1: \documentclass[11pt]{article}
2: \usepackage{fullpage,proof,amssymb,amsmath,epsfig,stmaryrd}
3: \include{macros}
4: \bibliographystyle{alpha}
5: \begin{document}
6:
7: \title{Discrete Quantum Causal Dynamics}
8:
9: \author{
10: \begin{tabular}[t]{c}
11: Richard F. Blute\thanks{
12: Research supported in part
13: by NSERC.
14: }\\Ivan T. Ivanov\\
15: {\small Department of Mathematics}\\
16: {\small \rule{0mm}{3mm} and Statistics}\\
17: {\small University of Ottawa}\\
18: {\small Ottawa, Ontario, Canada}\\
19: \texttt{rblute,iti@mathstat.uottawa.ca}\\
20: \end{tabular}
21: \and
22: \begin{tabular}[t]{c}
23: Prakash Panangaden\footnotemark[1]\\
24: {\small School of Computer Science}\\
25: {\small McGill University}\\
26: {\small Montr\'eal, Qu\'ebec, Canada}\\
27: \texttt{prakash@cs.mcgill.ca}
28: \end{tabular}
29: }
30:
31: \maketitle
32:
33: \renewcommand{\girpar}{\bindnasrepma}
34:
35:
36: \begin{abstract} We give a mathematical framework to
37: describe the evolution of an open quantum systems subjected
38: to finitely many interactions with classical apparatuses. The
39: systems in question may be composed of distinct, spatially
40: separated subsystems which evolve independently but may also
41: interact. This evolution, driven both by unitary operators
42: and measurements, is coded in a precise mathematical
43: structure in such a way that the crucial properties of
44: causality, covariance and entanglement are faithfully
45: represented. We show how our framework may be expressed
46: using the language of (poly)categories and functors.
47: Remarkably, important physical consequences - such as
48: covariance - follow directly from the functoriality of our
49: axioms.
50:
51: We establish strong links between the physical
52: picture we propose and linear logic. Specifically we show
53: that the refined logical connectives of linear logic can be
54: used to describe the entanglements of subsystems in a
55: precise way. Furthermore, we show that there is a precise
56: correspondence between the evolution of a given
57: system and deductions in a certain formal logical system
58: based on the rules of linear logic.
59:
60: This framework generalizes and enriches both
61: causal posets and the histories approach to quantum
62: mechanics.
63: \end{abstract}
64:
65:
66:
67: \section{Introduction}
68:
69: We propose a uniform scheme for describing a quantum system,
70: interacting with a network of classical objects. The
71: system in question may be composed of distinct spatially
72: separated subsystems which evolve independently, but may
73: also interact with each other at various points as well as
74: with the classical objects. When analyzing physical
75: laboratory experiments on quantum systems, we frequently
76: abstract away from the concrete experimental setup and from
77: the particular details of the machinery involved. What we
78: usually keep is the description of the quantum system - and
79: its spatially separated subsystems - in terms of wave
80: functions or density matrices and unitary operators as well
81: as the changes of the quantum system induced by the
82: interactions with classical devices. Crucial properties of
83: the evolution such as the causal ordering, covariance of the
84: description for different observers and quantum entanglement
85: between distinct subsystems should be completely reflected
86: in any such description.
87:
88: The basis of our representation is the graph of events and
89: causal links between them. An event could be one of the
90: following: a unitary evolution of some subsystem, an
91: interaction of a subsystem with a classical device (a
92: measurement) or perhaps just the coming together or
93: splitting apart of several spatially separated subsystems.
94: Events will be depicted as vertices of a directed graph. The
95: edges of the graph will represent the causal relations
96: between the different events. The vertices of the graph are
97: then naturally labelled with operators representing the
98: corresponding processes.
99:
100: Of course, the processes of unitary evolution and
101: measurement take a certain amount of time; but we are only
102: interested in the causal relations between such events and
103: this allows us to consider them as point-like vertices on the
104: graph. Thus we are thinking of the duration between events
105: as being longer than the duration of an event so that no
106: causal information is lost when we represent interactions as
107: events.
108:
109: The structure described thus far reflects the kinematical
110: properties of the quantum system. To describe the dynamics
111: we need a composition of the operators assigned to the
112: vertices of the graph. This composition is most conveniently
113: described in terms of a composition in a specific
114: mathematical structure, namely a polycategory generated by
115: the graph. The whole description could then be concisely
116: summarized by noticing that we have a functor from this
117: polycategory to the polycategory of Hilbert spaces. This
118: functor captures the dynamics of the system.
119:
120: Causal relations are made explicit and we prove that no
121: influences breaking causality arise in our scheme. The
122: possible entanglement between spatially separated subsystems
123: - represented by distinct edges of the graph - is also
124: accounted for. Thus, our framework allows one to represent
125: locality of interaction - i.e.\ causal influences do not
126: propagate outside the causal ``cone'' - while allowing the
127: expression of nonlocal correlations which occur when one has
128: quantum entanglement. The tension between causal evolution
129: and quantum entanglement is resolved.
130:
131: The categorical framework that we use is intimately
132: connected with linear logic. Linear logic was
133: originally introduced~\cite{Girard87} as a logic intended
134: for a finer analysis of the way resources are consumed
135: during the course of a proof. This logic has had a
136: significant impact on the theory of computation as well as
137: such far-flung areas as linguistics and pure mathematics.
138: In the present paper, the connectives of linear logic will
139: be used to express the existence or nonexistence of nonlocal
140: correlations. What we will introduce is a deductive system
141: based on the graph-theoretic structure of the system
142: that precisely picks out the spatial slices of physical
143: interest. Thus evolution of the system corresponds to
144: logical deductions within this deductive system.
145: For an expository introduction
146: to linear logic, see the review by Girard~\cite{Girard95} or
147: the brief exposition in the appendix.
148:
149: \subsection{Relation to other work}
150:
151: Next we outline the relations of our proposal to some
152: recent approaches to quantum mechanics and quantum gravity.
153:
154: \subsubsection{Consistent and decoherent histories}
155:
156: The \emph{consistent histories} approach to quantum
157: mechanics due to Griffiths and Omn\`es
158: \cite{Griffiths96,Omnes94} was formulated with the aim of
159: shedding new light on the conceptual difficulties of the
160: theory. A closely related proposal with different
161: motivation is the \emph{decoherent histories} approach to
162: quantum cosmology of Gell-Mann and Hartle
163: \cite{Gell-Mann93}. The basic ingredient in both approaches
164: is the notion of a \emph{history} of the quantum system
165: described by a sequence of projection operators in the
166: Hilbert space of the system, for a succession of times. The
167: goal of quantum mechanics is to determine the probability of
168: an event or a sequence of events, thus one might hope to
169: assign probabilities to the histories of the quantum
170: system. In order for the probabilities to be additive in
171: the usual sense, the histories have to be mutually
172: noninterfering. Sets of histories obeying this condition
173: are selected with the use of a special bilinear form on
174: histories - the decoherence functional.
175:
176: A particular history is mathematically represented as a
177: linearly ordered sequence of projection operators in the
178: Hilbert space of the quantum mechanical system. But the
179: linear causal ordering of the events in a history is too
180: restrictive in many experimental situations, in particular
181: when analyzing spatially separated entangled quantum
182: systems. This issue is even more pressing for quantum
183: cosmology considerations. An application of the histories
184: approach to quantum field theory on a curved space-time
185: \cite{Blencowe91} must assume the existence of a globally
186: hyperbolic manifold, and thus via the associated foliation,
187: a linear ordering of the histories of the quantum field.
188:
189: Our proposal for describing the evolution of an open quantum
190: system can be considered as describing a single history in a
191: set of histories. The important point is that events are no
192: longer linearly ordered by temporal order but, rather,
193: partially ordered with respect to the causal order. This
194: allows one to capture the notion of causal evolution in a
195: manifestly covariant fashion. The consistency/decoherence
196: condition for histories has an immediate generalization for
197: histories described by more general graphs as proposed here.
198:
199: \subsubsection{Causal sets}
200:
201: \emph{Causal sets} form the basis of an approach to quantum
202: gravity mainly advocated by R. Sorkin and collaborators
203: \cite{Bombelli87,Sorkin91}, where the basic idea is to take
204: the notion of causality as the primitive. In classical
205: relativity, the structure of the space-time manifold
206: together with a metric of Lorentzian signature determines
207: the causality relation. An important observation is that
208: the causal structure is conformally invariant, i.e.\
209: determined by only the conformal equivalence class of the
210: metric and hence more primitive than the metric. Various
211: proposals for quantum gravity - for example, the twistor
212: program~\cite{Penrose72} - have taken as their point of
213: departure the idea that the causal structure is more
214: fundamental than the metric structure.
215:
216: In the causal sets approach, one takes the point of view
217: that, at the smallest length scales, spacetime is inherently
218: discrete and that the causal structure, the ``light cones'',
219: are fundamental. This leads naturally to the idea of a
220: partially ordered set (poset for short) where the elements
221: are events and two events are related by causality. The
222: main interest is in approximating continuous spacetimes with
223: such structures and defining processes that would generate
224: these structures, with a view to an eventual theory of
225: quantum gravity. Though the aims are rather different the
226: issues connected with causality are closely related.
227:
228: Causal sets are further motivated by the idea that a
229: discrete structure would avoid the singularities that plague
230: physics (both classical and quantum). The assumption that
231: space-time should be a continuous manifold is one of the
232: ingredients that leads to the problematic singularities of
233: quantum field theory and general relativity. In the causal
234: sets approach, space-time is a discrete structure, thus
235: possibly avoiding these singularities, the idea being that
236: at the Planck scale, continuous geometry gives way to
237: discrete geometry.
238:
239: One way to think of this is that one approximates a manifold
240: as one ``sprinkles'' more and more points into the causal
241: set in a uniform fashion. More formally, one would want to
242: obtain a manifold as the categorical limit of a diagram of
243: posets and embeddings \cite{Maclane98}. Applications and
244: extensions of these ideas can be found in papers such as
245: \cite{Markopoulou00,Markopoulou97,Raptis00b}, although this
246: list is by no means exhaustive. In our approach we are not
247: thinking about generating the spacetime through such
248: limiting processes, but the idea of a causal set is implicit
249: in our work. For us, a finite causal set is the kinematical
250: framework on which we describe evolution and information
251: flow.
252:
253: \subsubsection{Quantum causal histories}
254:
255: The notion of \emph{quantum causal history} was introduced
256: by Markopoulou in \cite{Markopoulou00}. One begins with
257: a poset (causal set) and assigns Hilbert spaces to the
258: vertices and evolution operators to sets of edges. The
259: assignment must satisfy properties analogous to
260: functoriality. However, within this framework, one is
261: quickly led to violations of causality - as the author
262: herself notes - essentially because the slices used are
263: ``too global.'' She mentions the possibility of working
264: with a dual view. In fact, in our work, we take such a
265: dualized view as our starting point. In other words we
266: assign operators representing evolution or measurement to
267: vertices and Hilbert spaces to the edges, in a way
268: satisfying (poly)functoriality.
269:
270: \subsection{The Importance of Categories} A category can be
271: seen as a generalization of a poset in the following sense.
272: A poset merely records that an element $x$ is less than $y$
273: but a category keeps track of the different ways in which
274: $x$ might be less than
275: $y$. For example, in logic one might consider formulas
276: (denoted by Greek letters like $\phi$, $\psi$ etc.\ ) and
277: the relationship of provability between them. Thus one
278: would write $\phi\vdash\psi$ to mean that starting from the
279: assumption $\phi$ one can prove $\psi$. This gives rise to
280: a transitive and reflexive relation; if one considers
281: equivalence classes of formulas (two formulas being
282: equivalent if each can be used to prove the other) we get a
283: poset. However, if we are interested in distinguishing
284: distinct proofs we need to keep track of the different ways
285: in which $\phi$ can be used to prove $\psi$. Thus formulas
286: as objects and proofs as morphisms can be organized into a
287: category.
288:
289: In a poset when one writes $x\leq y$ then, depending on the
290: context, one is stating something like the following:
291: \begin{itemize}
292: \item $x$ is less than $y$;
293: \item $x$ precedes $y$;
294: \item $x$ implies $y$.
295: \end{itemize}
296: \noindent or any of several other possibilities. In a
297: causal set, we have in mind that $x$ causally precedes $y$.
298:
299: In the present work, we are particularly interested in
300: modelling the idea that information can flow from one event
301: to another in a number of different ways, \emph{along
302: different paths or channels}. We would like to keep track
303: of all these various independent paths. The structure of a
304: poset is inadequate for achieving this, as we would like to
305: say that $x\leq y$ in several different ways. This
306: naturally suggests that we pass from posets to more general
307: graphs and eventually to categories.
308:
309: Many recent experiments feature spatially distributed
310: quantum systems. When entangled quantum subsystems come back
311: together in the same spacetime region, the description of the
312: resulting system is causally influenced by all events in the
313: paths of the subsystems. In particular a past event could
314: influence the future events in several distinct ways through
315: different paths. Our scheme is well adapted for analyzing
316: experiments featuring spatially separated quantum entangled
317: entities and could be used in the field of quantum
318: information processing to analyze information flow
319: situations.
320:
321: \subsection{Contents of the present paper}
322:
323: Section~\ref{causal} presents the basic ideas of our scheme
324: via an example. Section~\ref{dyn} discusses the basic
325: physical ideas involved. In the first subsection we review
326: the notions of measurements and interventions. In the next
327: subsection we give the dynamical prescription in a special
328: case and in the final subsections we give the general
329: prescription and prove covariance. In section~\ref{poly} we
330: review basic facts about polycategories and their
331: construction. We also describe the polycategory of Hilbert spaces
332: and intervention operators
333: we will be using. In section~\ref{logic} we give a logical
334: presentation of polycategories and establish the connection
335: between our structures and linear logic. A functorial version of
336: our dynamical prescription is then presented.
337: We end with a discussion on further
338: applications of our scheme.
339: It is our hope that this paper will
340: interest members of several different communities within
341: mathematics, logic and physics.
342:
343: \section{Causal information flow via
344: examples}\label{causal}
345:
346: Consider a quantum system evolving in space-time while being
347: subjected to interactions with classical observers at a
348: number of points. The causal and spatio-temporal relations
349: in the system will be represented by a directed acyclic
350: graph (hereafter called a \emph{dag}). The vertices of the
351: graph - which will be drawn as boxes - represent the events
352: in the evolution of the system. An event could be a
353: measurement by a classical observer, a local unitary
354: evolution or just a splitting of a subsystem into several
355: spatially separated subsystems, which however could still
356: share an entangled common state. The propagation of the
357: different subsystems will be indicated by the edges of the
358: graph.
359:
360: There are a number of causal relations between edges and
361: vertices. A vertex $v_1$ is said to \emph{immediately
362: precede} $v_2$ if there is a (directed) edge from $v_1$ to
363: $v_2$. We write $v_1 \leq v_2$ for the reflexive transitive
364: closure of immediate precedence; thus $v \leq v$ always
365: holds and $v_1 \leq v_2$ means that there is a
366: \emph{directed path} from $v_1$ to $v_2$ (possibly of length
367: zero). When $v_1 \leq v_2$ we sometimes say $v_1$ is ``to
368: the past of'' $v_2$ and dually ``$v_2$ is to the future of
369: $v_1$.'' When we draw a poset we typically leave out the
370: self-loops and only draw the minimal number of edges needed
371: to infer all the others; the so-called ``Hasse diagram'' of
372: the poset. We note that our graphs will have initial and
373: final ``half-edges'', i.e. edges with only one endpoint.
374: Physically we have some quantum states incoming (or
375: ``prepared'') followed by some interactions and some
376: outgoing state.
377:
378: The relation between vertices induces a causal relation
379: between edges. We say that an edge $e_1$ is to the past of
380: another edge $e_2$ if the terminal vertex of $e_1$, say
381: $v_1$ and the initial vertex of $e_2$, say $v_2$, satisfy
382: $v_1 \leq v_2$. Note that we could have $v_1 = v_2$. An
383: initial edge is not to the future of any edge, nor is a
384: final edge to the past of any other edge. If two edges are
385: not causally related, we say that they are ``spacelike
386: separated'' or acausal. Note that two spacelike separated
387: edges could share a common terminal vertex or a common
388: initial vertex, (but since we have a graph, not both). A
389: \emph{space-like slice} is defined as a set of pairwise
390: acausal edges. Henceforth, whenever we say ``slice'' we
391: will always mean ``spacelike slice.'' Note that the initial
392: (or final) edges form a spacelike slice. We call this the
393: \emph{initial (final) slice}.
394:
395: For example for the graph of Figure~\ref{figureN} the set of
396: edges $\{e_c, e_d, e_e\}$ form a space-like slice. Another
397: example is the set $\{e_f, e_d, e_e\}$. The edges $e_a$ and
398: $e_b$ form the initial slice. The edges $e_a,e_b,e_f$ and
399: $e_g$ are half-edges, with $e_a$ and $e_b$ initial, and
400: $e_f$ and $e_g$ final.
401:
402: \begin{figure}[htb]
403: \begin{center}
404: \input{figure1}
405: \end{center}
406: \caption{}
407: \label{figureN}
408: \end{figure}
409:
410: Associated with any edge $e_i$ is an observer who has access
411: to a subsystem of the complete quantum system. Thus the
412: edges represent local information. Each edge $e_i$ is
413: assigned a density matrix $\rho_i$ in a Hilbert space
414: $\hi_i$\footnote{Throughout the paper, we assume that the
415: graph and the dimensions of all Hilbert spaces are
416: finite.}. The density matrix $\rho_i$ describes the
417: knowledge about the quantum system available to the local
418: observer at the edge $e_i$. More generally density matrices
419: will be associated to space-like slices. For a space-like
420: slice consisting of edges $\{e_{i_1}, \dots e_{i_p}\}$, the
421: assigned density matrix will be denoted $\rho_{i_1, \dots
422: i_p}$. This density matrix describes the subsystem of the
423: whole quantum system for that space-like slice. Every
424: space-like slice has also a Hilbert space which is the
425: tensor product of the Hilbert spaces of the edges forming
426: the slice. However the density matrix associated with the
427: slice is not in general a tensor product of the density
428: matrices on the edges. If it were, we could not capture
429: non-local quantum correlations.
430:
431: The graph of Figure~\ref{figureN}, represents a quantum
432: system $\mathit{Q}$ which starts evolving from a state in
433: which $\mathit{Q}$ consists of two spatially separated
434: subsystems $\mathit{Q}_a$ and $\mathit{Q}_b$ described by
435: density matrices $\rho_a$ and $\rho_b$ respectively, in
436: Hilbert spaces
437: ${\hi}_a$ and ${\hi}_b$. The initial edges $e_a$ and $e_b$
438: form the initial slice in this simple system. We will
439: follow the convention that if the initial slice consists of
440: several edges, the initial state of the whole system is a
441: tensor product state, i.e. the subsystems are not
442: entangled. For the above example, $\psi_{init} = \psi_a
443: \otimes \psi_b$ and $\rho_a =
444: \proj{\psi_a}$ and $\rho_b = \proj{\psi_b}$. Entangled
445: subsystems on distinct edges will always have at least one
446: event in the common past. Thus we always explicitly
447: represent the interaction which caused the entanglement.
448:
449: Each vertex $v_i$ of the graph is labelled with an operator
450: $T_i$ which describes the process taking place at the
451: corresponding event. The operator $T_i$ at a given event
452: $v_i$ takes density matrices on the tensor product of
453: Hilbert spaces living on the incoming edges at $v_i$ to
454: density matrices on the tensor product Hilbert space of
455: outgoing edges. The process at a vertex could be an
456: \emph{intervention}\footnote{Interventions are generalized
457: measurements where a quantum subsystem could be
458: discarded~\cite{Peres00}. This will be discussed more fully
459: below.} corresponding to a positive operator-valued measure
460: (POVM)~\cite{Nielsen00,Peres95} or a unitary
461: transformation. Or instead of an external or unitary action
462: there could be several quantum subsystems that come together
463: and then split apart, possibly in a different way. We will
464: consider this last case as a particular instance of a
465: unitary evolution with identity evolution operator. As a
466: simple example, in the case of an event corresponding to
467: unitary evolution by a unitary operator
468: $U$, we have the usual expression:
469:
470: \begin{equation}\label{inop}
471: \rho^{in} \ \mapsto \ \rho^{out} \ = \
472: U \rho^{in} U^\dagger
473: \end{equation}
474:
475: The general expression for an operator associated to an
476: event will be discussed fully in the next section, see
477: equation (\ref{inap}).
478:
479: Here we will discuss some of the conditions such a dynamical
480: scheme has to satisfy in order to reflect causality and
481: other physical properties of the quantum system. Causality
482: is the condition that the density matrix on a given edge
483: should not depend on the actions performed at vertices which
484: are acausal to this edge or are in its future. For example,
485: referring back to Figure~\ref{figureN}, we would like any
486: quantum evolution rule to say that the density matrix at
487: $e_g$ is unaffected by the intervention at $v_3$ or the
488: density matrix at $e_f$ is unaffected by the intervention at
489: $v_2$. A general unitary evolution between the states of
490: two space-like slices is easily shown to violate this
491: condition. Therefore we need to incorporate some sort of
492: locality condition into the evolution scheme.
493:
494: It is not hard to formulate such an evolution scheme. For
495: example, one could work with the dual picture and have
496: evolution occur along edges with density matrices at the
497: vertices. It is not hard to formulate rules which would
498: enforce causality properly in such a framework.
499: Unfortunately this rules out quantum correlations across
500: spatially separated subsystems. Thus, the evolution scheme
501: cannot be too local because entangled subsystems of the
502: quantum system could fly apart and later come together at a
503: vertex.
504:
505: Consider the system shown in Figure~\ref{figureD}.
506: \begin{figure}[htb]
507: \begin{center}
508: \input{figure2}
509: \end{center}
510: \caption{}
511: \label{figureD}
512: \end{figure} The quantum system represented in this graph is
513: as follows. The system is prepared in a state $\psi_a$ as
514: indicated by the density matrix $\rho_a =
515: \proj{\psi_a}$ on the incoming edge. At the vertex $v_1$
516: the system splits into two spatially separated subsystems on
517: the edges $e_b$ and $e_c$ which, in general, are still
518: described by a global entangled state. The local
519: transformations $T_2$ and $T_3$ will, in general, preserve
520: the entanglement and the global state will be still
521: entangled on the space-like slice
522: $\{e_d, e_e\}$. The two subsystems come together at the
523: vertex $v_4$. The two local density matrices $\rho_d$ and
524: $\rho_e$ are not sufficient to reconstruct the entangled
525: state of the system described by $\rho_f$. The off-diagonal
526: terms of $\rho_f$ are not reflected in the local density
527: matrices, $\rho_d$ and $\rho_e$. We need to include
528: information about the history of the state on the space-like
529: slice $\{e_d, e_e\}$ in order to reconstruct the global
530: state. One possibility is to work with global space-like
531: slices, and show that the scheme is generally covariant in
532: the sense of being slice-independent. In our functorial
533: approach, certain preferred (not necessarily global)
534: spacelike slices account for all entanglement.
535:
536: The rules for constructing and labeling the graphs given so
537: far reflect the kinematics of the quantum system.
538: Specifying the dynamics amounts to a prescription for how to
539: obtain the density matrices on every edge from the density
540: matrix on the initial slice and the operators at the
541: vertices of the graph. This prescription will be given
542: below in section~\ref{dyn}.
543:
544: \section{Dynamics on Graphs}\label{dyn}
545: \subsection{Measurements and Interventions} We begin with
546: some standard material on density matrices and positive
547: operator-valued measures (POVMs)
548: \cite{Nielsen00,Preskill98}, before introducing Peres'
549: notion of
550: \emph{intervention operator} \cite{Peres00}.
551:
552: Density matrices are used for describing quantum subsystems
553: which are part of larger quantum systems.
554: In particular a local observer who has
555: access only to a subsystem $Q_1$ of a quantum system $Q$
556: will associate a density matrix to his subsystem. Let
557: $\hi$ be the Hilbert space of state vectors of $Q$.
558:
559: If the overall system $Q$ is in a state described by a wave
560: function
561: $\ket{\psi} \in \hi$, then its density matrix is the
562: operator $\rho =
563: \proj{\psi} \in End(\hi)$. Since $Q$ can be decomposed into
564: subsystems, its Hilbert space is a tensor product $\hi =
565: \hi_{1} \otimes \hi_{2}$ of the Hilbert space $\hi_{1}$ of
566: the subsystem $Q_1$ and the Hilbert space
567: $\hi_{2}$ describing the remaining degrees of freedom. The
568: density matrix of the subsystem $Q_1$ is then given by a
569: partial trace with respect the Hilbert space $\hi_{2}$:
570: $\rho_{1} = Tr^{\hi_2} \rho$. If $\hi$ is any Hilbert
571: space, then the space of all density matrices will be denoted
572: $\mathsf{DM}(\hi)$.
573:
574:
575: The \emph{measurement} of a property of a quantum system
576: involves interaction with a classical apparatus. When a
577: classical apparatus measures an observable of a quantum
578: subsystem sitting inside a larger system the appropriate
579: mathematical formalism for such generalized measurement is
580: that of \emph{positive operator-valued measure} or POVM. Let
581: the possible outcomes of the measurement be labelled by the
582: letter $\mu
583: \in \{1 \dots N\}$. The measurement involves interaction
584: between the apparatus and the quantum system, described by a
585: unitary operator. The classical apparatus has a preferred
586: basis of states indexed by $\mu$. After the measurement,
587: the apparatus appears in one these preferred states. Since
588: we are only interested in describing our quantum subsystem
589: $Q_1$, we trace out all the remaining degrees of freedom.
590: Effectively to every outcome
591: $\mu$ is associated an operator $F_{\mu}$. The density
592: matrix of $Q_1$ after the measurement with outcome $\mu$ is
593: given by
594: \begin{equation}
595: \rho'_{\mu} = \frac{1}{p_{\mu}} F_{\mu} \rho
596: F^{\dagger}_{\mu}
597: \end{equation} where $\rho$ is the density matrix before the
598: measurement and $p_{\mu}$ is a numerical factor normalizing
599: the resulting density matrix to unit trace. Consider the
600: family of positive operators $E_{\mu} = F^{\dagger}_{\mu}
601: F_{\mu}
602: $. For a generalized measurement these have to satisfy the
603: condition
604: $\sum_{\mu} E_{\mu} = I$. The probability $p_{\mu}$ for
605: obtaining a measurement result labelled by $\mu$ is then
606: given by: $p_{\mu} = Tr E_{\mu} \rho$. This justifies the
607: name POVM.
608:
609: Even more general measurement processes could be considered
610: if the observer discards part of the quantum system during
611: the process of measurement. The appropriate mathematical
612: formalism for describing these generalized measurements is
613: that of {\it intervention operators} \cite{Peres00}. In
614: the process of measurement, the density matrix changes
615: according to:
616: \begin{equation}\label{inap}
617: \rho'_{\mu} \ = \
618: \frac{1}{p_{\mu}} \sum_m A_{\mu m} \ \rho \ A_{\mu
619: m}^{\dagger}
620: \end{equation} The families of maps $A_{\mu m}$ now act in
621: general from one Hilbert space to another, i.e for fixed
622: $\mu$ and $m$ they correspond to rectangular matrices.
623:
624: The label $\mu$ again distinguishes the set of possible
625: outcomes and the letter $m$ labels the degrees of freedom
626: discarded during this generalized measurement. Since the maps
627: $A_{\mu m}$ come from measurements realized by unitary
628: operator on some larger Hilbert space they again satisfy a
629: completeness condition: $\sum_{\mu m} A^{\dagger}_{\mu m}
630: A_{\mu m} = I$, where $I$ is the identity operator in the
631: appropriate Hilbert space. Notice that if the labels $\mu$
632: and $m$ are absent in (\ref{inap}) the equation describes
633: unitary evolution. Since the events we consider are
634: generalized measurements or unitary evolutions, equation
635: (\ref{inap}) is the appropriate mathematical representation
636: of those processes in full generality. Such maps
637: (\ref{inap}) on density matrices will be called {\it
638: intervention operators}.
639:
640: \subsection{The dynamical prescription}\label{dyna}
641:
642: We are now ready to start discussing the dynamics of a
643: quantum system represented by a dag $G$. Dynamics will be
644: described by supposing that we are given a density matrix on
645: the initial spacelike slice, and then giving a prescription
646: for calculating the density matrices of future spacelike
647: slices. In essence, we are propagating the initial data
648: throughout the system.
649:
650: To each vertex $i\in G$ will be assigned an operator $T_i$,
651: and to each edge $e_j$ will be assigned a Hilbert space
652: $\hi_j$. We note that all incoming (or outgoing) edges of a
653: given vertex are pairwise acausal and thus form a spacelike
654: slice. Thus there will be a density matrix
655: $\rho_i^{in}$ associated to the slice of the incoming
656: edges. Then one obtains the density matrix for the slice of
657: the outgoing edges by:
658: \[ \rho_i^{in}= T_i (\rho_i^{out}).\]
659:
660: Notice that more generally, for two acausal vertices, the sets of
661: incoming or outgoing edges are pairwise acausal. Thus, the
662: associated intervention operators will act on different
663: Hilbert spaces and hence commute.
664:
665: We begin with an illustrative example. Consider the dag of
666: Figure~\ref{fig3}.
667: \begin{figure}
668: \begin{center}
669: \epsfig{file=figure3.eps}
670: \end{center}
671: \caption{}
672: \label{fig3}
673: \end{figure} Given the state on the initial slice, the
674: operators at the events propagate the state to the future.
675: In the example of Figure~\ref{fig3} we have:
676: $\rho_c = T_1 (\rho_a)$,\ $\rho_{fde} = T_2 (\rho_b)$.
677: However the next intervention operator $T_3$ must act on the
678: so far undefined density matrix
679: $\rho_{cd}$. $T_3$ takes density matrices on
680: $\hi_c\ox\hi_d$ to those on
681: $\hi_g\ox\hi_h$. By extending
682: $T_3$ with the appropriate identity operators, we
683: can view it as a map from
684: $\mathsf{DM}(\hi_c\ox\hi_d\ox\hi_e\ox\hi_f)$ to
685: $\mathsf{DM}(\hi_e\ox\hi_f\ox\hi_g\ox\hi_h)$. Then we can
686: define the density matrix on another space-like slice,
687: namely $\rho_{fghe} = T_3 (\rho_c \otimes \rho_{fde})$.
688: Similarly $\rho_{fdi} = T_4 (\rho_{fde})$ and so on.
689: Starting from density matrices on the initial edges and
690: using the intervention operators associated with the
691: vertices - extended with identities as needed - we obtain
692: density matrices on specific space-like slices.
693:
694: The above inductive process for propagating density matrices
695: can be applied to any system described by a dag. However,
696: the procedure only gives the density matrices for certain
697: spacelike slices within the dag. For example, this
698: procedure does not yet yield a matrix for the slice $de$.
699: To calculate such density matrices, we will also have to
700: make use of the trace operator. Before extending the
701: procedure to such slices, we first consider those for which
702: the above process is sufficient. We call these slices
703: \textit{locative}.
704: \begin{defin}{\rm Let $G$ be a dag, and $L$ a slice of $G$.
705: Consider the set of all vertices $V$ which are to the past
706: of some edge in $L$. Let $I$ be the set of initial edges in
707: the past of $L$. Consider all paths of maximal length
708: beginning at an element of $I$ and only going through
709: vertices of $V$. Then $L$ is \emph{locative} if all such
710: paths end with an edge in $L$.}
711: \end{defin} In our example, the locative slices are the
712: following:
713: $$a, b, ab, c, cb, def, adef, cdef, efgh, adfi, cdfi, fghe,
714: fghi, fgk, hej, hij, jk$$ while, for example, $de$ is not
715: locative. Note that the fact that maximal slices are always
716: locative follows immediately from the definition of locative.
717:
718: We now describe the general rule for calculating the density
719: matrices on locative slices. Associated with each locative
720: slice $L$ is the set $I$ of initial edges in the past of
721: $L$. We choose a family of slices that begins with $I$ and
722: ends with $L$ in the following way. Consider the set of
723: vertices $V$ between the edges in $I$ and the edges in $L$.
724: Because $L$ is locative we know that propagating slices
725: forwards through the vertices in
726: $V$ will reproduce $L$. Let $M\subset V$ be such that the
727: vertices in $M$ are minimal in $V$ with respect to causal
728: ordering. We choose arbitrarily any vertex $u$ in $M$,
729: remove the incoming edges of $u$ and add the outgoing edges
730: of $u$ to the set $I$ obtaining a new set of edges $I_1$. It
731: is clear that $I_1$ is spacelike and locative. Proceeding
732: inductively in this fashion we obtain a sequence of slices
733: $I=I_0,I_1,I_2,\ldots,I_n = L$, where $n$ is the cardinality
734: of $V$. Of course, this family of slices is far from unique.
735:
736: The dynamics is obtained as follows. Recall that the states
737: on initial edges are assumed not to be entangled with each
738: other so that one can obtain the density matrix on any set
739: of initial edges, in particular $I$, by a tensor product.
740: Let $\rho_0$ be the density matrix on $I$. We look at the
741: vertex $u$ that was used to go from $I$ to $I_1$ and apply
742: the intervention operator $T$ assigned to this vertex -
743: possibly augmented with identity operators as in the example
744: above. Proceeding inductively along the family of slices,
745: we obtain the density matrix $\rho_n$ on $L$.
746:
747: The important point now is that $\rho_n$ does not depend on
748: the choice of slicing used in going from $I$ to $L$. This
749: can be argued as follows. Suppose we have a locative slice
750: $S$ and two vertices $u$ and $v$ which are both causally
751: minimal above $S$ and acausal with respect to each other.
752: Then we have four slices to consider, $S$, $S_u$, $S_v$ and
753: $S_{uv}$ where by $S_u$ we mean the slice obtained from $S$
754: by removing the incoming edges of $u$ and adding the
755: outgoing edges of $u$ to $S$ and similarly for the others.
756: It is clear, in this case, that the intervention operators
757: assigned to $u$ and to $v$ commute and the density matrix
758: computed on
759: $S_{uv}$ is independent of whether we evolved along the
760: sequence
761: $S\to S_u\to S_{uv}$ or $S\to S_v\to S_{uv}$. Now when we
762: constructed our slices at each stage we had the choice
763: between different minimal vertices to add to the slice. But
764: such vertices are clearly pairwise acausal and hence, by the
765: previous argument applied inductively, the evolution
766: prescription is independent of all possible choices.
767:
768: So far we have defined density matrices on locative slices
769: only. To define density matrices on general spacelike
770: slices we will need to consider partial tracing operations.
771:
772: \subsection{General Slices}
773:
774: Recall that when one has subsystems $Q_1$ and $Q_2$ of a
775: quantum system
776: $Q$, the Hilbert space for $Q$ may be decomposed as
777: $\hi_1\ox\hi_2$ where
778: $\hi_i$ represents $Q_i$. The density matrix for $Q_1$ is
779: obtained by tracing over $\hi_2$. To obtain a candidate for
780: the density matrix of a spacelike slice $L$, we should find
781: a locative slice $M$ that contains $L$ and trace over the
782: Hilbert spaces on edges in $M\setminus L$. Such a locative
783: slice $M$ always exists because maximal spacelike slices are
784: always locative. $M$ is not unique however, and thus - as
785: we did for locative slices - we must show that different
786: choices give the same result. To simplify the notation we
787: will discuss the case of density matrices associated with
788: single edges. The case of a general space-like slice is
789: similar.
790:
791: Consider an edge $e_i$ in a graph $G$. Let $V_i =
792: \{v_{i_1}, \dots, v_{i_p}\}$ be the set of vertices in the
793: past of $e_i$. Let $I_i =
794: \{e_{i_1}, \dots, e_{i_q}\}$ be the set of initial edges in
795: the past of
796: $e_i$. Constructing a sequence of slices by incrementally
797: incorporating the vertices of $V_i$ in a manner similar to
798: what we did in the previous subsection, we get a locative
799: slice $M_i$ containing $e_i$. Starting with the density
800: matrices on the edges of $I_i$ and applying the operators
801: associated with the vertices of $V_i$, we obtain the density
802: matrix on the locative slice $M_i$. It is clear that $M_i$
803: is in an evident sense the minimal locative slice containing $e_i$.
804:
805: \begin{defin}{\rm We shall refer to $M_i$ as the \emph{least
806: locative slice} of the edge $e_i$.}
807: \end{defin}
808:
809: Let the least locative slice $M_i$ of an edge $e_i$ consist
810: of edges $\{e_i, e_{j_1}, \dots, e_{j_r}\}$. The density
811: matrix $\rho_{i,j_1,
812: \dots, j_r}$ on $M_i$ is an element of the space $\ind(\hi_i
813: \otimes
814: \hi_{j_1} \otimes \dots \otimes \hi_{j _r})$. Let $Tr^{j_1
815: \dots j_r}$ be the partial trace operation $\ind(\hi_i
816: \otimes \hi_{j_1} \otimes \dots
817: \otimes \hi_{j_r}) \rightarrow \ind(\hi_i)$.
818: \begin{defin}[Density matrix associated with an
819: edge]\label{rho}\emph{The density matrix $\rho_i$ at the
820: edge $e_i$ is defined to be:
821: \begin{equation}
822: \rho_i \ = \ Tr^{j_1 \dots j_r} \ \rho_{i,j_1, \dots, j_r} .
823: \end{equation}}
824: \end{defin} If $M_i$ consists of the single edge $e_i$, then
825: no tracing is done.
826:
827: \begin{rem}
828: The causality condition for evolving the initial data on $G$
829: requires that the density matrix associated with a given
830: edge $e_i$ depends only on the initial data in the past of
831: $e_i$ and only those interventions to the past of $e_i$.
832: The density matrix $\rho_i$ as defined in~\ref{rho}
833: satisfies this requirement by construction and so our
834: prescription for dynamical evolution is causal.
835: \end{rem}
836:
837: In general, the edge $e_i$ is contained in many locative
838: slices and we could just as well have defined $\rho_i$ by
839: tracing over the complimentary degrees of freedom in any of
840: these locative slices. Independence of the resulting
841: density matrices is the discrete analog of Lorenz (or
842: general) covariance in our framework. To clarify the
843: discussion consider the quantum system represented by the
844: graph on Figure~\ref{figureF}.
845:
846: \begin{figure}[htb]
847: \begin{center}
848: \input{figure4}
849: \end{center}
850: \caption{}
851: \label{figureF}
852: \end{figure}
853:
854: Let the initial $\rho_a$ be the density matrix of a
855: maximally entangled state of two spin $1/2$ subsystems:
856: $\rho_a = |\psi_a\rangle\langle\psi_a|$, where $\psi_a =
857: 1/\sqrt{2} \ (\psau \otimes \psbu + \psad \otimes \psbd)$.
858: At the first vertex the two subsystems separate with no
859: classical intervention. Therefore $\rho_{bc} = \rho_a$.
860: The slice
861: $\{e_b, e_c\}$ is the least locative slice for the edge $e_b$
862: and we can compute the density matrix associated to this
863: edge: $\rho_b = T r^{c} \rho_{bc} = 1/2 \ (\psauk \psaub +
864: \psadk \psadb)$. Next, let the intervention at the second
865: vertex be a measurement on the corresponding subsystem with
866: the result that the spin was found to be in the state
867: $\psbu$. The intervention operator is the projection
868: operator on this state of the second subsystem: $T (\rho) =
869: 2\ P_2^{\uparrow} \rho P_2^{\uparrow}$. We obtain:
870: $\rho_{bd} = T (\rho_{bc}) = (\psauk \otimes \ \psbuk)
871: (\psaub \ \otimes
872: \psbub)$. If now we attempt to trace $\rho_{bd}$ over the
873: subsystem associated with the edge $e_d$, we will obtain an
874: incorrect result for $\rho_b$, namely $\psauk \psaub$. The
875: resolution is well known. Since a classical observer
876: located on the edge $e_b$ is not aware of the result of the
877: intervention at the second vertex, for him the density
878: matrix $\rho_{bd}$ has evolved from $\rho_{bc}$ by an
879: operator $\tilde T$ which includes all possible outcomes of
880: the measurement:
881: $\tilde\rho_{bd} = {\tilde T} (\rho_{bc}) =
882: \sum_{s=\uparrow,\downarrow} P_2^s \rho_{bc} P_2^s$.
883: Tracing out the
884: $d$-subsystem in the expression for $\tilde\rho_{bd}$, we
885: obtain the correct expression for $\rho_b$, namely
886: $\rho_b = 1/2 \ (\psauk \psaub + \psadk \psadb)$.
887:
888: Now we give the general prescription for computing the
889: density matrix on an edge $e_i$ from an arbitrary locative
890: slice $L$ containing this edge. We first compute a density
891: matrix $\tilde\rho_L$ for the slice $L$. But note this is
892: not the density matrix of definition~\ref{rho}.
893:
894: This density matrix is computed from the initial data by
895: applying intervention operators for the events in the past
896: of $L$ as before. But now, we will consider two types of
897: events in the past of $L$, those that are in the past of
898: $e_i$ and those that are not. For the events that are in
899: the past of the edge $e_i$, we use our regular intervention
900: operators without a summation over the set of possible
901: outcomes: $\rho \mapsto 1/p_{\mu} \sum_m A_{\mu m} \rho
902: A_{\mu m}^\dagger$. We do not sum over the outcomes in this
903: case precisely because the outcome is in fact known at
904: $e_i$. For the events that are in the past of the slice $L$
905: but not in the past of the edge $e_i$, we use operators
906: which sum over all possible outcomes: $\rho \mapsto
907: \sum_{\mu m} A_{\mu m} \rho A_{\mu m}^\dagger$. This time,
908: of course, the summation is there because the outcome cannot
909: be known at $e_i$ since these events are not in the past of
910: $e_i$.
911:
912: After we have obtained $\tilde\rho_L$, we trace out those
913: subsystems associated with edges in $L$ except for $e_i$ to
914: obtain the density matrix
915: $\tilde\rho_i$. This is the density matrix associated with
916: our preferred edge $e_i$, as computed from the slice $L$.
917: The independence of the result on the choice of $L$ is
918: expressed in the following proposition:
919:
920: \begin{prop}[Covariance] Let $e_i$ be an edge in the dag
921: $G$. The density matrix
922: $\rho_i$ associated with the edge $e_i$ does not depend on
923: the choice of locative slice used to compute it.
924: \end{prop}
925:
926: \begin{proof} \\ We have already demonstrated that to any
927: edge $e_i$, there is a unique least locative slice $M_i$
928: containing $e_i$. Let
929: $\rho_i$ be the density matrix for the edge $e_i$ as
930: computed from the least locative slice and let
931: $\tilde\rho_i$ be the density matrix for the same edge but
932: computed from an arbitrary locative slice, say $L$,
933: containing $e_i$. We will demonstrate the lemma by showing
934: that
935: $\rho_i=\tilde\rho_i$.
936:
937: First note that $M_i$ being less than $L$ implies that there
938: is a set $V$ of events between $M_i$ and $L$. The plan is
939: to remove the effect of these events and show that, at each
940: stage, the density matrix is unaffected. We begin by
941: picking a maximal event, say $k$, with the intervention
942: operator
943: $T_k$. Since $k$ is maximal and hence acausal with all
944: other maximal elements of $V$, as well as with all the
945: maximal elements to the past of
946: $e_i$, the intervention operator at $k$ commutes with all
947: the intervention operators at the vertices just mentioned.
948: Thus, we can choose the intervention operator $T_k$ to be
949: the outermost, i.e.\ the density matrix
950: $\rho_L$ obtained by propagating to $L$ can be written as
951: \[ \rho_L = T_k(\rho') \] where $\rho'$ is the density
952: matrix on the (locative) slice obtained by removing the
953: edges to the future of $k$ from $L$ and adding the edges to
954: the past of $k$. Using the explicit general form for an
955: intervention operator,
956: \[ \rho_L =
957: \sum_{\mu,m}A^{(k)}_{\mu,m}\rho'A^{\dagger(k)}_{\mu,m}.\] In
958: order to obtain the density matrix $\tilde\rho_i$, we trace
959: over all Hilbert spaces associated with edges in $L$ except
960: $e_i$. In particular, we trace over the outgoing edges
961: associated with $k$. Now we can use the cyclic property of
962: trace and rewrite the expression for $\tilde\rho_i$ as,
963: \[ \tilde\rho_i =
964: Tr(\sum_{\mu,m}A^{\dagger(k)}_{\mu,m}A^{(k)}_{\mu,m}\rho').
965: \] Now we use the identity
966: \[ \sum_{\mu m} A_{\mu m}^\dagger A_{\mu m}= I \] to get
967: \[ \tilde\rho_i = Tr(\rho').\]
968:
969: We have eliminated the effect of the intervention operator at
970: $k$. Proceeding inductively we can peel off the intervention
971: operators associated with the rest of the vertices in $V$,
972: thus
973: \[ \tilde\rho_i = \rho_i.\]
974: \end{proof}
975:
976: A similar argument for the case of a simple system
977: represented by the dag in Figure~\ref{figureD} is contained
978: in~\cite{Peres00b}.
979:
980: \section{Polycategories}\label{poly}
981:
982: We now wish to give a more axiomatic treatment of the above
983: construction. This will require the use of several concepts from
984: category theory and logic, which we now present.
985:
986: We begin by introducing the algebraic or categorical concepts
987: necessary for our formulation of the dynamics of quantum
988: information flow. While it might seem that these structures
989: are excessively abstract, this level of abstraction has
990: several advantages. First, it provides a great deal of
991: generality. Our definition can be applied in many contexts,
992: in particular it may be applied in situations other than the
993: sorts of information flow considered here. Second, the two
994: crucial properties of interest, causality and covariance,
995: now become straightforward consequences of the functoriality
996: of our axioms.
997:
998: \subsection{Posets, directed graphs and categories}
999:
1000: For comparison, we recall briefly that a poset is a set $P$
1001: together with a binary relation on $P$ (i.e. a subset of
1002: $P\times P$) denoted $\leq$ that satisfies the properties of
1003: antisymmetry, transitivity and reflexivity. It is a natural
1004: generalization of this idea to consider \emph{directed
1005: graphs}. A directed graph is simply a set $D$, the set of
1006: \emph{vertices} or \emph{nodes}, together with a binary
1007: relation $R$ on $D$. No properties of $R$ are required in
1008: the definition of directed graph. In particular there is no
1009: implicit transitivity assumed. A directed graph has a
1010: natural geometric visualization. One considers the nodes as
1011: points in the plane, and if $x$ and $y$ are nodes with
1012: $\langle x,y\rangle\in R$, we draw an arrow from $x$ to $y$.
1013:
1014: As already remarked, the nodes of our directed graph will be
1015: events, and arrows will represent propagation from one event
1016: to another. To avoid temporal loops, we will add the single
1017: requirement that our directed graphs be \emph{acyclic},
1018: i.e. there does not exist a sequence of edges
1019: $x_1,x_2,\ldots, x_n$ such that for all
1020: $i\in\{1,2,\ldots,n-1\}$, we have
1021: $\langle x_i,x_{i+1}\rangle\in R$, and $x_1=x_n$. This of
1022: course corresponds to there being no directed cycles in the
1023: geometric representation. Hereafter, a directed acyclic
1024: graph will be called a \emph{dag}. Note that every poset,
1025: considered as a directed graph, is acyclic. This is a
1026: consequence of transitivity and antisymmetry. But dags are
1027: a genuine generalization of posets.
1028:
1029: This difference will become more apparent when we consider
1030: the space of
1031: \emph{paths}. In a poset all the paths are already included
1032: (even if they are not explicitly drawn in the visualization
1033: of the poset). When we consider paths through a dag we may
1034: have multiple paths between the same two vertices. These
1035: multiple paths represent different ways that information
1036: flowed from one point to another, thus, we must regard them
1037: as distinct. Therefore - unlike the case with posets - we
1038: do not just want to regard the resulting structure as a
1039: binary relation, rather, we want to view it as a category.
1040:
1041: It is natural to associate to any dag $D$, indeed to any
1042: directed graph, a category. We first briefly remind the
1043: reader of the basic definitions. See \cite{Maclane98} for a
1044: more extensive introduction.
1045:
1046: \begin{defin}
1047: \rm{ A \textbf{category} \textsf{C} consists of two
1048: collections, the collection of {\it objects} and the
1049: collection of {\it morphisms}. Each morphism is assigned a
1050: domain and codomain, both being objects of
1051: \textsf{C}. Typically we write $f\colon A\rarr B$ to mean
1052: $f$ is a morphism with domain $A$ and codomain $B$. To
1053: every object $A$, we have a special morphism, the identity
1054: $id\colon A\rarr A$. There is also a composition law which
1055: takes morphisms $f\colon A\rarr B$ and $g\colon B\rarr C$
1056: and returns a morphism $gf\colon A\rarr C$. All this data
1057: must satisfy several evident equations, as described for
1058: example in
1059: \cite{Maclane98}. We also remind the reader that a
1060: \textbf{functor} is a morphism of categories, i.e. a
1061: functor, denoted $F\colon C\rarr D$ consists of a function
1062: taking objects $c\in C$ to objects $F(c)\in D$ and taking
1063: morphisms $f\colon c\rarr d$ in $C$ to morphisms $F(f)\colon
1064: F(c)\rarr F(d)$. A functor must preserve identities and
1065: composition. } \end{defin}
1066:
1067: To each dag $D$, we associate a category $\mathsf{C}(D)$.
1068: This is the category \emph{freely generated} by the dag.
1069: See for example \cite{Maclane98} Chapter 2, for a detailed
1070: description. The objects of our free category will be the
1071: vertices of $D$. If $x$ and $y$ are vertices, a morphism
1072: from $x$ to $y$ is a directed path in our dag. Identities are
1073: paths of length 0, and composition is given by concatenation
1074: of paths. The verification of the axioms for a category is
1075: straightforward.
1076:
1077: One of the key points of our work is that we are proposing
1078: passing from posets to categories. As we have remarked
1079: before, categories are more general than posets, indeed
1080: posets correspond to a degenerate class of categories in
1081: which there is at most one morphism between any two
1082: objects. The richer structure of categories allows us to
1083: retain more information about the system. Intuitively, the
1084: use of categories allows us not merely to note that $x$
1085: causally precedes
1086: $y$, but to keep track of the different ways that $x$ may
1087: evolve into $y$. To make this more precise, we need
1088: a slightly different construction on dags, which will
1089: yield polycategories as opposed to categories.
1090:
1091: \subsection{Polycategories}
1092:
1093: Roughly speaking, the distinction between categories and
1094: polycategories is the following: A category allows one to
1095: have morphisms which go from single objects to single
1096: objects. A polycategory allows one to have morphisms from
1097: lists of objects to lists of objects. A typical morphism in
1098: a polycategory (hereafter called a polymorphism) would be
1099: denoted:
1100:
1101: \[ f\colon A_1,A_2,\ldots,A_n\lrarr B_1,B_2,\ldots,B_m\]
1102:
1103: There are a number of contexts in which such a generalization would be
1104: useful. Before giving the formal definition, we discuss two such contexts.
1105: The first arises in algebra. Consider Hilbert spaces, vector spaces or any
1106: class of modules in which one can form a tensor product. Then we can
1107: define a polycategory as follows. Our objects will be such spaces, and a
1108: morphism of the above form will be a linear function:
1109:
1110: \[ f\colon A_1\ox A_2\ox\ldots\ox A_n\lrarr B_1\ox
1111: B_2\ox\ldots\ox B_m\]
1112:
1113: Thus polycategories have proven to be quite useful in the analysis of
1114: (ordinary) categories in which one can form tensor products of objects.
1115: Indeed this was the original motivation for their definition. See
1116: \cite{Lambek69,Szabo75}. Categories in which one has a reasonable notion
1117: of tensor product are called \emph{monoidal}, and have recently figured
1118: prominently in several areas of mathematical physics, most notably
1119: topological quantum field theory \cite{Atiyah90,Baez95}.
1120:
1121: The second well-known application of polycategories is to
1122: logic. Typically logicians are interested in the analysis of
1123: \emph{sequents}, written:
1124: \[ A_1,A_2,\ldots,A_n\vdash B_1,B_2,\ldots,B_m\]
1125: \noindent Now
1126: $A_1,A_2,\ldots,A_n,B_1,B_2,\ldots,B_m$ represent formulas
1127: in some logical system. We say that the above sequent holds
1128: if and only if the conjunction of $A_1,A_2,\ldots,A_n$
1129: logically entails the disjunction of $B_1,B_2,\ldots,B_m$.
1130: There is a well-established correspondence between the sort
1131: of logical entailments considered here and categorical
1132: structures. See for example
1133: \cite{Lambek86}.
1134:
1135: But notice the difference between this and our first
1136: example. When talking about vector spaces, the ``commas''
1137: on the left and right were both interpreted as the tensor
1138: product. However in the logic example, we have two
1139: different interpretations. Commas on the left are treated
1140: as conjunction, while commas on the right are treated as
1141: disjunction. Thus for a proper categorical interpretation
1142: of polycategories, one needs categories with two monoidal
1143: structures which interact in an appropriate fashion. Such
1144: categories are called \emph{linearly} or \emph{weakly
1145: distributive}, a notion due to Cockett and Seely
1146: \cite{Cockett97,Blute96}. Linearly distributive categories
1147: are the appropriate framework for considering a specific
1148: logical system known as \emph{linear logic}, introduced by
1149: Girard \cite{Girard87,Girard89}. For a brief exposition of
1150: linear logic, see the appendix. As we will see, the refined
1151: logical connectives of linear logic will be used to express
1152: the entanglements of our system.
1153:
1154: There is a very geometric or graphical calculus for
1155: representing morphisms in polycategories, which was
1156: introduced by Joyal and Street in
1157: \cite{Joyal91}, and given a logical interpretation in
1158: \cite{Blute96}. A polymorphism of the form:
1159: \[ f\colon A_1,A_2,\ldots,A_n\lrarr B_1,B_2,\ldots,B_m\]
1160: \noindent is represented as follows:
1161:
1162: \setlength{\unitlength}{.6in}
1163:
1164: \begin{picture}(6,3)(-2,0)
1165: \put(1.2,1){\framebox(3,.6){$f$}}
1166: \put(1.5,1){\vector(0,-1){.6}}
1167: \put(1.9,1){\vector(0,-1){.6}}
1168: \put(2.7,1){\vector(0,-1){.6}}
1169: \put(3,.7){\ldots}
1170: \put(2.2,.7){\ldots}
1171: \put(3.5,1){\vector(0,-1){.6}}
1172: \put(3.9,1){\vector(0,-1){.6}}
1173: \put(1.5,2.2){\vector(0,-1){.6}}
1174: \put(1.9,2.2){\vector(0,-1){.6}}
1175: \put(2.7,2.2){\vector(0,-1){.6}}
1176: \put(3,1.9){\ldots}
1177: \put(2.2,1.9){\ldots}
1178: \put(3.5,2.2){\vector(0,-1){.6}}
1179: \put(3.9,2.2){\vector(0,-1){.6}}
1180: \put(1.15,2.1){$A_1$}
1181: \put(1.55,2.1){$A_2$}
1182: \put(2.9,2.1){$A_{n-1}$}
1183: \put(4.0,2.1){$A_n$}
1184: \put(1.15,.5){$B_1$}
1185: \put(1.55,.5){$B_2$}
1186: \put(2.85,.5){$B_{m-1}$}
1187: \put(4.0,.5){$B_m$}
1188: \end{picture}
1189:
1190: Thus the polymorphism is represented as a box, with the
1191: incoming and outgoing arrows labelled by objects.
1192: Composition in polycategories then can be represented
1193: pictorially in a very natural fashion. Before giving a
1194: general discussion of composition in a polycategory, we
1195: illustrate this graphical representation. Suppose we are
1196: given two polymorphisms of the following form:
1197:
1198: \begin{center}
1199: $f\colon A_1,A_2,\ldots,A_n\lrarr B_1,B_2,\ldots,B_m,C$\\
1200: $g\colon C,D_1,D_2,\ldots,D_k\lrarr E_1,E_2,\ldots,E_j$
1201: \end{center}
1202:
1203: Note the single object $C$ common to the codomain of $f$ and
1204: the domain of $g$. Then under the definition of
1205: polycategory, we can compose these to get a morphism of
1206: form:
1207:
1208: \[g\circ_Cf\colon A_1,A_2,\ldots,A_n,D_1,D_2,\ldots,D_k\lrarr B_1,B_2,
1209: \ldots,B_m,E_1,E_2,\ldots,E_j\]
1210:
1211: The object $C$ which ``disappears'' after composition is called the \emph{cut
1212: object}, a terminology derived from logic. Note that we subscript the
1213: composition by the object being cut. This composition would be represented
1214: by the diagram on Figure~\ref{FigureComp}:
1215:
1216: \begin{figure}[htb]
1217: \begin{picture}(6,4)(-1,-1.5)
1218: \put(1.2,1){\framebox(3,.6){$f$}}
1219: \put(1.5,1){\vector(0,-1){2.12}}
1220: \put(1.9,1){\vector(0,-1){2.12}}
1221: \put(2.7,1){\vector(0,-1){2.12}}
1222: \put(3,.7){\ldots}
1223: \put(2.2,.7){\ldots}
1224: \put(3.5,1){\vector(0,-1){2.12}}
1225: \put(3.9,1){\line(0,-1){.6}}
1226: \put(1.5,2.2){\vector(0,-1){.6}}
1227: \put(1.9,2.2){\vector(0,-1){.6}}
1228: \put(2.7,2.2){\vector(0,-1){.6}}
1229: \put(3,1.9){\ldots}
1230: \put(2.2,1.9){\ldots}
1231: \put(3.5,2.2){\vector(0,-1){.6}}
1232: \put(3.9,2.2){\vector(0,-1){.6}}
1233:
1234: \put(3.6,-0.5){\framebox(3,.6){$g$}}
1235: \put(3.9,-.5){\vector(0,-1){.6}}
1236: \put(4.3,-.5){\vector(0,-1){.6}}
1237: \put(5.1,-.5){\vector(0,-1){.6}}
1238: \put(5.4,-.8){\ldots}
1239: \put(4.6,-.8){\ldots}
1240: \put(5.9,-.5){\vector(0,-1){.6}}
1241: \put(6.3,-.5){\vector(0,-1){.6}}
1242: \put(3.9,.7){\line(0,-1){.6}}
1243: \put(3.65,.5){$C$}
1244: \put(4.3,2.23){\vector(0,-1){2.13}}
1245: \put(5.1,2.23){\vector(0,-1){2.13}}
1246: \put(5.4,.4){\ldots}
1247: \put(4.6,.4){\ldots}
1248: \put(5.9,2.23){\vector(0,-1){2.13}}
1249: \put(6.3,2.23){\vector(0,-1){2.13}}
1250: \end{picture}
1251: \caption{}
1252: \label{FigureComp}
1253: \end{figure}
1254:
1255: We only label the segment corresponding to the cut object, for ease of
1256: reading. Thus composition in a polycategory is represented by the
1257: concatenation of the graphs of $f$ and $g$, followed by joining the
1258: incoming and outgoing edges corresponding to the cut object. There are
1259: several other possibilities for applications of the composition rule. In
1260: some cases, the graphical representation requires our arrows to cross.
1261: This corresponds to having a \emph{symmetric} polycategory. This is very
1262: much related to having a symmetric tensor or tensors, i.e. ones with the
1263: property that $A\ox B\cong B\ox A$. We will always assume our
1264: polycategories are symmetric.
1265:
1266: We now give a more formal definition of polycategory. We
1267: refer the reader to \cite{Cockett97,Szabo75} for further details.
1268:
1269: \begin{defin}
1270: \rm{
1271: A \textbf{polycategory} $\mathsf{C}$ consists of
1272: the following data:
1273:
1274: \begin{itemize}
1275: \item A set of objects, denoted $|\mathsf{C}|$.
1276: \item If $A_1,A_2,\ldots,A_n$ and $B_1,B_2,\ldots,B_m$ are
1277: finite sequences of objects, then we have a set of
1278: morphisms of the form
1279: $f\colon A_1,A_2,\ldots,A_n\lrarr B_1,B_2,\ldots,B_m$. We note that
1280: technically one must consider these sequences of objects as being defined
1281: only up to permutation.
1282: \item For every object $A$, we have an identity
1283: morphism $id_A\colon A\rarr A$.
1284: \end{itemize}
1285:
1286: \noindent The composition law was already described pictorially. The data
1287: of course are subject to a number of axioms, of which most important for us
1288: is the one which requires associativity of composition. The
1289: notion of \emph{polyfunctor} between polycategories is also straightforward
1290: to formulate. One first has a function $F$ taking objects to objects, and
1291: then given a morphism $f\colon A_1,A_2,\ldots,A_n \lrarr B_1,B_2, \ldots,
1292: B_m$, one assigns to it a morphism
1293: \begin{equation}
1294: F(f)\colon F(A_1),F(A_2),\ldots, F(A_n) \lrarr F(B_1),F(B_2),\ldots,F(B_m).
1295: \end{equation}
1296: Again, a number of axioms must be satisfied, in particular the polyfunctor
1297: must commute with the composition of polymorphisms.
1298: }
1299: \end{defin}
1300:
1301: As suggested by the above, there is a relationship between polycategories
1302: and monoidal categories. It is summarized in the following lemma, which
1303: can be found for example in \cite{Cockett97}:
1304:
1305: \begin{prop}\label{thelemma}
1306: Let $\mathsf{C}$ be a monoidal category. Then one can associate to
1307: $\mathsf{C}$ a polycategory (which will typically be denoted by
1308: $P(\mathsf{C})$ as follows:
1309: \begin{itemize}
1310: \item The objects of $P(\mathsf{C})$ will be the same as those of
1311: $\mathsf{C}$.
1312: \item A polymorphism of the form $f\colon A_1,A_2,\ldots,A_n\lrarr B_1,B_2,
1313: \ldots,B_m$ is a morphism $f\colon A_1\ox A_2\ox\ldots\ox A_n\lrarr B_1\ox
1314: B_2\ox\ldots\ox B_m$.
1315: \item Composition is induced by the composition in $\mathsf{C}$ in the
1316: following way. Suppose that we have two polymorphisms in $P(\mathsf{C})$
1317: as follows:
1318:
1319: \begin{center}
1320: $f\colon A_1,A_2,\ldots,A_n\lrarr B_1,B_2,\ldots,B_m,C$\\
1321: $g\colon C,D_1,D_2,\ldots,D_k\lrarr E_1,E_2,\ldots,E_j$
1322: \end{center}
1323:
1324: \noindent Then since we are in a monoidal category, we have morphisms
1325: \begin{center}
1326: $f\colon A_1\ox A_2\ox \ldots\ox A_n\lrarr B_1\ox B_2\ox\ldots\ox B_m\ox C$\\
1327: $g\colon C\ox D_1\ox D_2\ox \ldots\ox D_k\lrarr E_1\ox E_2\ox \ldots\ox E_j$
1328: \end{center}
1329:
1330: The composite in $P(\mathsf{C})$ is then given by:
1331: \begin{equation}\label{identi}
1332: g\circ_C f=(id_{B_1\ox B_2\ox\ldots\ox B_m}\ox g)\circ(f\ox id_{D_1\ox
1333: D_2\ldots\ox D_k})
1334: \end{equation}
1335: \end{itemize}
1336: \end{prop}
1337:
1338: We note that the concepts of polycategory and monoidal category are not
1339: equivalent. To obtain an equivalence, one needs to replace monoidal
1340: categories with the more general notion of linearly distributive category
1341: mentioned above.
1342:
1343: Now we will demonstrate that a dag generates a polycategory. In this
1344: construction, the nodes of the dag will be assigned morphisms and the edges
1345: will be assigned objects.
1346:
1347: We consider the dag example of Figure~\ref{figure6}. We have changed
1348: labels to be more appropriate for the present
1349: discussion.
1350:
1351:
1352: \begin{figure}[htb]
1353: \input{figure6}
1354: \caption{}
1355: \label{figure6}
1356: \end{figure}
1357:
1358: The idea behind the construction is that the nodes of the
1359: dag (the boxes in our picture) will correspond to
1360: polymorphisms. For example, in the above picture, the box
1361: $f_1$ determines a polymorphism:
1362:
1363: \[f_1\colon A\lrarr C,D\]
1364:
1365: Similarly, $f_4$ determines a polymorphism $f_4\colon
1366: D,E\rarr G$. Thus we see that one has a polymorphism
1367: corresponding to each node. The domain of that polymorphism
1368: will be the labels of the incoming arrows, and the codomain
1369: is determined by the labels of the outgoing arrows. These
1370: are the basic morphisms of the polycategory. As in the
1371: previous construction, one must adjoin morphisms
1372: corresponding to the allowable compositions. For example,
1373: in the above case, we can compose the morphisms $f_4$ and
1374: $f_1$ along the cut object $D$ to obtain a new polymorphism
1375: $f_4\circ_D f_1\colon A,E \rarr C,G$. One must also add
1376: identities and must force these composites to satisfy the
1377: appropriate equations. This construction yields the {\it
1378: polycategory freely generated by the dag}. More generally,
1379: we would have the following definition.
1380:
1381: \begin{defin}\label{free}{\rm We suppose that we are given a
1382: finite dag $G$. The {\em free polycategory generated by
1383: $G$}, denoted $P(G)$, is defined as follows. If a given
1384: vertex $v$ has incoming edges
1385: $A_1,A_2,\ldots, A_n$ and outgoing edges $B_1,B_2,\ldots,B_m$
1386: then the polycategory will have a polymorphism of the form
1387: $f_v\colon A_1,A_2,\ldots,A_n\rarr B_1,B_2,\ldots,B_m$. In
1388: general by induction, if $P(G)$ has polymorphisms of the
1389: form:
1390:
1391: \begin{center}
1392: $f\colon A_1,A_2,\ldots,A_n\lrarr B_1,B_2,\ldots,B_m,C$\\
1393: $g\colon C,D_1,D_2,\ldots,D_k\lrarr E_1,E_2,\ldots,E_j$
1394: \end{center}
1395:
1396: \noindent then we require the existence of a composite
1397: $g\circ_C f$ as a new polymorphism. We assume the existence
1398: of an identity morphism for each edge of $G$. Finally we
1399: impose on this data the necessary equations implied by the
1400: definition of polycategory. }
1401: \end{defin}
1402:
1403: \subsection{Categories of interventions} Next we describe an
1404: appropriate for our construction polycategory of intervention
1405: operators; there are several reasonable choices, this being
1406: the most straightforward. We start with the well known fact that the
1407: category $\mathsf{Hilb}$ of Hilbert spaces and bounded
1408: linear operators is a monoidal category. Hence by the
1409: construction of lemma \ref{thelemma}, we obtain a
1410: polycategory. However this is not the category we will
1411: ultimately use. We will introduce a category
1412: $\mathsf{Conj}$. Intuitively, the objects are Hilbert space
1413: endomorphisms and morphisms are conjugations. A more formal
1414: definition is as follows. Objects are finite-dimensional
1415: Hilbert spaces. A morphism from
1416: $\hi_1$ to $\hi_2$ is a finite family of maps $\{A_i\}_{i\in
1417: I}$ of linear morphisms $A_i\colon \hi_1\rarr\hi_2$.
1418: Composition is then described as follows. If we have the
1419: following pair of maps:
1420:
1421: \[ \hi_1\to^{\{A_i\}_{i\in I}}\hi_2
1422: \to^{\{B_j\}_{j\in J}}\hi_3\]
1423:
1424: \noindent then the composite is:
1425:
1426: \[ \hi_1\to^{\{B_j\circ A_i\}_{\langle i,j\rangle\in I\times J}}\hi_3\]
1427:
1428: A morphism in $\mathsf{Conj}$ can be seen as taking
1429: endomorphisms of
1430: $\hi_1$ to endomorphisms of $\hi_2$ by the formula
1431: $\cO\mapsto\sum_m A_m {\cal O} A_m^{\dagger}$. The monoidal
1432: structure on $\mathsf{Hilb}$ lifts to a monoidal structure
1433: on the category $\mathsf{Conj}$. The tensor product
1434: operator is the usual tensor product of operators on Hilbert
1435: spaces, on maps we take all possible pairings. We next
1436: restrict the class of morphisms by considering only those
1437: families suxh that the corresponding conjugation is trace
1438: preserving. We call
1439: the resulting category $\mathsf{Dio}$. This category also
1440: inherits a monoidal structure. As discussed in Lemma
1441: \ref{thelemma} any monoidal category canonically gives rise
1442: to a polycategory associated to it. We will denote by
1443: $\cP(\mathsf{Dio})$ the polycategory associated with
1444: $\mathsf{Dio}$.
1445:
1446: \section{The logic of polycategories}\label{logic}
1447:
1448: While definition~\ref{free} gives the free polycategory
1449: generated by a dag $G$, it will prove to be useful to have
1450: a more constructive description. Proof-theoretic
1451: techniques have proven to be useful in describing free polycategories.
1452: In our case, the logical structures necessary are quite
1453: simple, and so we digress briefly to put definition~\ref{free} in
1454: logical terms. Recall
1455: that one of the common interpretations of a polymorphism is
1456: as a logical sequent\footnote{We note that for
1457: purposes of this paper sequents should always be considered
1458: "up to permutation", i.e. one may rearrange the order of
1459: premises and conclusions as one sees fit.} of the form:
1460:
1461: \[ A_1,A_2,\ldots,A_n\vdash B_1,B_2,\ldots,B_m\]
1462:
1463: Our system will have only one inference rule, called
1464: the \textit{Cut rule}, which states:
1465: \begin{center}
1466: \mbox{
1467: \infer{\Gamma,\Gamma'\vdash\Delta,\Delta'}{\Gamma\vdash\Delta, A &
1468: \Gamma',A\vdash\Delta'}}
1469: \end{center}
1470:
1471: This should be interpreted as saying that if one has derived
1472: the two sequents above the line, then one can infer the
1473: sequent below the line. Proofs in the system always begin
1474: with \textit{axioms}. Axioms are of the form
1475: $A_1,A_2,\ldots,A_n\vdash B_1,B_2,\ldots,B_m$, where
1476: $A_1,A_2,\ldots, A_n$ are the incoming edges of some vertex
1477: in our dag, and
1478: $B_1,B_2,\ldots,B_m$ will be the outgoing edges. There will
1479: be one such axiom for each vertex in our dag. For example,
1480: consider Figure~\ref{fig3}. Then we will have the following
1481: axioms:
1482: \[
1483: a\stackrel{1}{\ent} c\;\;\;
1484: b\stackrel{2}{\ent} d,e,f\;\;\;
1485: c,d\stackrel{3}{\ent} g,h\;\;\;
1486: e\stackrel{4}{\ent} i\;\;\;
1487: f,g\stackrel{5}{\ent} j\;\;\;
1488: h,i\stackrel{6}{\ent} k
1489: \] where we have labelled each entailment symbol with the
1490: name of the corresponding vertex. The following is an
1491: example of a deduction in this system of the sequent
1492: $a,b\vdash f,g,h,i$.
1493:
1494: \begin{center}
1495: \mbox{
1496: \infer{a,b\vdash f,g,h,i}
1497: {\infer{a,b\vdash e,f,g,h}{b\vdash
1498: d,e,f & \infer{a,d\vdash g,h}{a\vdash c & c,d\vdash g,h}}
1499: & e\vdash i}
1500: }
1501: \end{center}
1502: This deduction corresponds to the fact that in the free
1503: polycategory generated by this dag, one has a morphism
1504: $a,b\rarr f,g,h,i$. In fact, it is easy to see that there is
1505: a precise correspondence between deductions in this logical
1506: system and nonidentity morphisms in the free polycategory.
1507:
1508: As a first attempt at capturing quantum evolution on a dag $G$
1509: axiomatically, one
1510: might consider taking a polyfunctor from $P(G)$ to
1511: $P(\mathsf{Hilb})$, where
1512: $\mathsf{Hilb}$ is the usual category of finite-dimensional
1513: Hilbert spaces with its usual tensor product. Note that such
1514: a polyfunctor must necessarily take a sequence of, say, incoming edges
1515: $A_1,A_2,\ldots,A_n$ to
1516: $\hi_1\ox\hi_2\ldots\ox\hi_n$ where $\hi_i$ corresponds to
1517: $A_i$. Then one would (tentatively) define a set
1518: $\Delta$ of edges to be {\it valid} if there is a
1519: deduction in the logic generated by $G$ of
1520: $\Gamma\vdash\Delta$ where
1521: $\Gamma$ is a set of initial edges. Equivalently there must
1522: be a morphism
1523: $\Gamma\rarr\Delta$ in $P(G)$. Then the polyfunctor would
1524: take this to a morphism of Hilbert spaces
1525: $T\colon\hi_\Gamma\rarr\hi_\Delta$. The initial density
1526: matrices would always be assumed to be given, and one would
1527: just apply $T$ to the appropriate initial density matrices
1528: to obtain the density matrix associated to $\Delta$. The locative
1529: slices are the ones on which density matrices can be obtained without
1530: the trace operation and we are looking to equate the notions of
1531: locative and valid for slices. This
1532: approach would be genuinely axiomatic, and would evidently
1533: be applicable to other situations by simply using a category
1534: other than Hilbert spaces as the target of the polyfunctor.
1535: Furthermore we would suggest that using logic as the means
1536: of calculating the matrices gives the approach a very
1537: canonical flavor.
1538:
1539: However, with this notion of validity, we would fail to
1540: capture all locative slices, and thus our tentative notion
1541: of validity will have to be modified. For example, consider
1542: the dag underlying the system of Figure~\ref{figureD} shown
1543: in Figure~\ref{figure5}.
1544:
1545: \begin{figure}[htb]
1546: \begin{center}
1547: \input{figure5}
1548: \end{center}
1549: \caption{}
1550: \label{figure5}
1551: \end{figure}
1552:
1553: Corresponding to this dag, we get the following basic
1554: morphisms (axioms):
1555: \[a\vdash b,c \,\,\,\,\,\,\, b\vdash d\,\,\,\,\,\,\, c\vdash
1556: e\,\,\,\,\,\,\, d,e\vdash f.\] Evidently, the set $\{f\}$ is
1557: a locative slice, and yet the sequent $a\vdash f$ is not
1558: derivable. The sequent $a\vdash d,e$ is derivable, and one
1559: would like to cut it against $d,e\vdash f$, but one is only
1560: allowed to cut a single formula. Such ``multicuts'' are
1561: expressly forbidden, as they lead to undesirable logical
1562: properties
1563: \cite{Blute93}.
1564:
1565: Physically, the reason for this problem is that the sequent
1566: $d,e\vdash f$ does not encode the information that the two
1567: states at $d$ and $e$ are correlated. It is precisely the
1568: fact that they are correlated that implies that one would
1569: need to use a multicut. To avoid this problem, one must
1570: introduce some notation, specifically a syntax
1571: for specifying such correlations. We will
1572: use the {\it logical connectives} of the multiplicative
1573: fragment of {\it linear logic} \cite{Girard87,Girard95}
1574: to this end. The multiplicative disjunction of linear
1575: logic, denoted $\girpar$ and called the {\it par} connective,
1576: will express such nonlocal correlations.
1577: In our example, we will write the sequent corresponding to
1578: vertex $4$ as
1579: $d\girpar e\vdash f$ to express the fact that the subsystems associated
1580: with these two edges are possibly entangled through interactions in their
1581: common past.
1582:
1583: Note that whenever two (or more)
1584: subsystems emerge from an interaction, they are correlated.
1585: In linear logic, this is reflected by the following rule
1586: called the (right) \emph{Par rule}:
1587:
1588: \begin{center}
1589: \mbox{
1590: \infer[]{\Gamma\vdash \Delta,A\girpar
1591: B}{\Gamma\vdash \Delta,A, B} }
1592: \end{center}
1593: Thus we can always introduce the symbol for correlation in the right
1594: hand side of the sequent.
1595:
1596: Notice that we can cut along a compound formula without
1597: violating any logical rules. So in the present
1598: setting, we would have the following deduction:
1599: \begin{center}
1600: \mbox{
1601: \infer{a\vdash f}
1602: {\infer{a\vdash d\girpar e}{\infer{a\vdash d, e}
1603: {\infer{a\vdash c,d}{a\vdash b,c & b\vdash d} & c\vdash e}}
1604: & d\girpar e\vdash f} }
1605: \end{center} All the cuts in this deduction are legitimate;
1606: instead of a multicut we are cutting along a compound
1607: formula in the last step. So the first step in modifying our
1608: general prescription is to extend our polycategory logic, which originally
1609: contained only the cut rule, to
1610: include the connective rules of linear logic. These are described in
1611: the appendix.
1612:
1613: The above logical rule determines how one introduces a par connective
1614: on the righthand side of a sequent. For the lefthand side,
1615: one introduces pars in the axioms by the following general prescription.
1616: Given a vertex in a
1617: multigraph, we suppose that it has incoming edges
1618: $a_1,a_2,\ldots,a_n$ and outgoing edges $b_1,b_2,\ldots,b_m$.
1619: In the previous formulation, this vertex would have been
1620: labelled with the axiom $\Gamma=a_1,a_2,\ldots,a_n\vdash
1621: b_1,b_2,\ldots,b_m$. We will now introduce several pars
1622: ($\girpar$) on the lefthand side to indicate entanglements
1623: of the sort described above. Begin by defining a relation
1624: $\sim$ by saying $a_i\sim a_j$ if there is an initial edge
1625: $c$ and directed paths from $c$ to $a_i$ and from $c$ to
1626: $a_j$. This is not an equivalence relation, but one takes
1627: the equivalence relation generated by the relation $\sim$.
1628: Call this new relation $\cong$. This equivalence relation,
1629: like all equivalence relations, partitions the set $\Gamma$
1630: into a set of equivalence classes. One then "pars" together
1631: the elements of each equivalence class, and this determines
1632: the structure of the lefthand side of our axiom. For
1633: example, consider vertices 5 and 6 in Figure~\ref{fig3}.
1634: Vertex 5 would be labelled by $f\girpar g\vdash j$ and
1635: vertex 6 would be labelled by $h\girpar i\vdash k$. On the
1636: other hand, vertex 3 would be labelled by $c,d\vdash g,h$.
1637:
1638: Just as the par connective indicates the existence of past
1639: correlations, we use the more familiar tensor symbol $\ox$,
1640: which is also a connective of linear logic, to indicate the lack of
1641: nonlocal correlation. This connective also has a logical rule:
1642:
1643:
1644: \begin{center}
1645: \mbox{
1646: \infer{\Gamma,\Gamma'\vdash \Delta,\Delta',A\ox B}
1647: {\Gamma\vdash \Delta,A & \Gamma'\vdash \Delta',B} }
1648: \end{center}
1649:
1650: But we note that unlike in ordinary logic, this rule can only
1651: be applied in situations that are physically meaningful. We will say
1652: that two deductions $\pi$ and $\pi'$ are {\it spacelike
1653: separated} if all the the vertices of $\pi$ and $\pi'$
1654: are pairwise spacelike separated. In the above formula, we
1655: require that the deductions of $\Gamma\vdash \Delta,A$ and
1656: $\Gamma'\vdash \Delta',B$ are spacelike separated.
1657: This restriction of application of
1658: inference rules is similar to the restrictions of
1659: {\it ludics} \cite{Girard01}.
1660: >From a categorical standpoint, the restrictions imply
1661: that the connectives are only partial functors, but this is
1662: only a minor issue.
1663:
1664: Summarizing, to every dag $G$ we associate its ``logic'',
1665: namely the edges are considered as formulas and vertices are
1666: axioms. We have the usual linear logical connective rules,
1667: including the cut rule which in our setting is interpreted physically
1668: as propagation. The par connective denotes correlation, and the tensor
1669: lack of correlation. Note that every deduction in our system
1670: will conclude with a sequent of the form $\Gamma\vdash\Delta$,
1671: where $\Gamma$ is a set of initial edges.
1672:
1673: Now one would
1674: like to modify the definition of validity to say that a set
1675: of edges $\Delta$ is {\it valid} if in our extended
1676: polycategory logic, one can derive a sequent
1677: $\Gamma\vdash\hat{\Delta}$ such that the list of edges
1678: appearing in $ \hat{\Delta}$ was precisely $\Delta$, and
1679: $\Gamma$ is a set of initial edges. However this is still
1680: not sufficient as an axiomatic approach to capturing all
1681: locative slices. We note the example in Figure~\ref{figure8}.
1682:
1683: \begin{figure}[htb]
1684: \begin{center}
1685: \input{figure8}
1686: \end{center}
1687: \caption{}
1688: \label{figure8}
1689: \end{figure}
1690:
1691: Evidently the slice $\{f,g\}$ is locative, but we claim that
1692: it cannot be derived even in our extended logic. To this
1693: directed graph, we would associate the following axioms:
1694:
1695: \[ a\vdash c,h \,\,\,\,\,\, b\vdash d,e\,\,\,\,\,\,
1696: c,d\vdash f\,\,\,\,\,\, h,e\vdash g\]
1697:
1698: Note that there are no correlations between $c$ and $d$ or
1699: between $h$ and $e$. Thus no $\girpar$-combinations can be introduced.
1700: Now if one attempts to derive $a,b\vdash f,g$, we proceed as
1701: follows:
1702:
1703: \begin{center}
1704: \mbox{
1705: \infer{a,b\vdash h,e,f}{\infer{a,b\vdash c\ox d,h,e}{a\vdash
1706: c,h & b\vdash d,e} & \infer{c\ox d\vdash f}{c,d\vdash f}}
1707: }\end{center}
1708:
1709: At this point, we are unable to proceed. Had we attempted
1710: the symmetric approach tensoring $h$ and $e$ together, we
1711: would have encountered the same problem.
1712:
1713: The problem is that our logical system is still missing one
1714: crucial aspect, and that is that correlations develop
1715: dynamically as the system evolves, or equivalently as the
1716: deduction proceeds. Thus our axioms must
1717: change dynamically as well. We give the following definition.
1718:
1719: \begin{defin}{\em
1720: Suppose we have a deduction $\pi$ of the sequent
1721: $\Gamma\vdash\Delta$ in the graph logic associated to the
1722: dag $G$, and that $T$ is a vertex in $G$ to the future or acausal
1723: to the edges of the set
1724: $\Delta$ with $a$ and $b$ among the incoming edges of $T$.
1725: Then $a$ and $b$ are {\em correlated} with respect to $\pi$
1726: if there exist outgoing edges $c$ and $d$ of the proof
1727: $\pi$ and directed paths from $c$ to $a$ and from $d$ to $b$.
1728: }\end{defin}
1729:
1730: So the point here is that when performing a deduction, one does
1731: not assign an axiom to a given vertex until it is necessary to use
1732: that axiom in the proof. Then one assigns that axiom using this new
1733: notion of correlation and the equivalence relation defined above.
1734: This prescription reflects the physical reality that entanglement of
1735: local quantum subsystems could develop as a result of a distant interaction
1736: between some other subsystems of the same quantum system.
1737: We are finally able to give the following crucial definition:
1738:
1739:
1740:
1741: \begin{defin}{\em
1742: A set
1743: $\Delta$ of edges in a dag $G$ is said to be {\em valid} if there is a
1744: deduction in the logic associated to $G$ of
1745: $\Gamma\vdash\hat{\Delta}$ where $\hat{\Delta}$ is a sequence of formulas
1746: whose underlying set of edges is precisely $\Delta$ and where
1747: $\Gamma$ is a set of initial edges, in fact the set of initial edges
1748: to the past of $\Delta$.}
1749: \end{defin}
1750:
1751: We are also ready to state the result relating the logical deduction and
1752: the dynamics of Section~\ref{dyna} in a graph.
1753:
1754: \begin{thm} A set of edges is valid if and only if it is locative. More
1755: specifically, if there is a deduction of $\Gamma\vdash\hat{\Delta}$
1756: as described above, then $\Delta$ is necessarily locative. Conversely,
1757: given any locative slice, one can find such a deduction.
1758: \end{thm}
1759: \begin{proof}\\ Recall that a locative slice $L$ is obtained from the set
1760: of initial edges in its past by an inductive procedure.
1761: At each step, we choose arbitrarily a minimal vertex $u$ in the past of $L$,
1762: remove the incoming edges of $u$ and add the outgoing edges. This step
1763: corresponds to the application of a cut rule, and the method we have used
1764: of assigning the par connective
1765: to the lefthand side of an axiom ensures that it is
1766: always a legal cut. The tensor rule is necessary in order to combine
1767: spacelike separated subsystems in order to prepare for the application of
1768: the cut rule.
1769: \end{proof}
1770:
1771: Thus we have successfully given an axiomatic logic-based approach
1772: to describing evolution. In summary, to find the density matrix
1773: associated to a locative slice $\Delta$, one finds a set
1774: of linear logic formulas whose underlying set of atoms is $\Delta$
1775: and a deduction of $\Gamma\vdash\hat{\Delta}$ where $\Gamma$ is
1776: as above. This deduction is interpreted as a morphism in the
1777: corresponding polycategory, and the polyfunctor to $\cP(\mathsf{Dio})$
1778: is applied to obtain a morphism in the category $\mathsf{Dio}$.
1779: (Note that in this context a polyfunctor is furthermore required to take
1780: any tensor or par connective in $\Gamma$ or $\hat{\Delta}$ to the
1781: usual tensor in $\mathsf{Dio}$.)
1782: One then plugs in the given initial data to obtain the density matrix
1783: corresponding to that slice. Given a nonlocative slice, one simply finds
1784: a locative slice containing it, repeats the above procedure and then
1785: traces out the extraneous edges.
1786:
1787: \section{Conclusions}\label{conc}
1788:
1789: We have presented an axiomatic system for the analysis
1790: of quantum evolution. The dynamics is local as to preserve causality, but at
1791: the same time entanglement of separated quantum systems is faithfully
1792: represented. One could apply these ideas
1793: to other situations
1794: by using a category other than the category of intervention
1795: operators
1796: as the target of the functor. An appropriate
1797: categorical structure for the target is the notion of
1798: a {\it traced monoidal category} \cite{Joyal96} or
1799: the notion of a {\it traced ideal} \cite{Blute99}.
1800: See also \cite{Blute00}.
1801: One particular situation which might be analyzed in this
1802: framework is the notion of classical probabilistic
1803: information. The paper \cite{Blute99} contains a category of
1804: {\it probabilistic relations} which might be of particular
1805: interest in this setting.
1806:
1807: Our work also suggests a natural
1808: extension of the notion of {\it consistent} or {\it
1809: decoherent histories} \cite{Gell-Mann93,Griffiths96}.
1810: Restricting the intervention operators at the vertices of our graph $G$
1811: to be projection operators we can consider $G$ to denote a particular
1812: history within a set of histories. This relaxes the usual linear ordering of
1813: events considered in the literature thus far.
1814: An exposition of histories on graphs is under preparation.
1815:
1816: \section*{Acknowledgements} The authors would like to thank
1817: NSERC for its financial support. We would also like to
1818: thank Phil Scott and Jean-Yves Girard for inviting us to
1819: present this work at the joint SMF-AMS conference in Lyon. The
1820: paper
1821: \cite{Markopoulou00}, which led to our initial consideration
1822: of these ideas, was pointed out to us by Ioannis Raptis. We
1823: would like to especially thank Rafael Sorkin for a lengthy
1824: discussion on causal sets and related topics. Finally, the
1825: second author would like to thank the University of Ottawa
1826: Department of Mathematics and Statistics for its support.
1827: \bibliography{causal}
1828:
1829:
1830: \appendix
1831: \section{Linear logic}
1832:
1833: This section can safely be skipped by logicians.
1834:
1835: Linear logic~\cite{Girard87} is a logic introduced by Girard
1836: in 1987 to allow a finer analysis of how ``resources'' are
1837: consumed in the course of a deduction. As already remarked
1838: in the text, the primary objects of study in logic and
1839: especially proof theory are {\it sequents}, and the
1840: constructors of sequents, the {\it inference rules}. Several
1841: examples have already been given such as the {\it cut rule}:
1842:
1843: \begin{center}
1844: \mbox{
1845: \infer[CUT]{\Gamma,\Gamma'\vdash\Delta,\Delta'}{\Gamma\vdash\Delta,
1846: A &
1847: \Gamma',A\vdash\Delta'}}
1848: \end{center}
1849:
1850: So typically an inference rule is a prescription for
1851: creating a more complex sequent from one or possibly several
1852: simpler ones. Two typical inference rules are the rules of
1853: {\it contraction} and {\it weakening}. These are as follows:
1854:
1855: \begin{center}
1856: \mbox{\infer[CONT]{\Gamma,A\vdash\Delta}
1857: {\Gamma,A,A\vdash\Delta}}
1858: \end{center}
1859:
1860: \begin{center}
1861: \mbox{\infer[WEAK]{\Gamma,A\vdash\Delta}
1862: {\Gamma\vdash\Delta}}
1863: \end{center}
1864:
1865: There are similar rules for the righthand side as well. These
1866: have long been standard in most logics, and indeed have a
1867: strong intuitive meaning. For example, contraction says that
1868: it is unnecessary to make the same assumption twice.
1869: However, in Girard's reexamination of the sequent calculus,
1870: he proposed an interpretation in which the formulas to the
1871: left of a sequent are resources to be consumed in the course
1872: of producing the output, i.e. the conclusions. From this
1873: perspective, the rules of contraction and weakening are
1874: quite dubious. The first step towards defining linear logic
1875: then is to recover these rules from the system. The result is
1876: a remarkably rich structure, the most notable aspect of
1877: which is that the usual connectives of logic, conjunction
1878: and disjunction, each split into two connectives. These
1879: connectives are naturally split into two classes, the {\it
1880: multiplicative} and the {\it additive} connectives. It is
1881: only the multiplicative connectives that will concern us
1882: here. Here are the rules for these connectives:
1883:
1884: \begin{center}
1885: \mbox{
1886: \infer[Right-\girpar]{\Gamma\vdash \Delta,A\girpar
1887: B}{\Gamma\vdash \Delta,A, B} }
1888: \end{center}
1889:
1890: \begin{center}
1891: \mbox{
1892: \infer[Left-\girpar]{\Gamma,\Gamma' A\wp B\vdash \Delta,\Delta'}
1893: {\Gamma,A\vdash \Delta & \Gamma',B\vdash \Delta'} }
1894: \end{center}
1895:
1896: \begin{center}
1897: \mbox{
1898: \infer[Right-\ox]{\Gamma,\Gamma'\vdash \Delta,\Delta',A\ox B}
1899: {\Gamma\vdash \Delta,A & \Gamma'\vdash \Delta',B} }
1900: \end{center}
1901:
1902: \begin{center}
1903: \mbox{
1904: \infer[Left-\ox]{\Gamma,A\ox B\vdash \Delta}{\Gamma,A,B\vdash \Delta} }
1905: \end{center}
1906:
1907:
1908: Categorically, the structure of linear logic has striking
1909: properties as well. As is traditional in categorical logic,
1910: one can form a category whose objects are formulas, and
1911: morphisms are proofs. This construction is described for
1912: example in \cite{Lambek69,Lambek86}. When one applies this
1913: construction to (multiplicative) linear logic ({\bf MLL}),
1914: one obtains a special class of symmetric monoidal closed
1915: categories called {\it $*$-autonomous}. These were defined
1916: by Barr in \cite{Barr79}.
1917:
1918: Subsequently it was demonstrated that the correspondence
1919: between proofs in {\bf MLL} and morphisms in the free
1920: $*$-autonomous category is quite sharp. See
1921: \cite{Blute93,Blute96}. This correspondence between
1922: morphisms and proofs is best expressed using {\it proof
1923: nets}, a graph-theoretic system for representing {\bf MLL}
1924: proofs \cite{Girard87}. Proof nets had already been seen to
1925: be a remarkable deductive system, exhibiting properties of
1926: great importance in the analysis of computation, especially
1927: concurrent computation. The precise connection between
1928: proof nets and free $*$-autonomous categories
1929: provides further evidence of their great utility.
1930:
1931:
1932:
1933:
1934: \end{document}
1935:
1936:
1937:
1938:
1939:
1940:
1941:
1942:
1943:
1944:
1945:
1946:
1947:
1948:
1949:
1950:
1951:
1952:
1953:
1954:
1955: