1883: |\Gamma_{\va'',\vb''}\>_{\vec\A_{234} \vec \B_{234}}.
1884: \ee
1885: % 
1886: Conditioning on their knowledge of $S(\va''),S(\vb'')$, Alice and Bob
1887: can now identify $k'\geq k(1-2\alpha_n)$ positions where
1888: $a_j''=b_j''=0$, and extract $k'$ copies of
1889: $\smfrac{1}{\sqrt{1{-}\pfail}}|\Gamma_{00}\>$.
1890: % 
1891: Note that leaking $S(\va''), S(\vb'')$ to the environment will not
1892: affect the extraction procedure, therefore, coherent computation and
1893: communication of $S(\va''), S(\vb'')$ is unnecessary.  (We have not
1894: explicitly included the environment's copy of $|S(\va'') S(\vb'')\>$ 
1895: in the equations to minimize clutter.) 
1896: % 
1897: After extracting $k'$ copies of
1898: $\smfrac{1}{\sqrt{1{-}\pfail}}|\Gamma_{00}\>$, we can safely discard
1899: the remainder of the state, which is now completely decoupled from both
1900: $\lbm \!  \smfrac{1}{\sqrt{1{-}\pfail}}|\Gamma_{00}\> \! \rbm
1901: ^{\otimes k'}$ and the message $|\vec a\>_{\A_0} |\vec b\>_{\A_1}
1902: |\vec b\>_{\B_0} |\vec a\>_{\B_1}$.
1903: 
1904: \item
1905: Alice and Bob perform entanglement concentration ${\cal E}_{\rm conc}$
1906: (using the techniques of \mscite{BBPS96}) on
1907: $\lbm \! \smfrac{1}{\sqrt{1{-}\pfail}}|\Gamma_{00}\> \! \rbm^{\otimes k'}$.  
1908: Note that since $\smfrac{1}{\sqrt{1{-}\pfail}}|\Gamma_{00}\>$ can be
1909: created using $U$ $n$ times and then using classical communication and
1910: postselection, it must have Schmidt rank $\leq {\rm Sch}(U)^n$, where
1911: ${\rm Sch}(U)$ is the Schmidt number of the gate $U$.
1912: Also recall that $E \lbm \! \smfrac{1}{\sqrt{1{-}\pfail}}|\Gamma_{00}\> 
1913: \!\rbm \geq C_1^{(n)} + C_2^{(n)} + \log (1{-}\eps_n)$.
1914: % 
1915: According to \mscite{BBPS96}, ${\cal E}_{\rm conc}$ requires no
1916: communication and with probability 
1917: % 
1918: $\geq 1 - \exp \lbm{-}{\rm Sch}(U)^n \lpm\! \sqrt{k'}-\log(k'{+}1) \!\rpm 
1919: \rbm$
1920: % 
1921: produces at least $k' \lbm C_1^{(n)} {+} C_2^{(n)} {+} \log (1{-}\eps_n) \rbm 
1922: - {\rm Sch}(U)^n \lbm \sqrt{k'} {-} \log(k'{+}1) \rbm$ ebits.
1923: 
1924: \end{enumerate}
1925: 
1926: \subsubsection{Error and resource accounting} 
1927: 
1928: $\cP_{nk}''$ consumes a total of \\[1ex]
1929: $~~~$(0) $nk$ uses of $U$ (in the $k$ executions of $\cP_n'$) \\
1930: $~~~$(1) $Rk$ uses of $U$
1931: 	(for communicating nontrivial syndrome locations) \\
1932: $~~~$(2) $k \lbm \!\! C_1^{(n)} {+} C_2^{(n)} \!\! \rbm \qq$  
1933: 	(for the encryption of classical messages). \\[1ex]
1934: % 
1935: $\cP_{nk}''$ produces, with probability and fidelity no less than 
1936: $$1{-}2^{{-}(k-1)}{-}2^{{-}(k-1)}-\exp \lbm \!\!  {-}{\rm Sch}(U)^n \lpm \!\!
1937: \sqrt{k'} {-} \log(k'{+}1) \!\!\rpm \!\!\rbm ,$$
1938:  at least \\[1ex]
1939: % 
1940: $~~~$(1) $l \, C_1^{(n)}\cof +l \, C_2^{(n)}\cob$  
1941: \\
1942: $~~~$(2) $k' \lpm \! C_1^{(n)}{+}C_2^{(n)} {+} \log (1{-}\eps_n)\! \rpm 
1943: - {\rm Sch}(U)^n \lpm \! \sqrt{k'}{-}\log(k'{+}1) \! \rpm \qq$ . 
1944: 
1945: We restate the constraints on the above parameters: 
1946: % 
1947: $\epsilon_n, \delta_n {\; \ra \; } 0$ as $n{\; \ra\; }\infty$;
1948: % 
1949: $C_1^{(n)} \geq n(C_1{-}\delta_n)$, $C_2^{(n)} \geq
1950: n(C_2{-}\delta_n)$;
1951: % 
1952: $\alpha_n \geq \max(1/C_1^{(n)}, 1/C_2^{(n)}, -2/\log\epsilon_n)$; 
1953: % 
1954: $k' \geq k(1{-}2\alpha_n)$; $l \geq k(1{-}3\alpha_n)$. 
1955:  
1956: We define ``error'' to include both infidelity and the probability of
1957: failure.  To leading orders of $k,n$, this is equal to $2^{{-}(k-2)} +
1958: \exp \lbm \!\!  {-} \sqrt{k} \; {\rm Sch}(U)^n \! \! \rbm $.  We
1959: define ``inefficiency'' to include extra uses of $U$, net consumption of entanglement, and the amount by which
1960: the coherent classical communication rates fall short of the classical
1961: capacities.  To leading order of $k,n$, these are respectively
1962: $Rk$, $2\alpha_nk(C_1^{(n)}{+}C_2^{(n)}) + \sqrt{k} \, {\rm Sch}(U)^n \approx 
1963:  2\alpha_nk n(C_1{+}C_2) + \sqrt{k} \, {\rm Sch}(U)^n$,
1964: and $nk(C_1{+}C_2) - l(C_1^{(n)} {+} C_2^{(n)}) \leq
1965: nk(3\alpha_n(C_1 {+} C_2) + 2\delta_n)$.  We would like the error to
1966: vanish, as well as the fractional inefficiency, defined as the
1967: inefficiency divided by $kn$, the number of uses of $U$.
1968: Equivalently, we can define $f(k,n)$ to be the {\em sum} of the error
1969: and the fractional inefficiency, and require that $f(k,n)\ra 0$ as
1970: $nk\ra \infty$.
1971: % 
1972: By the above arguments,
1973: % 
1974: \be f(k,n) \leq 2^{{-}(k-2)} + \exp(-\sqrt{k} \; {\rm Sch}(U)^n)
1975:  + 2\alpha_n(C_1 {+} C_2) + \smfrac{1}{n \sqrt{k}} \; {\rm Sch}(U)^n
1976:  + \frac{R}{n}
1977:  + 3\alpha_n(C_1 {+} C_2) + 2 \delta_n \,.
1978: \label{eq:frac-errors}\ee
1979: % 
1980: Note that for any fixed value of $n$, $\lim_{k\ra\infty} f(k,n) =
1981: 5\alpha_n(C_1{+}C_2)+2\delta_n + R/n$.  (This requires $k$
1982: to be sufficiently large and also $k \gg {\rm Sch}(U)^{2n}$.)  Now,
1983: allowing $n$ to grow, we have
1984: % 
1985: \be 
1986: 	\lim_{n\ra\infty}\lim_{k\ra\infty} f(k,n) = 0. 
1987: \ee 
1988: % 
1989: The order of limits in this equation is crucial due to the dependence 
1990: of $k$ on $n$. 
1991: 
1992: The only remaining problem is our catalytic use of 
1993: $O(nk)$ ebits.  In order to construct a protocol that uses only $U$,
1994: we need to first use $U$ $O(nk)$ times to generate the starting
1995: entanglement.  Then we repeat $\cP_n''$ $m$ times, reusing the same
1996: entanglement.  The catalyst results in an additional fractional
1997: inefficiency of $c/m$ (for some constant $c$ depending only on $U$)
1998: and the errors and 
1999: inefficiencies of $\cP_n''$ 
2000: add up to no more than $mf(k,n)$.  Choosing $m = \lfloor
2001: 1/\sqrt{f(k,n)} \rfloor$ will cause all of these errors and
2002: inefficiencies to simultaneously vanish.  (This technique is
2003: essentially equivalent to using Lemmas~\ref{lemma:noo} and
2004: \ref{lemma:cancel} and \thm{composability}.)  The actual error condition is
2005: that 
2006: % 
2007: \be
2008: \lim_{m\ra\infty}\lim_{n\ra\infty}\lim_{k\ra\infty} \; mf(k,n) +
2009: \frac{c}{m} \; = \; 0 \,.
2010: \ee 
2011: % 
2012: This proves the resource inequality
2013: \be U \geq C_1\cof + C_2\cob.\ee
2014: 
2015: \subsubsection{The $E<0$ and $E>0$ cases} 
2016: 
2017: If $E<0$ then entanglement is consumed in $\cP_n$, so there exists a
2018: sequence of integers $E^{(n)} \leq n(E+\delta_n)$ such that
2019: % 
2020: \be \cP_n \! \left( \ket{a}_{\A_1}\ket{b}_{\B_1}
2021: \ket{\Phi}^{E^{(n)}}_{\A_5 \B_5} \right)
2022: = 
2023: \sum_{a', b'} |b'\>_{\A_1} |a'\>_{\B_1} |\g_{a',b'}^{a,b}\>_{\A_2 \B_2}\,.
2024: \ee
2025: % 
2026: In this case, the analysis for $E^{(n)}=0$ goes through, only with
2027: additional entanglement consumed.  Almost all equations are the same,
2028: except now the Schmidt rank for $|\Gamma_{00}\>$ is upper-bounded by
2029: $\l[\Sch(U)2^{E+\delta_n}\r]^n$ instead of $\Sch(U)^n$.  This is still
2030: $\leq c^n$ for some constant $c$, so the same proof of correctness applies.
2031: 
2032: If instead $E>0$, entanglement is created, so for some
2033: $E^{(n)}\geq n(E-\delta_n)$ we have
2034: % 
2035: \be \cP_n \! \left( \ket{a}_{\A_1}\ket{b}_{\B_1} \right)
2036: = 
2037: \sum_{a', b'} |b'\>_{\A_1} |a'\>_{\B_1} |\g_{a',b'}^{a,b}\>_{\A_2 \B_2}\,.
2038: \ee
2039: % 
2040: for $E(|\g_{a,b}^{a,b}\>_{\A_2 \B_2}) \geq E^{(n)}$. 
2041: % 
2042: Again, the previous construction and analysis go through, with an
2043: extra $E^{(n)}$ ebits of entanglement of entropy in $|\Gamma_{00}\>$,
2044: and thus an extra fractional efficiency of $\leq 2\alpha_nE$ in
2045: \eq{frac-errors}.   The Schmidt rank of $|\Gamma_{00}\>$ is still
2046: upper bounded by Sch$(U)^n$ in this case. 
2047: \qed\bigskip
2048: 
2049: \begin{observation}
2050: If $(C_1,C_2,E)\in\CCE(U)$, but $(C_1,C_2,E+\delta)\not\in\CCE(U)$ for
2051: any $\delta>0$, then for any $\epsilon,\delta>0$ and for $n$
2052: sufficiently large there is a protocol $\cP_n$ and a state
2053: $\ket{\varphi}^{AB}$ on $\leq \kappa n\delta$ qubits (for a universal
2054: constant $\kappa$), such that for any
2055: $x\in\{0,1\}^{\lfloor n(C_1-\delta)\rfloor}, 
2056: y\in\{0,1\}^{\lfloor n(C_2-\delta)\rfloor}$ we have either
2057: $$ \cP_n \ket{x}^A\ket{y}^B \approx_\eps
2058: \ket{xy}^A\ket{xy}^B\ket{\Phi}^{\lfloor n(E-\delta)\rfloor}
2059: \ket{\varphi}$$
2060: if $E>0$ or
2061: $$ \cP_n \ket{x}^A\ket{y}^B \ket{\Phi}^{\lfloor -n(E-\delta)\rfloor}
2062: \approx_\eps
2063: \ket{xy}^A\ket{xy}^B\ket{\varphi}$$
2064: if $E<0$.
2065: 
2066: The key point here is that if $E$ taken to be the maximum possible for
2067: a given $C_1,C_2$, then the
2068: above proof of \thm{bidi-ccc} in fact produces ancilla systems of a
2069: sublinear size.
2070: \end{observation}
2071: 
2072: \section{Discussion}\label{sec:ccc-discuss}
2073: 
2074: Quantum information, like quantum computing, has often been studied
2075: under an implicit ``quantum co-processor'' model, in which quantum
2076: resources are used by some controlling classical computer.  Thus, we
2077: might use quantum computers or quantum channels to perform classical
2078: tasks, like solving computational problems, encrypting or
2079: authenticating a classical message, demonstrating nonlocal classical
2080: correlations, synchronizing classical clocks and
2081: so on.  On the other hand, since the quantum resources are manipulated
2082: by a classical computer, it is natural to think of conditioning
2083: quantum logical operations on classical information.
2084: 
2085: This framework has been quite useful for showing the strengths of
2086: quantum information relative to classical information processing
2087: techniques; e.g. we find that secure communication is possible,
2088: distributed computations require less communication and so on.
2089: However, in quantum Shannon theory, it is easy to be misled by the
2090: central role of classical information in the quantum co-processor
2091: model.  While classical communication may still be a useful {\em goal}
2092: of quantum Shannon theory, it is often inappropriate as an
2093: intermediate step.  Rather, we find in protocol after protocol that
2094: coherently decoupled cbits are better thought of
2095: as cobits.
2096: 
2097: Replacing cbits with cobits has significance beyond merely improving
2098: the efficiency of quantum protocols.  In many cases, cobits give rise
2099: to asymptotically reversible protocols, such as coherent teleportation
2100: and super-dense coding, or more interestingly, remote state
2101: preparation and HSW coding.  The resulting resource equalities go a
2102: long way towards simplifying the landscape of quantum Shannon theory: (1)
2103: The duality of teleportation and super-dense coding resolves a
2104: long-standing open question about how the original forms of these
2105: protocols could be individually optimal, but wasteful when composed;
2106: we now know that all the irreversibility from composing teleportation
2107: and super-dense coding is due to the map $\cof\geq\ctc$. (2) Coherent
2108: RSP and HSW coding give a resource equality that allows us to easily derive
2109: an expression for unitary gate capacity regions.  In the next chapter,
2110: we will see more examples of how making classical communication
2111: coherent leads to a wide variety of optimal coding theorems.
2112: 
2113: Although the implications of coherent classical communication are
2114: wide-ranging, the fundamental insight is quite simple: when studying
2115: quantum Shannon theory, we should set aside our intuition about the
2116: central role of classical communication, and instead examine carefully
2117: which systems are discarded and when communication can be coherently
2118: decoupled.
2119: