doc-src/HOL/HOL.tex
author Walther Neuper <neuper@ist.tugraz.at>
Thu, 12 Aug 2010 15:03:34 +0200
branchisac-from-Isabelle2009-2
changeset 37913 20e3616b2d9c
parent 31101 26c7bb764a38
child 43498 8749742785b8
permissions -rw-r--r--
prepare reactivation of isac-update-Isa09-2
     1 %% $Id$
     2 \chapter{Higher-Order Logic}
     3 \index{higher-order logic|(}
     4 \index{HOL system@{\sc hol} system}
     5 
     6 The theory~\thydx{HOL} implements higher-order logic.  It is based on
     7 Gordon's~{\sc hol} system~\cite{mgordon-hol}, which itself is based on
     8 Church's original paper~\cite{church40}.  Andrews's book~\cite{andrews86} is a
     9 full description of the original Church-style higher-order logic.  Experience
    10 with the {\sc hol} system has demonstrated that higher-order logic is widely
    11 applicable in many areas of mathematics and computer science, not just
    12 hardware verification, {\sc hol}'s original \textit{raison d'{\^e}tre\/}.  It
    13 is weaker than ZF set theory but for most applications this does not matter.
    14 If you prefer {\ML} to Lisp, you will probably prefer HOL to~ZF.
    15 
    16 The syntax of HOL\footnote{Earlier versions of Isabelle's HOL used a different
    17   syntax.  Ancient releases of Isabelle included still another version of~HOL,
    18   with explicit type inference rules~\cite{paulson-COLOG}.  This version no
    19   longer exists, but \thydx{ZF} supports a similar style of reasoning.}
    20 follows $\lambda$-calculus and functional programming.  Function application
    21 is curried.  To apply the function~$f$ of type $\tau@1\To\tau@2\To\tau@3$ to
    22 the arguments~$a$ and~$b$ in HOL, you simply write $f\,a\,b$.  There is no
    23 `apply' operator as in \thydx{ZF}.  Note that $f(a,b)$ means ``$f$ applied to
    24 the pair $(a,b)$'' in HOL.  We write ordered pairs as $(a,b)$, not $\langle
    25 a,b\rangle$ as in ZF.
    26 
    27 HOL has a distinct feel, compared with ZF and CTT.  It identifies object-level
    28 types with meta-level types, taking advantage of Isabelle's built-in
    29 type-checker.  It identifies object-level functions with meta-level functions,
    30 so it uses Isabelle's operations for abstraction and application.
    31 
    32 These identifications allow Isabelle to support HOL particularly nicely, but
    33 they also mean that HOL requires more sophistication from the user --- in
    34 particular, an understanding of Isabelle's type system.  Beginners should work
    35 with \texttt{show_types} (or even \texttt{show_sorts}) set to \texttt{true}.
    36 
    37 
    38 \begin{figure}
    39 \begin{constants}
    40   \it name      &\it meta-type  & \it description \\
    41   \cdx{Trueprop}& $bool\To prop$                & coercion to $prop$\\
    42   \cdx{Not}     & $bool\To bool$                & negation ($\lnot$) \\
    43   \cdx{True}    & $bool$                        & tautology ($\top$) \\
    44   \cdx{False}   & $bool$                        & absurdity ($\bot$) \\
    45   \cdx{If}      & $[bool,\alpha,\alpha]\To\alpha$ & conditional \\
    46   \cdx{Let}     & $[\alpha,\alpha\To\beta]\To\beta$ & let binder
    47 \end{constants}
    48 \subcaption{Constants}
    49 
    50 \begin{constants}
    51 \index{"@@{\tt\at} symbol}
    52 \index{*"! symbol}\index{*"? symbol}
    53 \index{*"?"! symbol}\index{*"E"X"! symbol}
    54   \it symbol &\it name     &\it meta-type & \it description \\
    55   \sdx{SOME} or \tt\at & \cdx{Eps}  & $(\alpha\To bool)\To\alpha$ & 
    56         Hilbert description ($\varepsilon$) \\
    57   \sdx{ALL} or {\tt!~} & \cdx{All}  & $(\alpha\To bool)\To bool$ & 
    58         universal quantifier ($\forall$) \\
    59   \sdx{EX} or {\tt?~}  & \cdx{Ex}   & $(\alpha\To bool)\To bool$ & 
    60         existential quantifier ($\exists$) \\
    61   \texttt{EX!} or {\tt?!} & \cdx{Ex1}  & $(\alpha\To bool)\To bool$ & 
    62         unique existence ($\exists!$)\\
    63   \texttt{LEAST}  & \cdx{Least}  & $(\alpha::ord \To bool)\To\alpha$ & 
    64         least element
    65 \end{constants}
    66 \subcaption{Binders} 
    67 
    68 \begin{constants}
    69 \index{*"= symbol}
    70 \index{&@{\tt\&} symbol}
    71 \index{*"| symbol}
    72 \index{*"-"-"> symbol}
    73   \it symbol    & \it meta-type & \it priority & \it description \\ 
    74   \sdx{o}       & $[\beta\To\gamma,\alpha\To\beta]\To (\alpha\To\gamma)$ & 
    75         Left 55 & composition ($\circ$) \\
    76   \tt =         & $[\alpha,\alpha]\To bool$ & Left 50 & equality ($=$) \\
    77   \tt <         & $[\alpha::ord,\alpha]\To bool$ & Left 50 & less than ($<$) \\
    78   \tt <=        & $[\alpha::ord,\alpha]\To bool$ & Left 50 & 
    79                 less than or equals ($\leq$)\\
    80   \tt \&        & $[bool,bool]\To bool$ & Right 35 & conjunction ($\conj$) \\
    81   \tt |         & $[bool,bool]\To bool$ & Right 30 & disjunction ($\disj$) \\
    82   \tt -->       & $[bool,bool]\To bool$ & Right 25 & implication ($\imp$)
    83 \end{constants}
    84 \subcaption{Infixes}
    85 \caption{Syntax of \texttt{HOL}} \label{hol-constants}
    86 \end{figure}
    87 
    88 
    89 \begin{figure}
    90 \index{*let symbol}
    91 \index{*in symbol}
    92 \dquotes
    93 \[\begin{array}{rclcl}
    94     term & = & \hbox{expression of class~$term$} \\
    95          & | & "SOME~" id " . " formula
    96          & | & "\at~" id " . " formula \\
    97          & | & 
    98     \multicolumn{3}{l}{"let"~id~"="~term";"\dots";"~id~"="~term~"in"~term} \\
    99          & | & 
   100     \multicolumn{3}{l}{"if"~formula~"then"~term~"else"~term} \\
   101          & | & "LEAST"~ id " . " formula \\[2ex]
   102  formula & = & \hbox{expression of type~$bool$} \\
   103          & | & term " = " term \\
   104          & | & term " \ttilde= " term \\
   105          & | & term " < " term \\
   106          & | & term " <= " term \\
   107          & | & "\ttilde\ " formula \\
   108          & | & formula " \& " formula \\
   109          & | & formula " | " formula \\
   110          & | & formula " --> " formula \\
   111          & | & "ALL~" id~id^* " . " formula
   112          & | & "!~~~" id~id^* " . " formula \\
   113          & | & "EX~~" id~id^* " . " formula 
   114          & | & "?~~~" id~id^* " . " formula \\
   115          & | & "EX!~" id~id^* " . " formula
   116          & | & "?!~~" id~id^* " . " formula \\
   117   \end{array}
   118 \]
   119 \caption{Full grammar for HOL} \label{hol-grammar}
   120 \end{figure} 
   121 
   122 
   123 \section{Syntax}
   124 
   125 Figure~\ref{hol-constants} lists the constants (including infixes and
   126 binders), while Fig.\ts\ref{hol-grammar} presents the grammar of
   127 higher-order logic.  Note that $a$\verb|~=|$b$ is translated to
   128 $\lnot(a=b)$.
   129 
   130 \begin{warn}
   131   HOL has no if-and-only-if connective; logical equivalence is expressed using
   132   equality.  But equality has a high priority, as befitting a relation, while
   133   if-and-only-if typically has the lowest priority.  Thus, $\lnot\lnot P=P$
   134   abbreviates $\lnot\lnot (P=P)$ and not $(\lnot\lnot P)=P$.  When using $=$
   135   to mean logical equivalence, enclose both operands in parentheses.
   136 \end{warn}
   137 
   138 \subsection{Types and overloading}
   139 The universal type class of higher-order terms is called~\cldx{term}.
   140 By default, explicit type variables have class \cldx{term}.  In
   141 particular the equality symbol and quantifiers are polymorphic over
   142 class \texttt{term}.
   143 
   144 The type of formulae, \tydx{bool}, belongs to class \cldx{term}; thus,
   145 formulae are terms.  The built-in type~\tydx{fun}, which constructs
   146 function types, is overloaded with arity {\tt(term,\thinspace
   147   term)\thinspace term}.  Thus, $\sigma\To\tau$ belongs to class~{\tt
   148   term} if $\sigma$ and~$\tau$ do, allowing quantification over
   149 functions.
   150 
   151 HOL allows new types to be declared as subsets of existing types;
   152 see~{\S}\ref{sec:HOL:Types}.  ML-like datatypes can also be declared; see
   153 ~{\S}\ref{sec:HOL:datatype}.
   154 
   155 Several syntactic type classes --- \cldx{plus}, \cldx{minus},
   156 \cldx{times} and
   157 \cldx{power} --- permit overloading of the operators {\tt+},\index{*"+
   158   symbol} {\tt-}\index{*"- symbol}, {\tt*}.\index{*"* symbol} 
   159 and \verb|^|.\index{^@\verb.^. symbol} 
   160 %
   161 They are overloaded to denote the obvious arithmetic operations on types
   162 \tdx{nat}, \tdx{int} and~\tdx{real}. (With the \verb|^| operator, the
   163 exponent always has type~\tdx{nat}.)  Non-arithmetic overloadings are also
   164 done: the operator {\tt-} can denote set difference, while \verb|^| can
   165 denote exponentiation of relations (iterated composition).  Unary minus is
   166 also written as~{\tt-} and is overloaded like its 2-place counterpart; it even
   167 can stand for set complement.
   168 
   169 The constant \cdx{0} is also overloaded.  It serves as the zero element of
   170 several types, of which the most important is \tdx{nat} (the natural
   171 numbers).  The type class \cldx{plus_ac0} comprises all types for which 0
   172 and~+ satisfy the laws $x+y=y+x$, $(x+y)+z = x+(y+z)$ and $0+x = x$.  These
   173 types include the numeric ones \tdx{nat}, \tdx{int} and~\tdx{real} and also
   174 multisets.  The summation operator \cdx{setsum} is available for all types in
   175 this class. 
   176 
   177 Theory \thydx{Ord} defines the syntactic class \cldx{ord} of order
   178 signatures.  The relations $<$ and $\leq$ are polymorphic over this
   179 class, as are the functions \cdx{mono}, \cdx{min} and \cdx{max}, and
   180 the \cdx{LEAST} operator. \thydx{Ord} also defines a subclass
   181 \cldx{order} of \cldx{ord} which axiomatizes the types that are partially
   182 ordered with respect to~$\leq$.  A further subclass \cldx{linorder} of
   183 \cldx{order} axiomatizes linear orderings.
   184 For details, see the file \texttt{Ord.thy}.
   185                                           
   186 If you state a goal containing overloaded functions, you may need to include
   187 type constraints.  Type inference may otherwise make the goal more
   188 polymorphic than you intended, with confusing results.  For example, the
   189 variables $i$, $j$ and $k$ in the goal $i \leq j \Imp i \leq j+k$ have type
   190 $\alpha::\{ord,plus\}$, although you may have expected them to have some
   191 numeric type, e.g. $nat$.  Instead you should have stated the goal as
   192 $(i::nat) \leq j \Imp i \leq j+k$, which causes all three variables to have
   193 type $nat$.
   194 
   195 \begin{warn}
   196   If resolution fails for no obvious reason, try setting
   197   \ttindex{show_types} to \texttt{true}, causing Isabelle to display
   198   types of terms.  Possibly set \ttindex{show_sorts} to \texttt{true} as
   199   well, causing Isabelle to display type classes and sorts.
   200 
   201   \index{unification!incompleteness of}
   202   Where function types are involved, Isabelle's unification code does not
   203   guarantee to find instantiations for type variables automatically.  Be
   204   prepared to use \ttindex{res_inst_tac} instead of \texttt{resolve_tac},
   205   possibly instantiating type variables.  Setting
   206   \ttindex{Unify.trace_types} to \texttt{true} causes Isabelle to report
   207   omitted search paths during unification.\index{tracing!of unification}
   208 \end{warn}
   209 
   210 
   211 \subsection{Binders}
   212 
   213 Hilbert's {\bf description} operator~$\varepsilon x. P[x]$ stands for some~$x$
   214 satisfying~$P$, if such exists.  Since all terms in HOL denote something, a
   215 description is always meaningful, but we do not know its value unless $P$
   216 defines it uniquely.  We may write descriptions as \cdx{Eps}($\lambda x.
   217 P[x]$) or use the syntax \hbox{\tt SOME~$x$.~$P[x]$}.
   218 
   219 Existential quantification is defined by
   220 \[ \exists x. P~x \;\equiv\; P(\varepsilon x. P~x). \]
   221 The unique existence quantifier, $\exists!x. P$, is defined in terms
   222 of~$\exists$ and~$\forall$.  An Isabelle binder, it admits nested
   223 quantifications.  For instance, $\exists!x\,y. P\,x\,y$ abbreviates
   224 $\exists!x. \exists!y. P\,x\,y$; note that this does not mean that there
   225 exists a unique pair $(x,y)$ satisfying~$P\,x\,y$.
   226 
   227 \medskip
   228 
   229 \index{*"! symbol}\index{*"? symbol}\index{HOL system@{\sc hol} system} The
   230 basic Isabelle/HOL binders have two notations.  Apart from the usual
   231 \texttt{ALL} and \texttt{EX} for $\forall$ and $\exists$, Isabelle/HOL also
   232 supports the original notation of Gordon's {\sc hol} system: \texttt{!}\ 
   233 and~\texttt{?}.  In the latter case, the existential quantifier \emph{must} be
   234 followed by a space; thus {\tt?x} is an unknown, while \verb'? x. f x=y' is a
   235 quantification.  Both notations are accepted for input.  The print mode
   236 ``\ttindexbold{HOL}'' governs the output notation.  If enabled (e.g.\ by
   237 passing option \texttt{-m HOL} to the \texttt{isabelle} executable),
   238 then~{\tt!}\ and~{\tt?}\ are displayed.
   239 
   240 \medskip
   241 
   242 If $\tau$ is a type of class \cldx{ord}, $P$ a formula and $x$ a
   243 variable of type $\tau$, then the term \cdx{LEAST}~$x. P[x]$ is defined
   244 to be the least (w.r.t.\ $\leq$) $x$ such that $P~x$ holds (see
   245 Fig.~\ref{hol-defs}).  The definition uses Hilbert's $\varepsilon$
   246 choice operator, so \texttt{Least} is always meaningful, but may yield
   247 nothing useful in case there is not a unique least element satisfying
   248 $P$.\footnote{Class $ord$ does not require much of its instances, so
   249   $\leq$ need not be a well-ordering, not even an order at all!}
   250 
   251 \medskip All these binders have priority 10.
   252 
   253 \begin{warn}
   254 The low priority of binders means that they need to be enclosed in
   255 parenthesis when they occur in the context of other operations.  For example,
   256 instead of $P \land \forall x. Q$ you need to write $P \land (\forall x. Q)$.
   257 \end{warn}
   258 
   259 
   260 \subsection{The let and case constructions}
   261 Local abbreviations can be introduced by a \texttt{let} construct whose
   262 syntax appears in Fig.\ts\ref{hol-grammar}.  Internally it is translated into
   263 the constant~\cdx{Let}.  It can be expanded by rewriting with its
   264 definition, \tdx{Let_def}.
   265 
   266 HOL also defines the basic syntax
   267 \[\dquotes"case"~e~"of"~c@1~"=>"~e@1~"|" \dots "|"~c@n~"=>"~e@n\] 
   268 as a uniform means of expressing \texttt{case} constructs.  Therefore \texttt{case}
   269 and \sdx{of} are reserved words.  Initially, this is mere syntax and has no
   270 logical meaning.  By declaring translations, you can cause instances of the
   271 \texttt{case} construct to denote applications of particular case operators.
   272 This is what happens automatically for each \texttt{datatype} definition
   273 (see~{\S}\ref{sec:HOL:datatype}).
   274 
   275 \begin{warn}
   276 Both \texttt{if} and \texttt{case} constructs have as low a priority as
   277 quantifiers, which requires additional enclosing parentheses in the context
   278 of most other operations.  For example, instead of $f~x = {\tt if\dots
   279 then\dots else}\dots$ you need to write $f~x = ({\tt if\dots then\dots
   280 else\dots})$.
   281 \end{warn}
   282 
   283 \section{Rules of inference}
   284 
   285 \begin{figure}
   286 \begin{ttbox}\makeatother
   287 \tdx{refl}          t = (t::'a)
   288 \tdx{subst}         [| s = t; P s |] ==> P (t::'a)
   289 \tdx{ext}           (!!x::'a. (f x :: 'b) = g x) ==> (\%x. f x) = (\%x. g x)
   290 \tdx{impI}          (P ==> Q) ==> P-->Q
   291 \tdx{mp}            [| P-->Q;  P |] ==> Q
   292 \tdx{iff}           (P-->Q) --> (Q-->P) --> (P=Q)
   293 \tdx{someI}         P(x::'a) ==> P(@x. P x)
   294 \tdx{True_or_False} (P=True) | (P=False)
   295 \end{ttbox}
   296 \caption{The \texttt{HOL} rules} \label{hol-rules}
   297 \end{figure}
   298 
   299 Figure~\ref{hol-rules} shows the primitive inference rules of~HOL, with
   300 their~{\ML} names.  Some of the rules deserve additional comments:
   301 \begin{ttdescription}
   302 \item[\tdx{ext}] expresses extensionality of functions.
   303 \item[\tdx{iff}] asserts that logically equivalent formulae are
   304   equal.
   305 \item[\tdx{someI}] gives the defining property of the Hilbert
   306   $\varepsilon$-operator.  It is a form of the Axiom of Choice.  The derived rule
   307   \tdx{some_equality} (see below) is often easier to use.
   308 \item[\tdx{True_or_False}] makes the logic classical.\footnote{In
   309     fact, the $\varepsilon$-operator already makes the logic classical, as
   310     shown by Diaconescu; see Paulson~\cite{paulson-COLOG} for details.}
   311 \end{ttdescription}
   312 
   313 
   314 \begin{figure}\hfuzz=4pt%suppress "Overfull \hbox" message
   315 \begin{ttbox}\makeatother
   316 \tdx{True_def}   True     == ((\%x::bool. x)=(\%x. x))
   317 \tdx{All_def}    All      == (\%P. P = (\%x. True))
   318 \tdx{Ex_def}     Ex       == (\%P. P(@x. P x))
   319 \tdx{False_def}  False    == (!P. P)
   320 \tdx{not_def}    not      == (\%P. P-->False)
   321 \tdx{and_def}    op &     == (\%P Q. !R. (P-->Q-->R) --> R)
   322 \tdx{or_def}     op |     == (\%P Q. !R. (P-->R) --> (Q-->R) --> R)
   323 \tdx{Ex1_def}    Ex1      == (\%P. ? x. P x & (! y. P y --> y=x))
   324 
   325 \tdx{o_def}      op o     == (\%(f::'b=>'c) g x::'a. f(g x))
   326 \tdx{if_def}     If P x y ==
   327               (\%P x y. @z::'a.(P=True --> z=x) & (P=False --> z=y))
   328 \tdx{Let_def}    Let s f  == f s
   329 \tdx{Least_def}  Least P  == @x. P(x) & (ALL y. P(y) --> x <= y)"
   330 \end{ttbox}
   331 \caption{The \texttt{HOL} definitions} \label{hol-defs}
   332 \end{figure}
   333 
   334 
   335 HOL follows standard practice in higher-order logic: only a few connectives
   336 are taken as primitive, with the remainder defined obscurely
   337 (Fig.\ts\ref{hol-defs}).  Gordon's {\sc hol} system expresses the
   338 corresponding definitions \cite[page~270]{mgordon-hol} using
   339 object-equality~({\tt=}), which is possible because equality in higher-order
   340 logic may equate formulae and even functions over formulae.  But theory~HOL,
   341 like all other Isabelle theories, uses meta-equality~({\tt==}) for
   342 definitions.
   343 \begin{warn}
   344 The definitions above should never be expanded and are shown for completeness
   345 only.  Instead users should reason in terms of the derived rules shown below
   346 or, better still, using high-level tactics
   347 (see~{\S}\ref{sec:HOL:generic-packages}).
   348 \end{warn}
   349 
   350 Some of the rules mention type variables; for example, \texttt{refl}
   351 mentions the type variable~{\tt'a}.  This allows you to instantiate
   352 type variables explicitly by calling \texttt{res_inst_tac}.
   353 
   354 
   355 \begin{figure}
   356 \begin{ttbox}
   357 \tdx{sym}         s=t ==> t=s
   358 \tdx{trans}       [| r=s; s=t |] ==> r=t
   359 \tdx{ssubst}      [| t=s; P s |] ==> P t
   360 \tdx{box_equals}  [| a=b;  a=c;  b=d |] ==> c=d  
   361 \tdx{arg_cong}    x = y ==> f x = f y
   362 \tdx{fun_cong}    f = g ==> f x = g x
   363 \tdx{cong}        [| f = g; x = y |] ==> f x = g y
   364 \tdx{not_sym}     t ~= s ==> s ~= t
   365 \subcaption{Equality}
   366 
   367 \tdx{TrueI}       True 
   368 \tdx{FalseE}      False ==> P
   369 
   370 \tdx{conjI}       [| P; Q |] ==> P&Q
   371 \tdx{conjunct1}   [| P&Q |] ==> P
   372 \tdx{conjunct2}   [| P&Q |] ==> Q 
   373 \tdx{conjE}       [| P&Q;  [| P; Q |] ==> R |] ==> R
   374 
   375 \tdx{disjI1}      P ==> P|Q
   376 \tdx{disjI2}      Q ==> P|Q
   377 \tdx{disjE}       [| P | Q; P ==> R; Q ==> R |] ==> R
   378 
   379 \tdx{notI}        (P ==> False) ==> ~ P
   380 \tdx{notE}        [| ~ P;  P |] ==> R
   381 \tdx{impE}        [| P-->Q;  P;  Q ==> R |] ==> R
   382 \subcaption{Propositional logic}
   383 
   384 \tdx{iffI}        [| P ==> Q;  Q ==> P |] ==> P=Q
   385 \tdx{iffD1}       [| P=Q; P |] ==> Q
   386 \tdx{iffD2}       [| P=Q; Q |] ==> P
   387 \tdx{iffE}        [| P=Q; [| P --> Q; Q --> P |] ==> R |] ==> R
   388 \subcaption{Logical equivalence}
   389 
   390 \end{ttbox}
   391 \caption{Derived rules for HOL} \label{hol-lemmas1}
   392 \end{figure}
   393 %
   394 %\tdx{eqTrueI}     P ==> P=True 
   395 %\tdx{eqTrueE}     P=True ==> P 
   396 
   397 
   398 \begin{figure}
   399 \begin{ttbox}\makeatother
   400 \tdx{allI}      (!!x. P x) ==> !x. P x
   401 \tdx{spec}      !x. P x ==> P x
   402 \tdx{allE}      [| !x. P x;  P x ==> R |] ==> R
   403 \tdx{all_dupE}  [| !x. P x;  [| P x; !x. P x |] ==> R |] ==> R
   404 
   405 \tdx{exI}       P x ==> ? x. P x
   406 \tdx{exE}       [| ? x. P x; !!x. P x ==> Q |] ==> Q
   407 
   408 \tdx{ex1I}      [| P a;  !!x. P x ==> x=a |] ==> ?! x. P x
   409 \tdx{ex1E}      [| ?! x. P x;  !!x. [| P x;  ! y. P y --> y=x |] ==> R 
   410           |] ==> R
   411 
   412 \tdx{some_equality}   [| P a;  !!x. P x ==> x=a |] ==> (@x. P x) = a
   413 \subcaption{Quantifiers and descriptions}
   414 
   415 \tdx{ccontr}          (~P ==> False) ==> P
   416 \tdx{classical}       (~P ==> P) ==> P
   417 \tdx{excluded_middle} ~P | P
   418 
   419 \tdx{disjCI}       (~Q ==> P) ==> P|Q
   420 \tdx{exCI}         (! x. ~ P x ==> P a) ==> ? x. P x
   421 \tdx{impCE}        [| P-->Q; ~ P ==> R; Q ==> R |] ==> R
   422 \tdx{iffCE}        [| P=Q;  [| P;Q |] ==> R;  [| ~P; ~Q |] ==> R |] ==> R
   423 \tdx{notnotD}      ~~P ==> P
   424 \tdx{swap}         ~P ==> (~Q ==> P) ==> Q
   425 \subcaption{Classical logic}
   426 
   427 \tdx{if_P}         P ==> (if P then x else y) = x
   428 \tdx{if_not_P}     ~ P ==> (if P then x else y) = y
   429 \tdx{split_if}     P(if Q then x else y) = ((Q --> P x) & (~Q --> P y))
   430 \subcaption{Conditionals}
   431 \end{ttbox}
   432 \caption{More derived rules} \label{hol-lemmas2}
   433 \end{figure}
   434 
   435 Some derived rules are shown in Figures~\ref{hol-lemmas1}
   436 and~\ref{hol-lemmas2}, with their {\ML} names.  These include natural rules
   437 for the logical connectives, as well as sequent-style elimination rules for
   438 conjunctions, implications, and universal quantifiers.  
   439 
   440 Note the equality rules: \tdx{ssubst} performs substitution in
   441 backward proofs, while \tdx{box_equals} supports reasoning by
   442 simplifying both sides of an equation.
   443 
   444 The following simple tactics are occasionally useful:
   445 \begin{ttdescription}
   446 \item[\ttindexbold{strip_tac} $i$] applies \texttt{allI} and \texttt{impI}
   447   repeatedly to remove all outermost universal quantifiers and implications
   448   from subgoal $i$.
   449 \item[\ttindexbold{case_tac} {\tt"}$P${\tt"} $i$] performs case distinction on
   450   $P$ for subgoal $i$: the latter is replaced by two identical subgoals with
   451   the added assumptions $P$ and $\lnot P$, respectively.
   452 \item[\ttindexbold{smp_tac} $j$ $i$] applies $j$ times \texttt{spec} and then
   453   \texttt{mp} in subgoal $i$, which is typically useful when forward-chaining 
   454   from an induction hypothesis. As a generalization of \texttt{mp_tac}, 
   455   if there are assumptions $\forall \vec{x}. P \vec{x} \imp Q \vec{x}$ and 
   456   $P \vec{a}$, ($\vec{x}$ being a vector of $j$ variables)
   457   then it replaces the universally quantified implication by $Q \vec{a}$. 
   458   It may instantiate unknowns. It fails if it can do nothing.
   459 \end{ttdescription}
   460 
   461 
   462 \begin{figure} 
   463 \begin{center}
   464 \begin{tabular}{rrr}
   465   \it name      &\it meta-type  & \it description \\ 
   466 \index{{}@\verb'{}' symbol}
   467   \verb|{}|     & $\alpha\,set$         & the empty set \\
   468   \cdx{insert}  & $[\alpha,\alpha\,set]\To \alpha\,set$
   469         & insertion of element \\
   470   \cdx{Collect} & $(\alpha\To bool)\To\alpha\,set$
   471         & comprehension \\
   472   \cdx{INTER} & $[\alpha\,set,\alpha\To\beta\,set]\To\beta\,set$
   473         & intersection over a set\\
   474   \cdx{UNION} & $[\alpha\,set,\alpha\To\beta\,set]\To\beta\,set$
   475         & union over a set\\
   476   \cdx{Inter} & $(\alpha\,set)set\To\alpha\,set$
   477         &set of sets intersection \\
   478   \cdx{Union} & $(\alpha\,set)set\To\alpha\,set$
   479         &set of sets union \\
   480   \cdx{Pow}   & $\alpha\,set \To (\alpha\,set)set$
   481         & powerset \\[1ex]
   482   \cdx{range}   & $(\alpha\To\beta )\To\beta\,set$
   483         & range of a function \\[1ex]
   484   \cdx{Ball}~~\cdx{Bex} & $[\alpha\,set,\alpha\To bool]\To bool$
   485         & bounded quantifiers
   486 \end{tabular}
   487 \end{center}
   488 \subcaption{Constants}
   489 
   490 \begin{center}
   491 \begin{tabular}{llrrr} 
   492   \it symbol &\it name     &\it meta-type & \it priority & \it description \\
   493   \sdx{INT}  & \cdx{INTER1}  & $(\alpha\To\beta\,set)\To\beta\,set$ & 10 & 
   494         intersection\\
   495   \sdx{UN}  & \cdx{UNION1}  & $(\alpha\To\beta\,set)\To\beta\,set$ & 10 & 
   496         union 
   497 \end{tabular}
   498 \end{center}
   499 \subcaption{Binders} 
   500 
   501 \begin{center}
   502 \index{*"`"` symbol}
   503 \index{*": symbol}
   504 \index{*"<"= symbol}
   505 \begin{tabular}{rrrr} 
   506   \it symbol    & \it meta-type & \it priority & \it description \\ 
   507   \tt ``        & $[\alpha\To\beta ,\alpha\,set]\To  \beta\,set$
   508         & Left 90 & image \\
   509   \sdx{Int}     & $[\alpha\,set,\alpha\,set]\To\alpha\,set$
   510         & Left 70 & intersection ($\int$) \\
   511   \sdx{Un}      & $[\alpha\,set,\alpha\,set]\To\alpha\,set$
   512         & Left 65 & union ($\un$) \\
   513   \tt:          & $[\alpha ,\alpha\,set]\To bool$       
   514         & Left 50 & membership ($\in$) \\
   515   \tt <=        & $[\alpha\,set,\alpha\,set]\To bool$
   516         & Left 50 & subset ($\subseteq$) 
   517 \end{tabular}
   518 \end{center}
   519 \subcaption{Infixes}
   520 \caption{Syntax of the theory \texttt{Set}} \label{hol-set-syntax}
   521 \end{figure} 
   522 
   523 
   524 \begin{figure} 
   525 \begin{center} \tt\frenchspacing
   526 \index{*"! symbol}
   527 \begin{tabular}{rrr} 
   528   \it external          & \it internal  & \it description \\ 
   529   $a$ \ttilde: $b$      & \ttilde($a$ : $b$)    & \rm not in\\
   530   {\ttlbrace}$a@1$, $\ldots${\ttrbrace}  &  insert $a@1$ $\ldots$ {\ttlbrace}{\ttrbrace} & \rm finite set \\
   531   {\ttlbrace}$x$. $P[x]${\ttrbrace}        &  Collect($\lambda x. P[x]$) &
   532         \rm comprehension \\
   533   \sdx{INT} $x$:$A$. $B[x]$      & INTER $A$ $\lambda x. B[x]$ &
   534         \rm intersection \\
   535   \sdx{UN}{\tt\ }  $x$:$A$. $B[x]$      & UNION $A$ $\lambda x. B[x]$ &
   536         \rm union \\
   537   \sdx{ALL} $x$:$A$.\ $P[x]$ or \texttt{!} $x$:$A$.\ $P[x]$ &
   538         Ball $A$ $\lambda x.\ P[x]$ & 
   539         \rm bounded $\forall$ \\
   540   \sdx{EX}{\tt\ } $x$:$A$.\ $P[x]$ or \texttt{?} $x$:$A$.\ $P[x]$ & 
   541         Bex $A$ $\lambda x.\ P[x]$ & \rm bounded $\exists$
   542 \end{tabular}
   543 \end{center}
   544 \subcaption{Translations}
   545 
   546 \dquotes
   547 \[\begin{array}{rclcl}
   548     term & = & \hbox{other terms\ldots} \\
   549          & | & "{\ttlbrace}{\ttrbrace}" \\
   550          & | & "{\ttlbrace} " term\; ("," term)^* " {\ttrbrace}" \\
   551          & | & "{\ttlbrace} " id " . " formula " {\ttrbrace}" \\
   552          & | & term " `` " term \\
   553          & | & term " Int " term \\
   554          & | & term " Un " term \\
   555          & | & "INT~~"  id ":" term " . " term \\
   556          & | & "UN~~~"  id ":" term " . " term \\
   557          & | & "INT~~"  id~id^* " . " term \\
   558          & | & "UN~~~"  id~id^* " . " term \\[2ex]
   559  formula & = & \hbox{other formulae\ldots} \\
   560          & | & term " : " term \\
   561          & | & term " \ttilde: " term \\
   562          & | & term " <= " term \\
   563          & | & "ALL " id ":" term " . " formula
   564          & | & "!~" id ":" term " . " formula \\
   565          & | & "EX~~" id ":" term " . " formula
   566          & | & "?~" id ":" term " . " formula \\
   567   \end{array}
   568 \]
   569 \subcaption{Full Grammar}
   570 \caption{Syntax of the theory \texttt{Set} (continued)} \label{hol-set-syntax2}
   571 \end{figure} 
   572 
   573 
   574 \section{A formulation of set theory}
   575 Historically, higher-order logic gives a foundation for Russell and
   576 Whitehead's theory of classes.  Let us use modern terminology and call them
   577 {\bf sets}, but note that these sets are distinct from those of ZF set theory,
   578 and behave more like ZF classes.
   579 \begin{itemize}
   580 \item
   581 Sets are given by predicates over some type~$\sigma$.  Types serve to
   582 define universes for sets, but type-checking is still significant.
   583 \item
   584 There is a universal set (for each type).  Thus, sets have complements, and
   585 may be defined by absolute comprehension.
   586 \item
   587 Although sets may contain other sets as elements, the containing set must
   588 have a more complex type.
   589 \end{itemize}
   590 Finite unions and intersections have the same behaviour in HOL as they do
   591 in~ZF.  In HOL the intersection of the empty set is well-defined, denoting the
   592 universal set for the given type.
   593 
   594 \subsection{Syntax of set theory}\index{*set type}
   595 HOL's set theory is called \thydx{Set}.  The type $\alpha\,set$ is essentially
   596 the same as $\alpha\To bool$.  The new type is defined for clarity and to
   597 avoid complications involving function types in unification.  The isomorphisms
   598 between the two types are declared explicitly.  They are very natural:
   599 \texttt{Collect} maps $\alpha\To bool$ to $\alpha\,set$, while \hbox{\tt op :}
   600 maps in the other direction (ignoring argument order).
   601 
   602 Figure~\ref{hol-set-syntax} lists the constants, infixes, and syntax
   603 translations.  Figure~\ref{hol-set-syntax2} presents the grammar of the new
   604 constructs.  Infix operators include union and intersection ($A\un B$
   605 and $A\int B$), the subset and membership relations, and the image
   606 operator~{\tt``}\@.  Note that $a$\verb|~:|$b$ is translated to
   607 $\lnot(a\in b)$.  
   608 
   609 The $\{a@1,\ldots\}$ notation abbreviates finite sets constructed in
   610 the obvious manner using~\texttt{insert} and~$\{\}$:
   611 \begin{eqnarray*}
   612   \{a, b, c\} & \equiv &
   613   \texttt{insert} \, a \, ({\tt insert} \, b \, ({\tt insert} \, c \, \{\}))
   614 \end{eqnarray*}
   615 
   616 The set \hbox{\tt{\ttlbrace}$x$.\ $P[x]${\ttrbrace}} consists of all $x$ (of
   617 suitable type) that satisfy~$P[x]$, where $P[x]$ is a formula that may contain
   618 free occurrences of~$x$.  This syntax expands to \cdx{Collect}$(\lambda x.
   619 P[x])$.  It defines sets by absolute comprehension, which is impossible in~ZF;
   620 the type of~$x$ implicitly restricts the comprehension.
   621 
   622 The set theory defines two {\bf bounded quantifiers}:
   623 \begin{eqnarray*}
   624    \forall x\in A. P[x] &\hbox{abbreviates}& \forall x. x\in A\imp P[x] \\
   625    \exists x\in A. P[x] &\hbox{abbreviates}& \exists x. x\in A\conj P[x]
   626 \end{eqnarray*}
   627 The constants~\cdx{Ball} and~\cdx{Bex} are defined
   628 accordingly.  Instead of \texttt{Ball $A$ $P$} and \texttt{Bex $A$ $P$} we may
   629 write\index{*"! symbol}\index{*"? symbol}
   630 \index{*ALL symbol}\index{*EX symbol} 
   631 %
   632 \hbox{\tt ALL~$x$:$A$.\ $P[x]$} and \hbox{\tt EX~$x$:$A$.\ $P[x]$}.  The
   633 original notation of Gordon's {\sc hol} system is supported as well:
   634 \texttt{!}\ and \texttt{?}.
   635 
   636 Unions and intersections over sets, namely $\bigcup@{x\in A}B[x]$ and
   637 $\bigcap@{x\in A}B[x]$, are written 
   638 \sdx{UN}~\hbox{\tt$x$:$A$.\ $B[x]$} and
   639 \sdx{INT}~\hbox{\tt$x$:$A$.\ $B[x]$}.  
   640 
   641 Unions and intersections over types, namely $\bigcup@x B[x]$ and $\bigcap@x
   642 B[x]$, are written \sdx{UN}~\hbox{\tt$x$.\ $B[x]$} and
   643 \sdx{INT}~\hbox{\tt$x$.\ $B[x]$}.  They are equivalent to the previous
   644 union and intersection operators when $A$ is the universal set.
   645 
   646 The operators $\bigcup A$ and $\bigcap A$ act upon sets of sets.  They are
   647 not binders, but are equal to $\bigcup@{x\in A}x$ and $\bigcap@{x\in A}x$,
   648 respectively.
   649 
   650 
   651 
   652 \begin{figure} \underscoreon
   653 \begin{ttbox}
   654 \tdx{mem_Collect_eq}    (a : {\ttlbrace}x. P x{\ttrbrace}) = P a
   655 \tdx{Collect_mem_eq}    {\ttlbrace}x. x:A{\ttrbrace} = A
   656 
   657 \tdx{empty_def}         {\ttlbrace}{\ttrbrace}          == {\ttlbrace}x. False{\ttrbrace}
   658 \tdx{insert_def}        insert a B  == {\ttlbrace}x. x=a{\ttrbrace} Un B
   659 \tdx{Ball_def}          Ball A P    == ! x. x:A --> P x
   660 \tdx{Bex_def}           Bex A P     == ? x. x:A & P x
   661 \tdx{subset_def}        A <= B      == ! x:A. x:B
   662 \tdx{Un_def}            A Un B      == {\ttlbrace}x. x:A | x:B{\ttrbrace}
   663 \tdx{Int_def}           A Int B     == {\ttlbrace}x. x:A & x:B{\ttrbrace}
   664 \tdx{set_diff_def}      A - B       == {\ttlbrace}x. x:A & x~:B{\ttrbrace}
   665 \tdx{Compl_def}         -A          == {\ttlbrace}x. ~ x:A{\ttrbrace}
   666 \tdx{INTER_def}         INTER A B   == {\ttlbrace}y. ! x:A. y: B x{\ttrbrace}
   667 \tdx{UNION_def}         UNION A B   == {\ttlbrace}y. ? x:A. y: B x{\ttrbrace}
   668 \tdx{INTER1_def}        INTER1 B    == INTER {\ttlbrace}x. True{\ttrbrace} B 
   669 \tdx{UNION1_def}        UNION1 B    == UNION {\ttlbrace}x. True{\ttrbrace} B 
   670 \tdx{Inter_def}         Inter S     == (INT x:S. x)
   671 \tdx{Union_def}         Union S     == (UN  x:S. x)
   672 \tdx{Pow_def}           Pow A       == {\ttlbrace}B. B <= A{\ttrbrace}
   673 \tdx{image_def}         f``A        == {\ttlbrace}y. ? x:A. y=f x{\ttrbrace}
   674 \tdx{range_def}         range f     == {\ttlbrace}y. ? x. y=f x{\ttrbrace}
   675 \end{ttbox}
   676 \caption{Rules of the theory \texttt{Set}} \label{hol-set-rules}
   677 \end{figure}
   678 
   679 
   680 \begin{figure} \underscoreon
   681 \begin{ttbox}
   682 \tdx{CollectI}        [| P a |] ==> a : {\ttlbrace}x. P x{\ttrbrace}
   683 \tdx{CollectD}        [| a : {\ttlbrace}x. P x{\ttrbrace} |] ==> P a
   684 \tdx{CollectE}        [| a : {\ttlbrace}x. P x{\ttrbrace};  P a ==> W |] ==> W
   685 
   686 \tdx{ballI}           [| !!x. x:A ==> P x |] ==> ! x:A. P x
   687 \tdx{bspec}           [| ! x:A. P x;  x:A |] ==> P x
   688 \tdx{ballE}           [| ! x:A. P x;  P x ==> Q;  ~ x:A ==> Q |] ==> Q
   689 
   690 \tdx{bexI}            [| P x;  x:A |] ==> ? x:A. P x
   691 \tdx{bexCI}           [| ! x:A. ~ P x ==> P a;  a:A |] ==> ? x:A. P x
   692 \tdx{bexE}            [| ? x:A. P x;  !!x. [| x:A; P x |] ==> Q  |] ==> Q
   693 \subcaption{Comprehension and Bounded quantifiers}
   694 
   695 \tdx{subsetI}         (!!x. x:A ==> x:B) ==> A <= B
   696 \tdx{subsetD}         [| A <= B;  c:A |] ==> c:B
   697 \tdx{subsetCE}        [| A <= B;  ~ (c:A) ==> P;  c:B ==> P |] ==> P
   698 
   699 \tdx{subset_refl}     A <= A
   700 \tdx{subset_trans}    [| A<=B;  B<=C |] ==> A<=C
   701 
   702 \tdx{equalityI}       [| A <= B;  B <= A |] ==> A = B
   703 \tdx{equalityD1}      A = B ==> A<=B
   704 \tdx{equalityD2}      A = B ==> B<=A
   705 \tdx{equalityE}       [| A = B;  [| A<=B; B<=A |] ==> P |]  ==>  P
   706 
   707 \tdx{equalityCE}      [| A = B;  [| c:A; c:B |] ==> P;  
   708                            [| ~ c:A; ~ c:B |] ==> P 
   709                 |]  ==>  P
   710 \subcaption{The subset and equality relations}
   711 \end{ttbox}
   712 \caption{Derived rules for set theory} \label{hol-set1}
   713 \end{figure}
   714 
   715 
   716 \begin{figure} \underscoreon
   717 \begin{ttbox}
   718 \tdx{emptyE}   a : {\ttlbrace}{\ttrbrace} ==> P
   719 
   720 \tdx{insertI1} a : insert a B
   721 \tdx{insertI2} a : B ==> a : insert b B
   722 \tdx{insertE}  [| a : insert b A;  a=b ==> P;  a:A ==> P |] ==> P
   723 
   724 \tdx{ComplI}   [| c:A ==> False |] ==> c : -A
   725 \tdx{ComplD}   [| c : -A |] ==> ~ c:A
   726 
   727 \tdx{UnI1}     c:A ==> c : A Un B
   728 \tdx{UnI2}     c:B ==> c : A Un B
   729 \tdx{UnCI}     (~c:B ==> c:A) ==> c : A Un B
   730 \tdx{UnE}      [| c : A Un B;  c:A ==> P;  c:B ==> P |] ==> P
   731 
   732 \tdx{IntI}     [| c:A;  c:B |] ==> c : A Int B
   733 \tdx{IntD1}    c : A Int B ==> c:A
   734 \tdx{IntD2}    c : A Int B ==> c:B
   735 \tdx{IntE}     [| c : A Int B;  [| c:A; c:B |] ==> P |] ==> P
   736 
   737 \tdx{UN_I}     [| a:A;  b: B a |] ==> b: (UN x:A. B x)
   738 \tdx{UN_E}     [| b: (UN x:A. B x);  !!x.[| x:A;  b:B x |] ==> R |] ==> R
   739 
   740 \tdx{INT_I}    (!!x. x:A ==> b: B x) ==> b : (INT x:A. B x)
   741 \tdx{INT_D}    [| b: (INT x:A. B x);  a:A |] ==> b: B a
   742 \tdx{INT_E}    [| b: (INT x:A. B x);  b: B a ==> R;  ~ a:A ==> R |] ==> R
   743 
   744 \tdx{UnionI}   [| X:C;  A:X |] ==> A : Union C
   745 \tdx{UnionE}   [| A : Union C;  !!X.[| A:X;  X:C |] ==> R |] ==> R
   746 
   747 \tdx{InterI}   [| !!X. X:C ==> A:X |] ==> A : Inter C
   748 \tdx{InterD}   [| A : Inter C;  X:C |] ==> A:X
   749 \tdx{InterE}   [| A : Inter C;  A:X ==> R;  ~ X:C ==> R |] ==> R
   750 
   751 \tdx{PowI}     A<=B ==> A: Pow B
   752 \tdx{PowD}     A: Pow B ==> A<=B
   753 
   754 \tdx{imageI}   [| x:A |] ==> f x : f``A
   755 \tdx{imageE}   [| b : f``A;  !!x.[| b=f x;  x:A |] ==> P |] ==> P
   756 
   757 \tdx{rangeI}   f x : range f
   758 \tdx{rangeE}   [| b : range f;  !!x.[| b=f x |] ==> P |] ==> P
   759 \end{ttbox}
   760 \caption{Further derived rules for set theory} \label{hol-set2}
   761 \end{figure}
   762 
   763 
   764 \subsection{Axioms and rules of set theory}
   765 Figure~\ref{hol-set-rules} presents the rules of theory \thydx{Set}.  The
   766 axioms \tdx{mem_Collect_eq} and \tdx{Collect_mem_eq} assert
   767 that the functions \texttt{Collect} and \hbox{\tt op :} are isomorphisms.  Of
   768 course, \hbox{\tt op :} also serves as the membership relation.
   769 
   770 All the other axioms are definitions.  They include the empty set, bounded
   771 quantifiers, unions, intersections, complements and the subset relation.
   772 They also include straightforward constructions on functions: image~({\tt``})
   773 and \texttt{range}.
   774 
   775 %The predicate \cdx{inj_on} is used for simulating type definitions.
   776 %The statement ${\tt inj_on}~f~A$ asserts that $f$ is injective on the
   777 %set~$A$, which specifies a subset of its domain type.  In a type
   778 %definition, $f$ is the abstraction function and $A$ is the set of valid
   779 %representations; we should not expect $f$ to be injective outside of~$A$.
   780 
   781 %\begin{figure} \underscoreon
   782 %\begin{ttbox}
   783 %\tdx{Inv_f_f}    inj f ==> Inv f (f x) = x
   784 %\tdx{f_Inv_f}    y : range f ==> f(Inv f y) = y
   785 %
   786 %\tdx{Inv_injective}
   787 %    [| Inv f x=Inv f y; x: range f;  y: range f |] ==> x=y
   788 %
   789 %
   790 %\tdx{monoI}      [| !!A B. A <= B ==> f A <= f B |] ==> mono f
   791 %\tdx{monoD}      [| mono f;  A <= B |] ==> f A <= f B
   792 %
   793 %\tdx{injI}       [| !! x y. f x = f y ==> x=y |] ==> inj f
   794 %\tdx{inj_inverseI}              (!!x. g(f x) = x) ==> inj f
   795 %\tdx{injD}       [| inj f; f x = f y |] ==> x=y
   796 %
   797 %\tdx{inj_onI}  (!!x y. [| f x=f y; x:A; y:A |] ==> x=y) ==> inj_on f A
   798 %\tdx{inj_onD}  [| inj_on f A;  f x=f y;  x:A;  y:A |] ==> x=y
   799 %
   800 %\tdx{inj_on_inverseI}
   801 %    (!!x. x:A ==> g(f x) = x) ==> inj_on f A
   802 %\tdx{inj_on_contraD}
   803 %    [| inj_on f A;  x~=y;  x:A;  y:A |] ==> ~ f x=f y
   804 %\end{ttbox}
   805 %\caption{Derived rules involving functions} \label{hol-fun}
   806 %\end{figure}
   807 
   808 
   809 \begin{figure} \underscoreon
   810 \begin{ttbox}
   811 \tdx{Union_upper}     B:A ==> B <= Union A
   812 \tdx{Union_least}     [| !!X. X:A ==> X<=C |] ==> Union A <= C
   813 
   814 \tdx{Inter_lower}     B:A ==> Inter A <= B
   815 \tdx{Inter_greatest}  [| !!X. X:A ==> C<=X |] ==> C <= Inter A
   816 
   817 \tdx{Un_upper1}       A <= A Un B
   818 \tdx{Un_upper2}       B <= A Un B
   819 \tdx{Un_least}        [| A<=C;  B<=C |] ==> A Un B <= C
   820 
   821 \tdx{Int_lower1}      A Int B <= A
   822 \tdx{Int_lower2}      A Int B <= B
   823 \tdx{Int_greatest}    [| C<=A;  C<=B |] ==> C <= A Int B
   824 \end{ttbox}
   825 \caption{Derived rules involving subsets} \label{hol-subset}
   826 \end{figure}
   827 
   828 
   829 \begin{figure} \underscoreon   \hfuzz=4pt%suppress "Overfull \hbox" message
   830 \begin{ttbox}
   831 \tdx{Int_absorb}        A Int A = A
   832 \tdx{Int_commute}       A Int B = B Int A
   833 \tdx{Int_assoc}         (A Int B) Int C  =  A Int (B Int C)
   834 \tdx{Int_Un_distrib}    (A Un B)  Int C  =  (A Int C) Un (B Int C)
   835 
   836 \tdx{Un_absorb}         A Un A = A
   837 \tdx{Un_commute}        A Un B = B Un A
   838 \tdx{Un_assoc}          (A Un B)  Un C  =  A Un (B Un C)
   839 \tdx{Un_Int_distrib}    (A Int B) Un C  =  (A Un C) Int (B Un C)
   840 
   841 \tdx{Compl_disjoint}    A Int (-A) = {\ttlbrace}x. False{\ttrbrace}
   842 \tdx{Compl_partition}   A Un  (-A) = {\ttlbrace}x. True{\ttrbrace}
   843 \tdx{double_complement} -(-A) = A
   844 \tdx{Compl_Un}          -(A Un B)  = (-A) Int (-B)
   845 \tdx{Compl_Int}         -(A Int B) = (-A) Un (-B)
   846 
   847 \tdx{Union_Un_distrib}  Union(A Un B) = (Union A) Un (Union B)
   848 \tdx{Int_Union}         A Int (Union B) = (UN C:B. A Int C)
   849 
   850 \tdx{Inter_Un_distrib}  Inter(A Un B) = (Inter A) Int (Inter B)
   851 \tdx{Un_Inter}          A Un (Inter B) = (INT C:B. A Un C)
   852 
   853 \end{ttbox}
   854 \caption{Set equalities} \label{hol-equalities}
   855 \end{figure}
   856 %\tdx{Un_Union_image}    (UN x:C.(A x) Un (B x)) = Union(A``C) Un Union(B``C)
   857 %\tdx{Int_Inter_image}   (INT x:C.(A x) Int (B x)) = Inter(A``C) Int Inter(B``C)
   858 
   859 Figures~\ref{hol-set1} and~\ref{hol-set2} present derived rules.  Most are
   860 obvious and resemble rules of Isabelle's ZF set theory.  Certain rules, such
   861 as \tdx{subsetCE}, \tdx{bexCI} and \tdx{UnCI}, are designed for classical
   862 reasoning; the rules \tdx{subsetD}, \tdx{bexI}, \tdx{Un1} and~\tdx{Un2} are
   863 not strictly necessary but yield more natural proofs.  Similarly,
   864 \tdx{equalityCE} supports classical reasoning about extensionality, after the
   865 fashion of \tdx{iffCE}.  See the file \texttt{HOL/Set.ML} for proofs
   866 pertaining to set theory.
   867 
   868 Figure~\ref{hol-subset} presents lattice properties of the subset relation.
   869 Unions form least upper bounds; non-empty intersections form greatest lower
   870 bounds.  Reasoning directly about subsets often yields clearer proofs than
   871 reasoning about the membership relation.  See the file \texttt{HOL/subset.ML}.
   872 
   873 Figure~\ref{hol-equalities} presents many common set equalities.  They
   874 include commutative, associative and distributive laws involving unions,
   875 intersections and complements.  For a complete listing see the file {\tt
   876 HOL/equalities.ML}.
   877 
   878 \begin{warn}
   879 \texttt{Blast_tac} proves many set-theoretic theorems automatically.
   880 Hence you seldom need to refer to the theorems above.
   881 \end{warn}
   882 
   883 \begin{figure}
   884 \begin{center}
   885 \begin{tabular}{rrr}
   886   \it name      &\it meta-type  & \it description \\ 
   887   \cdx{inj}~~\cdx{surj}& $(\alpha\To\beta )\To bool$
   888         & injective/surjective \\
   889   \cdx{inj_on}        & $[\alpha\To\beta ,\alpha\,set]\To bool$
   890         & injective over subset\\
   891   \cdx{inv} & $(\alpha\To\beta)\To(\beta\To\alpha)$ & inverse function
   892 \end{tabular}
   893 \end{center}
   894 
   895 \underscoreon
   896 \begin{ttbox}
   897 \tdx{inj_def}         inj f      == ! x y. f x=f y --> x=y
   898 \tdx{surj_def}        surj f     == ! y. ? x. y=f x
   899 \tdx{inj_on_def}      inj_on f A == !x:A. !y:A. f x=f y --> x=y
   900 \tdx{inv_def}         inv f      == (\%y. @x. f(x)=y)
   901 \end{ttbox}
   902 \caption{Theory \thydx{Fun}} \label{fig:HOL:Fun}
   903 \end{figure}
   904 
   905 \subsection{Properties of functions}\nopagebreak
   906 Figure~\ref{fig:HOL:Fun} presents a theory of simple properties of functions.
   907 Note that ${\tt inv}~f$ uses Hilbert's $\varepsilon$ to yield an inverse
   908 of~$f$.  See the file \texttt{HOL/Fun.ML} for a complete listing of the derived
   909 rules.  Reasoning about function composition (the operator~\sdx{o}) and the
   910 predicate~\cdx{surj} is done simply by expanding the definitions.
   911 
   912 There is also a large collection of monotonicity theorems for constructions
   913 on sets in the file \texttt{HOL/mono.ML}.
   914 
   915 
   916 \section{Generic packages}
   917 \label{sec:HOL:generic-packages}
   918 
   919 HOL instantiates most of Isabelle's generic packages, making available the
   920 simplifier and the classical reasoner.
   921 
   922 \subsection{Simplification and substitution}
   923 
   924 Simplification tactics tactics such as \texttt{Asm_simp_tac} and \texttt{Full_simp_tac} use the default simpset
   925 (\texttt{simpset()}), which works for most purposes.  A quite minimal
   926 simplification set for higher-order logic is~\ttindexbold{HOL_ss};
   927 even more frugal is \ttindexbold{HOL_basic_ss}.  Equality~($=$), which
   928 also expresses logical equivalence, may be used for rewriting.  See
   929 the file \texttt{HOL/simpdata.ML} for a complete listing of the basic
   930 simplification rules.
   931 
   932 See \iflabelundefined{chap:classical}{the {\em Reference Manual\/}}%
   933 {Chaps.\ts\ref{substitution} and~\ref{simp-chap}} for details of substitution
   934 and simplification.
   935 
   936 \begin{warn}\index{simplification!of conjunctions}%
   937   Reducing $a=b\conj P(a)$ to $a=b\conj P(b)$ is sometimes advantageous.  The
   938   left part of a conjunction helps in simplifying the right part.  This effect
   939   is not available by default: it can be slow.  It can be obtained by
   940   including \ttindex{conj_cong} in a simpset, \verb$addcongs [conj_cong]$.
   941 \end{warn}
   942 
   943 \begin{warn}\index{simplification!of \texttt{if}}\label{if-simp}%
   944   By default only the condition of an \ttindex{if} is simplified but not the
   945   \texttt{then} and \texttt{else} parts. Of course the latter are simplified
   946   once the condition simplifies to \texttt{True} or \texttt{False}. To ensure
   947   full simplification of all parts of a conditional you must remove
   948   \ttindex{if_weak_cong} from the simpset, \verb$delcongs [if_weak_cong]$.
   949 \end{warn}
   950 
   951 If the simplifier cannot use a certain rewrite rule --- either because
   952 of nontermination or because its left-hand side is too flexible ---
   953 then you might try \texttt{stac}:
   954 \begin{ttdescription}
   955 \item[\ttindexbold{stac} $thm$ $i,$] where $thm$ is of the form $lhs = rhs$,
   956   replaces in subgoal $i$ instances of $lhs$ by corresponding instances of
   957   $rhs$.  In case of multiple instances of $lhs$ in subgoal $i$, backtracking
   958   may be necessary to select the desired ones.
   959 
   960 If $thm$ is a conditional equality, the instantiated condition becomes an
   961 additional (first) subgoal.
   962 \end{ttdescription}
   963 
   964 HOL provides the tactic \ttindex{hyp_subst_tac}, which substitutes for an
   965 equality throughout a subgoal and its hypotheses.  This tactic uses HOL's
   966 general substitution rule.
   967 
   968 \subsubsection{Case splitting}
   969 \label{subsec:HOL:case:splitting}
   970 
   971 HOL also provides convenient means for case splitting during rewriting. Goals
   972 containing a subterm of the form \texttt{if}~$b$~{\tt then\dots else\dots}
   973 often require a case distinction on $b$. This is expressed by the theorem
   974 \tdx{split_if}:
   975 $$
   976 \Var{P}(\mbox{\tt if}~\Var{b}~{\tt then}~\Var{x}~\mbox{\tt else}~\Var{y})~=~
   977 ((\Var{b} \to \Var{P}(\Var{x})) \land (\lnot \Var{b} \to \Var{P}(\Var{y})))
   978 \eqno{(*)}
   979 $$
   980 For example, a simple instance of $(*)$ is
   981 \[
   982 x \in (\mbox{\tt if}~x \in A~{\tt then}~A~\mbox{\tt else}~\{x\})~=~
   983 ((x \in A \to x \in A) \land (x \notin A \to x \in \{x\}))
   984 \]
   985 Because $(*)$ is too general as a rewrite rule for the simplifier (the
   986 left-hand side is not a higher-order pattern in the sense of
   987 \iflabelundefined{chap:simplification}{the {\em Reference Manual\/}}%
   988 {Chap.\ts\ref{chap:simplification}}), there is a special infix function 
   989 \ttindexbold{addsplits} of type \texttt{simpset * thm list -> simpset}
   990 (analogous to \texttt{addsimps}) that adds rules such as $(*)$ to a
   991 simpset, as in
   992 \begin{ttbox}
   993 by(simp_tac (simpset() addsplits [split_if]) 1);
   994 \end{ttbox}
   995 The effect is that after each round of simplification, one occurrence of
   996 \texttt{if} is split acording to \texttt{split_if}, until all occurences of
   997 \texttt{if} have been eliminated.
   998 
   999 It turns out that using \texttt{split_if} is almost always the right thing to
  1000 do. Hence \texttt{split_if} is already included in the default simpset. If
  1001 you want to delete it from a simpset, use \ttindexbold{delsplits}, which is
  1002 the inverse of \texttt{addsplits}:
  1003 \begin{ttbox}
  1004 by(simp_tac (simpset() delsplits [split_if]) 1);
  1005 \end{ttbox}
  1006 
  1007 In general, \texttt{addsplits} accepts rules of the form
  1008 \[
  1009 \Var{P}(c~\Var{x@1}~\dots~\Var{x@n})~=~ rhs
  1010 \]
  1011 where $c$ is a constant and $rhs$ is arbitrary. Note that $(*)$ is of the
  1012 right form because internally the left-hand side is
  1013 $\Var{P}(\mathtt{If}~\Var{b}~\Var{x}~~\Var{y})$. Important further examples
  1014 are splitting rules for \texttt{case} expressions (see~{\S}\ref{subsec:list}
  1015 and~{\S}\ref{subsec:datatype:basics}).
  1016 
  1017 Analogous to \texttt{Addsimps} and \texttt{Delsimps}, there are also
  1018 imperative versions of \texttt{addsplits} and \texttt{delsplits}
  1019 \begin{ttbox}
  1020 \ttindexbold{Addsplits}: thm list -> unit
  1021 \ttindexbold{Delsplits}: thm list -> unit
  1022 \end{ttbox}
  1023 for adding splitting rules to, and deleting them from the current simpset.
  1024 
  1025 \subsection{Classical reasoning}
  1026 
  1027 HOL derives classical introduction rules for $\disj$ and~$\exists$, as well as
  1028 classical elimination rules for~$\imp$ and~$\bimp$, and the swap rule; recall
  1029 Fig.\ts\ref{hol-lemmas2} above.
  1030 
  1031 The classical reasoner is installed.  Tactics such as \texttt{Blast_tac} and
  1032 {\tt Best_tac} refer to the default claset (\texttt{claset()}), which works
  1033 for most purposes.  Named clasets include \ttindexbold{prop_cs}, which
  1034 includes the propositional rules, and \ttindexbold{HOL_cs}, which also
  1035 includes quantifier rules.  See the file \texttt{HOL/cladata.ML} for lists of
  1036 the classical rules,
  1037 and \iflabelundefined{chap:classical}{the {\em Reference Manual\/}}%
  1038 {Chap.\ts\ref{chap:classical}} for more discussion of classical proof methods.
  1039 
  1040 
  1041 %FIXME outdated, both from the Isabelle and SVC perspective
  1042 % \section{Calling the decision procedure SVC}\label{sec:HOL:SVC}
  1043 
  1044 % \index{SVC decision procedure|(}
  1045 
  1046 % The Stanford Validity Checker (SVC) is a tool that can check the validity of
  1047 % certain types of formulae.  If it is installed on your machine, then
  1048 % Isabelle/HOL can be configured to call it through the tactic
  1049 % \ttindex{svc_tac}.  It is ideal for large tautologies and complex problems in
  1050 % linear arithmetic.  Subexpressions that SVC cannot handle are automatically
  1051 % replaced by variables, so you can call the tactic on any subgoal.  See the
  1052 % file \texttt{HOL/ex/svc_test.ML} for examples.
  1053 % \begin{ttbox} 
  1054 % svc_tac   : int -> tactic
  1055 % Svc.trace : bool ref      \hfill{\bf initially false}
  1056 % \end{ttbox}
  1057 
  1058 % \begin{ttdescription}
  1059 % \item[\ttindexbold{svc_tac} $i$] attempts to prove subgoal~$i$ by translating
  1060 %   it into a formula recognized by~SVC\@.  If it succeeds then the subgoal is
  1061 %   removed.  It fails if SVC is unable to prove the subgoal.  It crashes with
  1062 %   an error message if SVC appears not to be installed.  Numeric variables may
  1063 %   have types \texttt{nat}, \texttt{int} or \texttt{real}.
  1064 
  1065 % \item[\ttindexbold{Svc.trace}] is a flag that, if set, causes \texttt{svc_tac}
  1066 %   to trace its operations: abstraction of the subgoal, translation to SVC
  1067 %   syntax, SVC's response.
  1068 % \end{ttdescription}
  1069 
  1070 % Here is an example, with tracing turned on:
  1071 % \begin{ttbox}
  1072 % set Svc.trace;
  1073 % {\out val it : bool = true}
  1074 % Goal "(#3::nat)*a <= #2 + #4*b + #6*c  & #11 <= #2*a + b + #2*c & \ttback
  1075 % \ttback     a + #3*b <= #5 + #2*c  --> #2 + #3*b <= #2*a + #6*c";
  1076 
  1077 % by (svc_tac 1);
  1078 % {\out Subgoal abstracted to}
  1079 % {\out #3 * a <= #2 + #4 * b + #6 * c &}
  1080 % {\out #11 <= #2 * a + b + #2 * c & a + #3 * b <= #5 + #2 * c -->}
  1081 % {\out #2 + #3 * b <= #2 * a + #6 * c}
  1082 % {\out Calling SVC:}
  1083 % {\out (=> (<= 0  (F_c) )  (=> (<= 0  (F_b) )  (=> (<= 0  (F_a) )}
  1084 % {\out   (=> (AND (<= {* 3  (F_a) }  {+ {+ 2  {* 4  (F_b) } }  }
  1085 % {\out {* 6  (F_c) } } )  (AND (<= 11  {+ {+ {* 2  (F_a) }  (F_b) }}
  1086 % {\out   {* 2  (F_c) } } )  (<= {+ (F_a)  {* 3  (F_b) } }  {+ 5  }
  1087 % {\out {* 2  (F_c) } } ) ) )  (< {+ 2  {* 3  (F_b) } }  {+ 1  {+ }
  1088 % {\out {* 2  (F_a) }  {* 6  (F_c) } } } ) ) ) ) ) }
  1089 % {\out SVC Returns:}
  1090 % {\out VALID}
  1091 % {\out Level 1}
  1092 % {\out #3 * a <= #2 + #4 * b + #6 * c &}
  1093 % {\out #11 <= #2 * a + b + #2 * c & a + #3 * b <= #5 + #2 * c -->}
  1094 % {\out #2 + #3 * b <= #2 * a + #6 * c}
  1095 % {\out No subgoals!}
  1096 % \end{ttbox}
  1097 
  1098 
  1099 % \begin{warn}
  1100 % Calling \ttindex{svc_tac} entails an above-average risk of
  1101 % unsoundness.  Isabelle does not check SVC's result independently.  Moreover,
  1102 % the tactic translates the submitted formula using code that lies outside
  1103 % Isabelle's inference core.  Theorems that depend upon results proved using SVC
  1104 % (and other oracles) are displayed with the annotation \texttt{[!]} attached.
  1105 % You can also use \texttt{\#der (rep_thm $th$)} to examine the proof object of
  1106 % theorem~$th$, as described in the \emph{Reference Manual}.  
  1107 % \end{warn}
  1108 
  1109 % To start, first download SVC from the Internet at URL
  1110 % \begin{ttbox}
  1111 % http://agamemnon.stanford.edu/~levitt/vc/index.html
  1112 % \end{ttbox}
  1113 % and install it using the instructions supplied.  SVC requires two environment
  1114 % variables:
  1115 % \begin{ttdescription}
  1116 % \item[\ttindexbold{SVC_HOME}] is an absolute pathname to the SVC
  1117 %     distribution directory. 
  1118     
  1119 %   \item[\ttindexbold{SVC_MACHINE}] identifies the type of computer and
  1120 %     operating system.  
  1121 % \end{ttdescription}
  1122 % You can set these environment variables either using the Unix shell or through
  1123 % an Isabelle \texttt{settings} file.  Isabelle assumes SVC to be installed if
  1124 % \texttt{SVC_HOME} is defined.
  1125 
  1126 % \paragraph*{Acknowledgement.}  This interface uses code supplied by S{\o}ren
  1127 % Heilmann.
  1128 
  1129 
  1130 % \index{SVC decision procedure|)}
  1131 
  1132 
  1133 
  1134 
  1135 \section{Types}\label{sec:HOL:Types}
  1136 This section describes HOL's basic predefined types ($\alpha \times \beta$,
  1137 $\alpha + \beta$, $nat$ and $\alpha \; list$) and ways for introducing new
  1138 types in general.  The most important type construction, the
  1139 \texttt{datatype}, is treated separately in {\S}\ref{sec:HOL:datatype}.
  1140 
  1141 
  1142 \subsection{Product and sum types}\index{*"* type}\index{*"+ type}
  1143 \label{subsec:prod-sum}
  1144 
  1145 \begin{figure}[htbp]
  1146 \begin{constants}
  1147   \it symbol    & \it meta-type &           & \it description \\ 
  1148   \cdx{Pair}    & $[\alpha,\beta]\To \alpha\times\beta$
  1149         & & ordered pairs $(a,b)$ \\
  1150   \cdx{fst}     & $\alpha\times\beta \To \alpha$        & & first projection\\
  1151   \cdx{snd}     & $\alpha\times\beta \To \beta$         & & second projection\\
  1152   \cdx{split}   & $[[\alpha,\beta]\To\gamma, \alpha\times\beta] \To \gamma$ 
  1153         & & generalized projection\\
  1154   \cdx{Sigma}  & 
  1155         $[\alpha\,set, \alpha\To\beta\,set]\To(\alpha\times\beta)set$ &
  1156         & general sum of sets
  1157 \end{constants}
  1158 %\tdx{fst_def}      fst p     == @a. ? b. p = (a,b)
  1159 %\tdx{snd_def}      snd p     == @b. ? a. p = (a,b)
  1160 %\tdx{split_def}    split c p == c (fst p) (snd p)
  1161 \begin{ttbox}\makeatletter
  1162 \tdx{Sigma_def}    Sigma A B == UN x:A. UN y:B x. {\ttlbrace}(x,y){\ttrbrace}
  1163 
  1164 \tdx{Pair_eq}      ((a,b) = (a',b')) = (a=a' & b=b')
  1165 \tdx{Pair_inject}  [| (a, b) = (a',b');  [| a=a';  b=b' |] ==> R |] ==> R
  1166 \tdx{PairE}        [| !!x y. p = (x,y) ==> Q |] ==> Q
  1167 
  1168 \tdx{fst_conv}     fst (a,b) = a
  1169 \tdx{snd_conv}     snd (a,b) = b
  1170 \tdx{surjective_pairing}  p = (fst p,snd p)
  1171 
  1172 \tdx{split}        split c (a,b) = c a b
  1173 \tdx{split_split}  R(split c p) = (! x y. p = (x,y) --> R(c x y))
  1174 
  1175 \tdx{SigmaI}    [| a:A;  b:B a |] ==> (a,b) : Sigma A B
  1176 
  1177 \tdx{SigmaE}    [| c:Sigma A B; !!x y.[| x:A; y:B x; c=(x,y) |] ==> P 
  1178           |] ==> P
  1179 \end{ttbox}
  1180 \caption{Type $\alpha\times\beta$}\label{hol-prod}
  1181 \end{figure} 
  1182 
  1183 Theory \thydx{Prod} (Fig.\ts\ref{hol-prod}) defines the product type
  1184 $\alpha\times\beta$, with the ordered pair syntax $(a, b)$.  General
  1185 tuples are simulated by pairs nested to the right:
  1186 \begin{center}
  1187 \begin{tabular}{c|c}
  1188 external & internal \\
  1189 \hline
  1190 $\tau@1 \times \dots \times \tau@n$ & $\tau@1 \times (\dots (\tau@{n-1} \times \tau@n)\dots)$ \\
  1191 \hline
  1192 $(t@1,\dots,t@n)$ & $(t@1,(\dots,(t@{n-1},t@n)\dots)$ \\
  1193 \end{tabular}
  1194 \end{center}
  1195 In addition, it is possible to use tuples
  1196 as patterns in abstractions:
  1197 \begin{center}
  1198 {\tt\%($x$,$y$). $t$} \quad stands for\quad \texttt{split(\%$x$\thinspace$y$.\ $t$)} 
  1199 \end{center}
  1200 Nested patterns are also supported.  They are translated stepwise:
  1201 \begin{eqnarray*}
  1202 \hbox{\tt\%($x$,$y$,$z$).\ $t$} 
  1203    & \leadsto & \hbox{\tt\%($x$,($y$,$z$)).\ $t$} \\
  1204    & \leadsto & \hbox{\tt split(\%$x$.\%($y$,$z$).\ $t$)}\\
  1205    & \leadsto & \hbox{\tt split(\%$x$.\ split(\%$y$ $z$.\ $t$))}
  1206 \end{eqnarray*}
  1207 The reverse translation is performed upon printing.
  1208 \begin{warn}
  1209   The translation between patterns and \texttt{split} is performed automatically
  1210   by the parser and printer.  Thus the internal and external form of a term
  1211   may differ, which can affects proofs.  For example the term {\tt
  1212   (\%(x,y).(y,x))(a,b)} requires the theorem \texttt{split} (which is in the
  1213   default simpset) to rewrite to {\tt(b,a)}.
  1214 \end{warn}
  1215 In addition to explicit $\lambda$-abstractions, patterns can be used in any
  1216 variable binding construct which is internally described by a
  1217 $\lambda$-abstraction.  Some important examples are
  1218 \begin{description}
  1219 \item[Let:] \texttt{let {\it pattern} = $t$ in $u$}
  1220 \item[Quantifiers:] \texttt{ALL~{\it pattern}:$A$.~$P$}
  1221 \item[Choice:] {\underscoreon \tt SOME~{\it pattern}.~$P$}
  1222 \item[Set operations:] \texttt{UN~{\it pattern}:$A$.~$B$}
  1223 \item[Sets:] \texttt{{\ttlbrace}{\it pattern}.~$P${\ttrbrace}}
  1224 \end{description}
  1225 
  1226 There is a simple tactic which supports reasoning about patterns:
  1227 \begin{ttdescription}
  1228 \item[\ttindexbold{split_all_tac} $i$] replaces in subgoal $i$ all
  1229   {\tt!!}-quantified variables of product type by individual variables for
  1230   each component.  A simple example:
  1231 \begin{ttbox}
  1232 {\out 1. !!p. (\%(x,y,z). (x, y, z)) p = p}
  1233 by(split_all_tac 1);
  1234 {\out 1. !!x xa ya. (\%(x,y,z). (x, y, z)) (x, xa, ya) = (x, xa, ya)}
  1235 \end{ttbox}
  1236 \end{ttdescription}
  1237 
  1238 Theory \texttt{Prod} also introduces the degenerate product type \texttt{unit}
  1239 which contains only a single element named {\tt()} with the property
  1240 \begin{ttbox}
  1241 \tdx{unit_eq}       u = ()
  1242 \end{ttbox}
  1243 \bigskip
  1244 
  1245 Theory \thydx{Sum} (Fig.~\ref{hol-sum}) defines the sum type $\alpha+\beta$
  1246 which associates to the right and has a lower priority than $*$: $\tau@1 +
  1247 \tau@2 + \tau@3*\tau@4$ means $\tau@1 + (\tau@2 + (\tau@3*\tau@4))$.
  1248 
  1249 The definition of products and sums in terms of existing types is not
  1250 shown.  The constructions are fairly standard and can be found in the
  1251 respective theory files. Although the sum and product types are
  1252 constructed manually for foundational reasons, they are represented as
  1253 actual datatypes later (see {\S}\ref{subsec:datatype:rep_datatype}).
  1254 Therefore, the theory \texttt{Datatype} should be used instead of
  1255 \texttt{Sum} or \texttt{Prod}.
  1256 
  1257 \begin{figure}
  1258 \begin{constants}
  1259   \it symbol    & \it meta-type &           & \it description \\ 
  1260   \cdx{Inl}     & $\alpha \To \alpha+\beta$    & & first injection\\
  1261   \cdx{Inr}     & $\beta \To \alpha+\beta$     & & second injection\\
  1262   \cdx{sum_case} & $[\alpha\To\gamma, \beta\To\gamma, \alpha+\beta] \To\gamma$
  1263         & & conditional
  1264 \end{constants}
  1265 \begin{ttbox}\makeatletter
  1266 \tdx{Inl_not_Inr}    Inl a ~= Inr b
  1267 
  1268 \tdx{inj_Inl}        inj Inl
  1269 \tdx{inj_Inr}        inj Inr
  1270 
  1271 \tdx{sumE}           [| !!x. P(Inl x);  !!y. P(Inr y) |] ==> P s
  1272 
  1273 \tdx{sum_case_Inl}   sum_case f g (Inl x) = f x
  1274 \tdx{sum_case_Inr}   sum_case f g (Inr x) = g x
  1275 
  1276 \tdx{surjective_sum} sum_case (\%x. f(Inl x)) (\%y. f(Inr y)) s = f s
  1277 \tdx{sum.split_case} R(sum_case f g s) = ((! x. s = Inl(x) --> R(f(x))) &
  1278                                      (! y. s = Inr(y) --> R(g(y))))
  1279 \end{ttbox}
  1280 \caption{Type $\alpha+\beta$}\label{hol-sum}
  1281 \end{figure}
  1282 
  1283 \begin{figure}
  1284 \index{*"< symbol}
  1285 \index{*"* symbol}
  1286 \index{*div symbol}
  1287 \index{*mod symbol}
  1288 \index{*dvd symbol}
  1289 \index{*"+ symbol}
  1290 \index{*"- symbol}
  1291 \begin{constants}
  1292   \it symbol    & \it meta-type & \it priority & \it description \\ 
  1293   \cdx{0}       & $\alpha$  & & zero \\
  1294   \cdx{Suc}     & $nat \To nat$ & & successor function\\
  1295   \tt *    & $[\alpha,\alpha]\To \alpha$    &  Left 70 & multiplication \\
  1296   \tt div  & $[\alpha,\alpha]\To \alpha$    &  Left 70 & division\\
  1297   \tt mod  & $[\alpha,\alpha]\To \alpha$    &  Left 70 & modulus\\
  1298   \tt dvd  & $[\alpha,\alpha]\To bool$     &  Left 70 & ``divides'' relation\\
  1299   \tt +    & $[\alpha,\alpha]\To \alpha$    &  Left 65 & addition\\
  1300   \tt -    & $[\alpha,\alpha]\To \alpha$    &  Left 65 & subtraction
  1301 \end{constants}
  1302 \subcaption{Constants and infixes}
  1303 
  1304 \begin{ttbox}\makeatother
  1305 \tdx{nat_induct}     [| P 0; !!n. P n ==> P(Suc n) |]  ==> P n
  1306 
  1307 \tdx{Suc_not_Zero}   Suc m ~= 0
  1308 \tdx{inj_Suc}        inj Suc
  1309 \tdx{n_not_Suc_n}    n~=Suc n
  1310 \subcaption{Basic properties}
  1311 \end{ttbox}
  1312 \caption{The type of natural numbers, \tydx{nat}} \label{hol-nat1}
  1313 \end{figure}
  1314 
  1315 
  1316 \begin{figure}
  1317 \begin{ttbox}\makeatother
  1318               0+n           = n
  1319               (Suc m)+n     = Suc(m+n)
  1320 
  1321               m-0           = m
  1322               0-n           = n
  1323               Suc(m)-Suc(n) = m-n
  1324 
  1325               0*n           = 0
  1326               Suc(m)*n      = n + m*n
  1327 
  1328 \tdx{mod_less}      m<n ==> m mod n = m
  1329 \tdx{mod_geq}       [| 0<n;  ~m<n |] ==> m mod n = (m-n) mod n
  1330 
  1331 \tdx{div_less}      m<n ==> m div n = 0
  1332 \tdx{div_geq}       [| 0<n;  ~m<n |] ==> m div n = Suc((m-n) div n)
  1333 \end{ttbox}
  1334 \caption{Recursion equations for the arithmetic operators} \label{hol-nat2}
  1335 \end{figure}
  1336 
  1337 \subsection{The type of natural numbers, \textit{nat}}
  1338 \index{nat@{\textit{nat}} type|(}
  1339 
  1340 The theory \thydx{Nat} defines the natural numbers in a roundabout but
  1341 traditional way.  The axiom of infinity postulates a type~\tydx{ind} of
  1342 individuals, which is non-empty and closed under an injective operation.  The
  1343 natural numbers are inductively generated by choosing an arbitrary individual
  1344 for~0 and using the injective operation to take successors.  This is a least
  1345 fixedpoint construction.  
  1346 
  1347 Type~\tydx{nat} is an instance of class~\cldx{ord}, which makes the overloaded
  1348 functions of this class (especially \cdx{<} and \cdx{<=}, but also \cdx{min},
  1349 \cdx{max} and \cdx{LEAST}) available on \tydx{nat}.  Theory \thydx{Nat} 
  1350 also shows that {\tt<=} is a linear order, so \tydx{nat} is
  1351 also an instance of class \cldx{linorder}.
  1352 
  1353 Theory \thydx{NatArith} develops arithmetic on the natural numbers.  It defines
  1354 addition, multiplication and subtraction.  Theory \thydx{Divides} defines
  1355 division, remainder and the ``divides'' relation.  The numerous theorems
  1356 proved include commutative, associative, distributive, identity and
  1357 cancellation laws.  See Figs.\ts\ref{hol-nat1} and~\ref{hol-nat2}.  The
  1358 recursion equations for the operators \texttt{+}, \texttt{-} and \texttt{*} on
  1359 \texttt{nat} are part of the default simpset.
  1360 
  1361 Functions on \tydx{nat} can be defined by primitive or well-founded recursion;
  1362 see {\S}\ref{sec:HOL:recursive}.  A simple example is addition.
  1363 Here, \texttt{op +} is the name of the infix operator~\texttt{+}, following
  1364 the standard convention.
  1365 \begin{ttbox}
  1366 \sdx{primrec}
  1367       "0 + n = n"
  1368   "Suc m + n = Suc (m + n)"
  1369 \end{ttbox}
  1370 There is also a \sdx{case}-construct
  1371 of the form
  1372 \begin{ttbox}
  1373 case \(e\) of 0 => \(a\) | Suc \(m\) => \(b\)
  1374 \end{ttbox}
  1375 Note that Isabelle insists on precisely this format; you may not even change
  1376 the order of the two cases.
  1377 Both \texttt{primrec} and \texttt{case} are realized by a recursion operator
  1378 \cdx{nat_rec}, which is available because \textit{nat} is represented as
  1379 a datatype (see {\S}\ref{subsec:datatype:rep_datatype}).
  1380 
  1381 %The predecessor relation, \cdx{pred_nat}, is shown to be well-founded.
  1382 %Recursion along this relation resembles primitive recursion, but is
  1383 %stronger because we are in higher-order logic; using primitive recursion to
  1384 %define a higher-order function, we can easily Ackermann's function, which
  1385 %is not primitive recursive \cite[page~104]{thompson91}.
  1386 %The transitive closure of \cdx{pred_nat} is~$<$.  Many functions on the
  1387 %natural numbers are most easily expressed using recursion along~$<$.
  1388 
  1389 Tactic {\tt\ttindex{induct_tac} "$n$" $i$} performs induction on variable~$n$
  1390 in subgoal~$i$ using theorem \texttt{nat_induct}.  There is also the derived
  1391 theorem \tdx{less_induct}:
  1392 \begin{ttbox}
  1393 [| !!n. [| ! m. m<n --> P m |] ==> P n |]  ==>  P n
  1394 \end{ttbox}
  1395 
  1396 
  1397 \subsection{Numerical types and numerical reasoning}
  1398 
  1399 The integers (type \tdx{int}) are also available in HOL, and the reals (type
  1400 \tdx{real}) are available in the logic image \texttt{HOL-Complex}.  They support
  1401 the expected operations of addition (\texttt{+}), subtraction (\texttt{-}) and
  1402 multiplication (\texttt{*}), and much else.  Type \tdx{int} provides the
  1403 \texttt{div} and \texttt{mod} operators, while type \tdx{real} provides real
  1404 division and other operations.  Both types belong to class \cldx{linorder}, so
  1405 they inherit the relational operators and all the usual properties of linear
  1406 orderings.  For full details, please survey the theories in subdirectories
  1407 \texttt{Integ}, \texttt{Real}, and \texttt{Complex}.
  1408 
  1409 All three numeric types admit numerals of the form \texttt{$sd\ldots d$},
  1410 where $s$ is an optional minus sign and $d\ldots d$ is a string of digits.
  1411 Numerals are represented internally by a datatype for binary notation, which
  1412 allows numerical calculations to be performed by rewriting.  For example, the
  1413 integer division of \texttt{54342339} by \texttt{3452} takes about five
  1414 seconds.  By default, the simplifier cancels like terms on the opposite sites
  1415 of relational operators (reducing \texttt{z+x<x+y} to \texttt{z<y}, for
  1416 instance.  The simplifier also collects like terms, replacing \texttt{x+y+x*3}
  1417 by \texttt{4*x+y}.
  1418 
  1419 Sometimes numerals are not wanted, because for example \texttt{n+3} does not
  1420 match a pattern of the form \texttt{Suc $k$}.  You can re-arrange the form of
  1421 an arithmetic expression by proving (via \texttt{subgoal_tac}) a lemma such as
  1422 \texttt{n+3 = Suc (Suc (Suc n))}.  As an alternative, you can disable the
  1423 fancier simplifications by using a basic simpset such as \texttt{HOL_ss}
  1424 rather than the default one, \texttt{simpset()}.
  1425 
  1426 Reasoning about arithmetic inequalities can be tedious.  Fortunately, HOL
  1427 provides a decision procedure for \emph{linear arithmetic}: formulae involving
  1428 addition and subtraction. The simplifier invokes a weak version of this
  1429 decision procedure automatically. If this is not sufficent, you can invoke the
  1430 full procedure \ttindex{Lin_Arith.tac} explicitly.  It copes with arbitrary
  1431 formulae involving {\tt=}, {\tt<}, {\tt<=}, {\tt+}, {\tt-}, {\tt Suc}, {\tt
  1432   min}, {\tt max} and numerical constants. Other subterms are treated as
  1433 atomic, while subformulae not involving numerical types are ignored. Quantified
  1434 subformulae are ignored unless they are positive universal or negative
  1435 existential. The running time is exponential in the number of
  1436 occurrences of {\tt min}, {\tt max}, and {\tt-} because they require case
  1437 distinctions.
  1438 If {\tt k} is a numeral, then {\tt div k}, {\tt mod k} and
  1439 {\tt k dvd} are also supported. The former two are eliminated
  1440 by case distinctions, again blowing up the running time.
  1441 If the formula involves explicit quantifiers, \texttt{Lin_Arith.tac} may take
  1442 super-exponential time and space.
  1443 
  1444 If \texttt{Lin_Arith.tac} fails, try to find relevant arithmetic results in
  1445 the library.  The theories \texttt{Nat} and \texttt{NatArith} contain
  1446 theorems about {\tt<}, {\tt<=}, \texttt{+}, \texttt{-} and \texttt{*}.
  1447 Theory \texttt{Divides} contains theorems about \texttt{div} and
  1448 \texttt{mod}.  Use Proof General's \emph{find} button (or other search
  1449 facilities) to locate them.
  1450 
  1451 \index{nat@{\textit{nat}} type|)}
  1452 
  1453 
  1454 \begin{figure}
  1455 \index{#@{\tt[]} symbol}
  1456 \index{#@{\tt\#} symbol}
  1457 \index{"@@{\tt\at} symbol}
  1458 \index{*"! symbol}
  1459 \begin{constants}
  1460   \it symbol & \it meta-type & \it priority & \it description \\
  1461   \tt[]    & $\alpha\,list$ & & empty list\\
  1462   \tt \#   & $[\alpha,\alpha\,list]\To \alpha\,list$ & Right 65 & 
  1463         list constructor \\
  1464   \cdx{null}    & $\alpha\,list \To bool$ & & emptiness test\\
  1465   \cdx{hd}      & $\alpha\,list \To \alpha$ & & head \\
  1466   \cdx{tl}      & $\alpha\,list \To \alpha\,list$ & & tail \\
  1467   \cdx{last}    & $\alpha\,list \To \alpha$ & & last element \\
  1468   \cdx{butlast} & $\alpha\,list \To \alpha\,list$ & & drop last element \\
  1469   \tt\at  & $[\alpha\,list,\alpha\,list]\To \alpha\,list$ & Left 65 & append \\
  1470   \cdx{map}     & $(\alpha\To\beta) \To (\alpha\,list \To \beta\,list)$
  1471         & & apply to all\\
  1472   \cdx{filter}  & $(\alpha \To bool) \To (\alpha\,list \To \alpha\,list)$
  1473         & & filter functional\\
  1474   \cdx{set}& $\alpha\,list \To \alpha\,set$ & & elements\\
  1475   \sdx{mem}  & $\alpha \To \alpha\,list \To bool$  &  Left 55   & membership\\
  1476   \cdx{foldl}   & $(\beta\To\alpha\To\beta) \To \beta \To \alpha\,list \To \beta$ &
  1477   & iteration \\
  1478   \cdx{concat}   & $(\alpha\,list)list\To \alpha\,list$ & & concatenation \\
  1479   \cdx{rev}     & $\alpha\,list \To \alpha\,list$ & & reverse \\
  1480   \cdx{length}  & $\alpha\,list \To nat$ & & length \\
  1481   \tt! & $\alpha\,list \To nat \To \alpha$ & Left 100 & indexing \\
  1482   \cdx{take}, \cdx{drop} & $nat \To \alpha\,list \To \alpha\,list$ &&
  1483     take/drop a prefix \\
  1484   \cdx{takeWhile},\\
  1485   \cdx{dropWhile} &
  1486     $(\alpha \To bool) \To \alpha\,list \To \alpha\,list$ &&
  1487     take/drop a prefix
  1488 \end{constants}
  1489 \subcaption{Constants and infixes}
  1490 
  1491 \begin{center} \tt\frenchspacing
  1492 \begin{tabular}{rrr} 
  1493   \it external        & \it internal  & \it description \\{}
  1494   [$x@1$, $\dots$, $x@n$]  &  $x@1$ \# $\cdots$ \# $x@n$ \# [] &
  1495         \rm finite list \\{}
  1496   [$x$:$l$. $P$]  & filter ($\lambda x{.}P$) $l$ & 
  1497         \rm list comprehension
  1498 \end{tabular}
  1499 \end{center}
  1500 \subcaption{Translations}
  1501 \caption{The theory \thydx{List}} \label{hol-list}
  1502 \end{figure}
  1503 
  1504 
  1505 \begin{figure}
  1506 \begin{ttbox}\makeatother
  1507 null [] = True
  1508 null (x#xs) = False
  1509 
  1510 hd (x#xs) = x
  1511 
  1512 tl (x#xs) = xs
  1513 tl [] = []
  1514 
  1515 [] @ ys = ys
  1516 (x#xs) @ ys = x # xs @ ys
  1517 
  1518 set [] = \ttlbrace\ttrbrace
  1519 set (x#xs) = insert x (set xs)
  1520 
  1521 x mem [] = False
  1522 x mem (y#ys) = (if y=x then True else x mem ys)
  1523 
  1524 concat([]) = []
  1525 concat(x#xs) = x @ concat(xs)
  1526 
  1527 rev([]) = []
  1528 rev(x#xs) = rev(xs) @ [x]
  1529 
  1530 length([]) = 0
  1531 length(x#xs) = Suc(length(xs))
  1532 
  1533 xs!0 = hd xs
  1534 xs!(Suc n) = (tl xs)!n
  1535 \end{ttbox}
  1536 \caption{Simple list processing functions}
  1537 \label{fig:HOL:list-simps}
  1538 \end{figure}
  1539 
  1540 \begin{figure}
  1541 \begin{ttbox}\makeatother
  1542 map f [] = []
  1543 map f (x#xs) = f x # map f xs
  1544 
  1545 filter P [] = []
  1546 filter P (x#xs) = (if P x then x#filter P xs else filter P xs)
  1547 
  1548 foldl f a [] = a
  1549 foldl f a (x#xs) = foldl f (f a x) xs
  1550 
  1551 take n [] = []
  1552 take n (x#xs) = (case n of 0 => [] | Suc(m) => x # take m xs)
  1553 
  1554 drop n [] = []
  1555 drop n (x#xs) = (case n of 0 => x#xs | Suc(m) => drop m xs)
  1556 
  1557 takeWhile P [] = []
  1558 takeWhile P (x#xs) = (if P x then x#takeWhile P xs else [])
  1559 
  1560 dropWhile P [] = []
  1561 dropWhile P (x#xs) = (if P x then dropWhile P xs else xs)
  1562 \end{ttbox}
  1563 \caption{Further list processing functions}
  1564 \label{fig:HOL:list-simps2}
  1565 \end{figure}
  1566 
  1567 
  1568 \subsection{The type constructor for lists, \textit{list}}
  1569 \label{subsec:list}
  1570 \index{list@{\textit{list}} type|(}
  1571 
  1572 Figure~\ref{hol-list} presents the theory \thydx{List}: the basic list
  1573 operations with their types and syntax.  Type $\alpha \; list$ is
  1574 defined as a \texttt{datatype} with the constructors {\tt[]} and {\tt\#}.
  1575 As a result the generic structural induction and case analysis tactics
  1576 \texttt{induct\_tac} and \texttt{cases\_tac} also become available for
  1577 lists.  A \sdx{case} construct of the form
  1578 \begin{center}\tt
  1579 case $e$ of [] => $a$  |  \(x\)\#\(xs\) => b
  1580 \end{center}
  1581 is defined by translation.  For details see~{\S}\ref{sec:HOL:datatype}. There
  1582 is also a case splitting rule \tdx{split_list_case}
  1583 \[
  1584 \begin{array}{l}
  1585 P(\mathtt{case}~e~\mathtt{of}~\texttt{[] =>}~a ~\texttt{|}~
  1586                x\texttt{\#}xs~\texttt{=>}~f~x~xs) ~= \\
  1587 ((e = \texttt{[]} \to P(a)) \land
  1588  (\forall x~ xs. e = x\texttt{\#}xs \to P(f~x~xs)))
  1589 \end{array}
  1590 \]
  1591 which can be fed to \ttindex{addsplits} just like
  1592 \texttt{split_if} (see~{\S}\ref{subsec:HOL:case:splitting}).
  1593 
  1594 \texttt{List} provides a basic library of list processing functions defined by
  1595 primitive recursion (see~{\S}\ref{sec:HOL:primrec}).  The recursion equations
  1596 are shown in Figs.\ts\ref{fig:HOL:list-simps} and~\ref{fig:HOL:list-simps2}.
  1597 
  1598 \index{list@{\textit{list}} type|)}
  1599 
  1600 
  1601 \subsection{Introducing new types} \label{sec:typedef}
  1602 
  1603 The HOL-methodology dictates that all extensions to a theory should be
  1604 \textbf{definitional}.  The type definition mechanism that meets this
  1605 criterion is \ttindex{typedef}.  Note that \emph{type synonyms}, which are
  1606 inherited from Pure and described elsewhere, are just syntactic abbreviations
  1607 that have no logical meaning.
  1608 
  1609 \begin{warn}
  1610   Types in HOL must be non-empty; otherwise the quantifier rules would be
  1611   unsound, because $\exists x. x=x$ is a theorem \cite[{\S}7]{paulson-COLOG}.
  1612 \end{warn}
  1613 A \bfindex{type definition} identifies the new type with a subset of
  1614 an existing type.  More precisely, the new type is defined by
  1615 exhibiting an existing type~$\tau$, a set~$A::\tau\,set$, and a
  1616 theorem of the form $x:A$.  Thus~$A$ is a non-empty subset of~$\tau$,
  1617 and the new type denotes this subset.  New functions are defined that
  1618 establish an isomorphism between the new type and the subset.  If
  1619 type~$\tau$ involves type variables $\alpha@1$, \ldots, $\alpha@n$,
  1620 then the type definition creates a type constructor
  1621 $(\alpha@1,\ldots,\alpha@n)ty$ rather than a particular type.
  1622 
  1623 \begin{figure}[htbp]
  1624 \begin{rail}
  1625 typedef  : 'typedef' ( () | '(' name ')') type '=' set witness;
  1626 
  1627 type    : typevarlist name ( () | '(' infix ')' );
  1628 set     : string;
  1629 witness : () | '(' id ')';
  1630 \end{rail}
  1631 \caption{Syntax of type definitions}
  1632 \label{fig:HOL:typedef}
  1633 \end{figure}
  1634 
  1635 The syntax for type definitions is shown in Fig.~\ref{fig:HOL:typedef}.  For
  1636 the definition of `typevarlist' and `infix' see
  1637 \iflabelundefined{chap:classical}
  1638 {the appendix of the {\em Reference Manual\/}}%
  1639 {Appendix~\ref{app:TheorySyntax}}.  The remaining nonterminals have the
  1640 following meaning:
  1641 \begin{description}
  1642 \item[\it type:] the new type constructor $(\alpha@1,\dots,\alpha@n)ty$ with
  1643   optional infix annotation.
  1644 \item[\it name:] an alphanumeric name $T$ for the type constructor
  1645   $ty$, in case $ty$ is a symbolic name.  Defaults to $ty$.
  1646 \item[\it set:] the representing subset $A$.
  1647 \item[\it witness:] name of a theorem of the form $a:A$ proving
  1648   non-emptiness.  It can be omitted in case Isabelle manages to prove
  1649   non-emptiness automatically.
  1650 \end{description}
  1651 If all context conditions are met (no duplicate type variables in
  1652 `typevarlist', no extra type variables in `set', and no free term variables
  1653 in `set'), the following components are added to the theory:
  1654 \begin{itemize}
  1655 \item a type $ty :: (term,\dots,term)term$
  1656 \item constants
  1657 \begin{eqnarray*}
  1658 T &::& \tau\;set \\
  1659 Rep_T &::& (\alpha@1,\dots,\alpha@n)ty \To \tau \\
  1660 Abs_T &::& \tau \To (\alpha@1,\dots,\alpha@n)ty
  1661 \end{eqnarray*}
  1662 \item a definition and three axioms
  1663 \[
  1664 \begin{array}{ll}
  1665 T{\tt_def} & T \equiv A \\
  1666 {\tt Rep_}T & Rep_T\,x \in T \\
  1667 {\tt Rep_}T{\tt_inverse} & Abs_T\,(Rep_T\,x) = x \\
  1668 {\tt Abs_}T{\tt_inverse} & y \in T \Imp Rep_T\,(Abs_T\,y) = y
  1669 \end{array}
  1670 \]
  1671 stating that $(\alpha@1,\dots,\alpha@n)ty$ is isomorphic to $A$ by $Rep_T$
  1672 and its inverse $Abs_T$.
  1673 \end{itemize}
  1674 Below are two simple examples of HOL type definitions.  Non-emptiness is
  1675 proved automatically here.
  1676 \begin{ttbox}
  1677 typedef unit = "{\ttlbrace}True{\ttrbrace}"
  1678 
  1679 typedef (prod)
  1680   ('a, 'b) "*"    (infixr 20)
  1681       = "{\ttlbrace}f . EX (a::'a) (b::'b). f = (\%x y. x = a & y = b){\ttrbrace}"
  1682 \end{ttbox}
  1683 
  1684 Type definitions permit the introduction of abstract data types in a safe
  1685 way, namely by providing models based on already existing types.  Given some
  1686 abstract axiomatic description $P$ of a type, this involves two steps:
  1687 \begin{enumerate}
  1688 \item Find an appropriate type $\tau$ and subset $A$ which has the desired
  1689   properties $P$, and make a type definition based on this representation.
  1690 \item Prove that $P$ holds for $ty$ by lifting $P$ from the representation.
  1691 \end{enumerate}
  1692 You can now forget about the representation and work solely in terms of the
  1693 abstract properties $P$.
  1694 
  1695 \begin{warn}
  1696 If you introduce a new type (constructor) $ty$ axiomatically, i.e.\ by
  1697 declaring the type and its operations and by stating the desired axioms, you
  1698 should make sure the type has a non-empty model.  You must also have a clause
  1699 \par
  1700 \begin{ttbox}
  1701 arities \(ty\) :: (term,\thinspace\(\dots\),{\thinspace}term){\thinspace}term
  1702 \end{ttbox}
  1703 in your theory file to tell Isabelle that $ty$ is in class \texttt{term}, the
  1704 class of all HOL types.
  1705 \end{warn}
  1706 
  1707 
  1708 \section{Datatype definitions}
  1709 \label{sec:HOL:datatype}
  1710 \index{*datatype|(}
  1711 
  1712 Inductive datatypes, similar to those of \ML, frequently appear in
  1713 applications of Isabelle/HOL.  In principle, such types could be defined by
  1714 hand via \texttt{typedef} (see {\S}\ref{sec:typedef}), but this would be far too
  1715 tedious.  The \ttindex{datatype} definition package of Isabelle/HOL (cf.\ 
  1716 \cite{Berghofer-Wenzel:1999:TPHOL}) automates such chores.  It generates an
  1717 appropriate \texttt{typedef} based on a least fixed-point construction, and
  1718 proves freeness theorems and induction rules, as well as theorems for
  1719 recursion and case combinators.  The user just has to give a simple
  1720 specification of new inductive types using a notation similar to {\ML} or
  1721 Haskell.
  1722 
  1723 The current datatype package can handle both mutual and indirect recursion.
  1724 It also offers to represent existing types as datatypes giving the advantage
  1725 of a more uniform view on standard theories.
  1726 
  1727 
  1728 \subsection{Basics}
  1729 \label{subsec:datatype:basics}
  1730 
  1731 A general \texttt{datatype} definition is of the following form:
  1732 \[
  1733 \begin{array}{llcl}
  1734 \mathtt{datatype} & (\vec{\alpha})t@1 & = &
  1735   C^1@1~\tau^1@{1,1}~\ldots~\tau^1@{1,m^1@1} ~\mid~ \ldots ~\mid~
  1736     C^1@{k@1}~\tau^1@{k@1,1}~\ldots~\tau^1@{k@1,m^1@{k@1}} \\
  1737  & & \vdots \\
  1738 \mathtt{and} & (\vec{\alpha})t@n & = &
  1739   C^n@1~\tau^n@{1,1}~\ldots~\tau^n@{1,m^n@1} ~\mid~ \ldots ~\mid~
  1740     C^n@{k@n}~\tau^n@{k@n,1}~\ldots~\tau^n@{k@n,m^n@{k@n}}
  1741 \end{array}
  1742 \]
  1743 where $\vec{\alpha} = (\alpha@1,\ldots,\alpha@h)$ is a list of type variables,
  1744 $C^j@i$ are distinct constructor names and $\tau^j@{i,i'}$ are {\em
  1745   admissible} types containing at most the type variables $\alpha@1, \ldots,
  1746 \alpha@h$. A type $\tau$ occurring in a \texttt{datatype} definition is {\em
  1747   admissible} if and only if
  1748 \begin{itemize}
  1749 \item $\tau$ is non-recursive, i.e.\ $\tau$ does not contain any of the
  1750 newly defined type constructors $t@1,\ldots,t@n$, or
  1751 \item $\tau = (\vec{\alpha})t@{j'}$ where $1 \leq j' \leq n$, or
  1752 \item $\tau = (\tau'@1,\ldots,\tau'@{h'})t'$, where $t'$ is
  1753 the type constructor of an already existing datatype and $\tau'@1,\ldots,\tau'@{h'}$
  1754 are admissible types.
  1755 \item $\tau = \sigma \to \tau'$, where $\tau'$ is an admissible
  1756 type and $\sigma$ is non-recursive (i.e. the occurrences of the newly defined
  1757 types are {\em strictly positive})
  1758 \end{itemize}
  1759 If some $(\vec{\alpha})t@{j'}$ occurs in a type $\tau^j@{i,i'}$
  1760 of the form
  1761 \[
  1762 (\ldots,\ldots ~ (\vec{\alpha})t@{j'} ~ \ldots,\ldots)t'
  1763 \]
  1764 this is called a {\em nested} (or \emph{indirect}) occurrence. A very simple
  1765 example of a datatype is the type \texttt{list}, which can be defined by
  1766 \begin{ttbox}
  1767 datatype 'a list = Nil
  1768                  | Cons 'a ('a list)
  1769 \end{ttbox}
  1770 Arithmetic expressions \texttt{aexp} and boolean expressions \texttt{bexp} can be modelled
  1771 by the mutually recursive datatype definition
  1772 \begin{ttbox}
  1773 datatype 'a aexp = If_then_else ('a bexp) ('a aexp) ('a aexp)
  1774                  | Sum ('a aexp) ('a aexp)
  1775                  | Diff ('a aexp) ('a aexp)
  1776                  | Var 'a
  1777                  | Num nat
  1778 and      'a bexp = Less ('a aexp) ('a aexp)
  1779                  | And ('a bexp) ('a bexp)
  1780                  | Or ('a bexp) ('a bexp)
  1781 \end{ttbox}
  1782 The datatype \texttt{term}, which is defined by
  1783 \begin{ttbox}
  1784 datatype ('a, 'b) term = Var 'a
  1785                        | App 'b ((('a, 'b) term) list)
  1786 \end{ttbox}
  1787 is an example for a datatype with nested recursion. Using nested recursion
  1788 involving function spaces, we may also define infinitely branching datatypes, e.g.
  1789 \begin{ttbox}
  1790 datatype 'a tree = Atom 'a | Branch "nat => 'a tree"
  1791 \end{ttbox}
  1792 
  1793 \medskip
  1794 
  1795 Types in HOL must be non-empty. Each of the new datatypes
  1796 $(\vec{\alpha})t@j$ with $1 \leq j \leq n$ is non-empty if and only if it has a
  1797 constructor $C^j@i$ with the following property: for all argument types
  1798 $\tau^j@{i,i'}$ of the form $(\vec{\alpha})t@{j'}$ the datatype
  1799 $(\vec{\alpha})t@{j'}$ is non-empty.
  1800 
  1801 If there are no nested occurrences of the newly defined datatypes, obviously
  1802 at least one of the newly defined datatypes $(\vec{\alpha})t@j$
  1803 must have a constructor $C^j@i$ without recursive arguments, a \emph{base
  1804   case}, to ensure that the new types are non-empty. If there are nested
  1805 occurrences, a datatype can even be non-empty without having a base case
  1806 itself. Since \texttt{list} is a non-empty datatype, \texttt{datatype t = C (t
  1807   list)} is non-empty as well.
  1808 
  1809 
  1810 \subsubsection{Freeness of the constructors}
  1811 
  1812 The datatype constructors are automatically defined as functions of their
  1813 respective type:
  1814 \[ C^j@i :: [\tau^j@{i,1},\dots,\tau^j@{i,m^j@i}] \To (\alpha@1,\dots,\alpha@h)t@j \]
  1815 These functions have certain {\em freeness} properties.  They construct
  1816 distinct values:
  1817 \[
  1818 C^j@i~x@1~\dots~x@{m^j@i} \neq C^j@{i'}~y@1~\dots~y@{m^j@{i'}} \qquad
  1819 \mbox{for all}~ i \neq i'.
  1820 \]
  1821 The constructor functions are injective:
  1822 \[
  1823 (C^j@i~x@1~\dots~x@{m^j@i} = C^j@i~y@1~\dots~y@{m^j@i}) =
  1824 (x@1 = y@1 \land \dots \land x@{m^j@i} = y@{m^j@i})
  1825 \]
  1826 Since the number of distinctness inequalities is quadratic in the number of
  1827 constructors, the datatype package avoids proving them separately if there are
  1828 too many constructors. Instead, specific inequalities are proved by a suitable
  1829 simplification procedure on demand.\footnote{This procedure, which is already part
  1830 of the default simpset, may be referred to by the ML identifier
  1831 \texttt{DatatypePackage.distinct_simproc}.}
  1832 
  1833 \subsubsection{Structural induction}
  1834 
  1835 The datatype package also provides structural induction rules.  For
  1836 datatypes without nested recursion, this is of the following form:
  1837 \[
  1838 \infer{P@1~x@1 \land \dots \land P@n~x@n}
  1839   {\begin{array}{lcl}
  1840      \Forall x@1 \dots x@{m^1@1}.
  1841        \List{P@{s^1@{1,1}}~x@{r^1@{1,1}}; \dots;
  1842          P@{s^1@{1,l^1@1}}~x@{r^1@{1,l^1@1}}} & \Imp &
  1843            P@1~\left(C^1@1~x@1~\dots~x@{m^1@1}\right) \\
  1844      & \vdots \\
  1845      \Forall x@1 \dots x@{m^1@{k@1}}.
  1846        \List{P@{s^1@{k@1,1}}~x@{r^1@{k@1,1}}; \dots;
  1847          P@{s^1@{k@1,l^1@{k@1}}}~x@{r^1@{k@1,l^1@{k@1}}}} & \Imp &
  1848            P@1~\left(C^1@{k@1}~x@1~\ldots~x@{m^1@{k@1}}\right) \\
  1849      & \vdots \\
  1850      \Forall x@1 \dots x@{m^n@1}.
  1851        \List{P@{s^n@{1,1}}~x@{r^n@{1,1}}; \dots;
  1852          P@{s^n@{1,l^n@1}}~x@{r^n@{1,l^n@1}}} & \Imp &
  1853            P@n~\left(C^n@1~x@1~\ldots~x@{m^n@1}\right) \\
  1854      & \vdots \\
  1855      \Forall x@1 \dots x@{m^n@{k@n}}.
  1856        \List{P@{s^n@{k@n,1}}~x@{r^n@{k@n,1}}; \dots
  1857          P@{s^n@{k@n,l^n@{k@n}}}~x@{r^n@{k@n,l^n@{k@n}}}} & \Imp &
  1858            P@n~\left(C^n@{k@n}~x@1~\ldots~x@{m^n@{k@n}}\right)
  1859    \end{array}}
  1860 \]
  1861 where
  1862 \[
  1863 \begin{array}{rcl}
  1864 Rec^j@i & := &
  1865    \left\{\left(r^j@{i,1},s^j@{i,1}\right),\ldots,
  1866      \left(r^j@{i,l^j@i},s^j@{i,l^j@i}\right)\right\} = \\[2ex]
  1867 && \left\{(i',i'')~\left|~
  1868      1\leq i' \leq m^j@i \land 1 \leq i'' \leq n \land
  1869        \tau^j@{i,i'} = (\alpha@1,\ldots,\alpha@h)t@{i''}\right.\right\}
  1870 \end{array}
  1871 \]
  1872 i.e.\ the properties $P@j$ can be assumed for all recursive arguments.
  1873 
  1874 For datatypes with nested recursion, such as the \texttt{term} example from
  1875 above, things are a bit more complicated.  Conceptually, Isabelle/HOL unfolds
  1876 a definition like
  1877 \begin{ttbox}
  1878 datatype ('a,'b) term = Var 'a
  1879                       | App 'b ((('a, 'b) term) list)
  1880 \end{ttbox}
  1881 to an equivalent definition without nesting:
  1882 \begin{ttbox}
  1883 datatype ('a,'b) term      = Var
  1884                            | App 'b (('a, 'b) term_list)
  1885 and      ('a,'b) term_list = Nil'
  1886                            | Cons' (('a,'b) term) (('a,'b) term_list)
  1887 \end{ttbox}
  1888 Note however, that the type \texttt{('a,'b) term_list} and the constructors {\tt
  1889   Nil'} and \texttt{Cons'} are not really introduced.  One can directly work with
  1890 the original (isomorphic) type \texttt{(('a, 'b) term) list} and its existing
  1891 constructors \texttt{Nil} and \texttt{Cons}. Thus, the structural induction rule for
  1892 \texttt{term} gets the form
  1893 \[
  1894 \infer{P@1~x@1 \land P@2~x@2}
  1895   {\begin{array}{l}
  1896      \Forall x.~P@1~(\mathtt{Var}~x) \\
  1897      \Forall x@1~x@2.~P@2~x@2 \Imp P@1~(\mathtt{App}~x@1~x@2) \\
  1898      P@2~\mathtt{Nil} \\
  1899      \Forall x@1~x@2. \List{P@1~x@1; P@2~x@2} \Imp P@2~(\mathtt{Cons}~x@1~x@2)
  1900    \end{array}}
  1901 \]
  1902 Note that there are two predicates $P@1$ and $P@2$, one for the type \texttt{('a,'b) term}
  1903 and one for the type \texttt{(('a, 'b) term) list}.
  1904 
  1905 For a datatype with function types such as \texttt{'a tree}, the induction rule
  1906 is of the form
  1907 \[
  1908 \infer{P~t}
  1909   {\Forall a.~P~(\mathtt{Atom}~a) &
  1910    \Forall ts.~(\forall x.~P~(ts~x)) \Imp P~(\mathtt{Branch}~ts)}
  1911 \]
  1912 
  1913 \medskip In principle, inductive types are already fully determined by
  1914 freeness and structural induction.  For convenience in applications,
  1915 the following derived constructions are automatically provided for any
  1916 datatype.
  1917 
  1918 \subsubsection{The \sdx{case} construct}
  1919 
  1920 The type comes with an \ML-like \texttt{case}-construct:
  1921 \[
  1922 \begin{array}{rrcl}
  1923 \mbox{\tt case}~e~\mbox{\tt of} & C^j@1~x@{1,1}~\dots~x@{1,m^j@1} & \To & e@1 \\
  1924                            \vdots \\
  1925                            \mid & C^j@{k@j}~x@{k@j,1}~\dots~x@{k@j,m^j@{k@j}} & \To & e@{k@j}
  1926 \end{array}
  1927 \]
  1928 where the $x@{i,j}$ are either identifiers or nested tuple patterns as in
  1929 {\S}\ref{subsec:prod-sum}.
  1930 \begin{warn}
  1931   All constructors must be present, their order is fixed, and nested patterns
  1932   are not supported (with the exception of tuples).  Violating this
  1933   restriction results in strange error messages.
  1934 \end{warn}
  1935 
  1936 To perform case distinction on a goal containing a \texttt{case}-construct,
  1937 the theorem $t@j.$\texttt{split} is provided:
  1938 \[
  1939 \begin{array}{@{}rcl@{}}
  1940 P(t@j_\mathtt{case}~f@1~\dots~f@{k@j}~e) &\!\!\!=&
  1941 \!\!\! ((\forall x@1 \dots x@{m^j@1}. e = C^j@1~x@1\dots x@{m^j@1} \to
  1942                              P(f@1~x@1\dots x@{m^j@1})) \\
  1943 &&\!\!\! ~\land~ \dots ~\land \\
  1944 &&~\!\!\! (\forall x@1 \dots x@{m^j@{k@j}}. e = C^j@{k@j}~x@1\dots x@{m^j@{k@j}} \to
  1945                              P(f@{k@j}~x@1\dots x@{m^j@{k@j}})))
  1946 \end{array}
  1947 \]
  1948 where $t@j$\texttt{_case} is the internal name of the \texttt{case}-construct.
  1949 This theorem can be added to a simpset via \ttindex{addsplits}
  1950 (see~{\S}\ref{subsec:HOL:case:splitting}).
  1951 
  1952 Case splitting on assumption works as well, by using the rule
  1953 $t@j.$\texttt{split_asm} in the same manner.  Both rules are available under
  1954 $t@j.$\texttt{splits} (this name is \emph{not} bound in ML, though).
  1955 
  1956 \begin{warn}\index{simplification!of \texttt{case}}%
  1957   By default only the selector expression ($e$ above) in a
  1958   \texttt{case}-construct is simplified, in analogy with \texttt{if} (see
  1959   page~\pageref{if-simp}). Only if that reduces to a constructor is one of
  1960   the arms of the \texttt{case}-construct exposed and simplified. To ensure
  1961   full simplification of all parts of a \texttt{case}-construct for datatype
  1962   $t$, remove $t$\texttt{.}\ttindexbold{case_weak_cong} from the simpset, for
  1963   example by \texttt{delcongs [thm "$t$.weak_case_cong"]}.
  1964 \end{warn}
  1965 
  1966 \subsubsection{The function \cdx{size}}\label{sec:HOL:size}
  1967 
  1968 Theory \texttt{NatArith} declares a generic function \texttt{size} of type
  1969 $\alpha\To nat$.  Each datatype defines a particular instance of \texttt{size}
  1970 by overloading according to the following scheme:
  1971 %%% FIXME: This formula is too big and is completely unreadable
  1972 \[
  1973 size(C^j@i~x@1~\dots~x@{m^j@i}) = \!
  1974 \left\{
  1975 \begin{array}{ll}
  1976 0 & \!\mbox{if $Rec^j@i = \emptyset$} \\
  1977 1+\sum\limits@{h=1}^{l^j@i}size~x@{r^j@{i,h}} &
  1978  \!\mbox{if $Rec^j@i = \left\{\left(r^j@{i,1},s^j@{i,1}\right),\ldots,
  1979   \left(r^j@{i,l^j@i},s^j@{i,l^j@i}\right)\right\}$}
  1980 \end{array}
  1981 \right.
  1982 \]
  1983 where $Rec^j@i$ is defined above.  Viewing datatypes as generalised trees, the
  1984 size of a leaf is 0 and the size of a node is the sum of the sizes of its
  1985 subtrees ${}+1$.
  1986 
  1987 \subsection{Defining datatypes}
  1988 
  1989 The theory syntax for datatype definitions is shown in
  1990 Fig.~\ref{datatype-grammar}.  In order to be well-formed, a datatype
  1991 definition has to obey the rules stated in the previous section.  As a result
  1992 the theory is extended with the new types, the constructors, and the theorems
  1993 listed in the previous section.
  1994 
  1995 \begin{figure}
  1996 \begin{rail}
  1997 datatype : 'datatype' typedecls;
  1998 
  1999 typedecls: ( newtype '=' (cons + '|') ) + 'and'
  2000          ;
  2001 newtype  : typevarlist id ( () | '(' infix ')' )
  2002          ;
  2003 cons     : name (argtype *) ( () | ( '(' mixfix ')' ) )
  2004          ;
  2005 argtype  : id | tid | ('(' typevarlist id ')')
  2006          ;
  2007 \end{rail}
  2008 \caption{Syntax of datatype declarations}
  2009 \label{datatype-grammar}
  2010 \end{figure}
  2011 
  2012 Most of the theorems about datatypes become part of the default simpset and
  2013 you never need to see them again because the simplifier applies them
  2014 automatically.  Only induction or case distinction are usually invoked by hand.
  2015 \begin{ttdescription}
  2016 \item[\ttindexbold{induct_tac} {\tt"}$x${\tt"} $i$]
  2017  applies structural induction on variable $x$ to subgoal $i$, provided the
  2018  type of $x$ is a datatype.
  2019 \item[\texttt{induct_tac}
  2020   {\tt"}$x@1$ $\ldots$ $x@n${\tt"} $i$] applies simultaneous
  2021   structural induction on the variables $x@1,\ldots,x@n$ to subgoal $i$.  This
  2022   is the canonical way to prove properties of mutually recursive datatypes
  2023   such as \texttt{aexp} and \texttt{bexp}, or datatypes with nested recursion such as
  2024   \texttt{term}.
  2025 \end{ttdescription}
  2026 In some cases, induction is overkill and a case distinction over all
  2027 constructors of the datatype suffices.
  2028 \begin{ttdescription}
  2029 \item[\ttindexbold{case_tac} {\tt"}$u${\tt"} $i$]
  2030  performs a case analysis for the term $u$ whose type  must be a datatype.
  2031  If the datatype has $k@j$ constructors  $C^j@1$, \dots $C^j@{k@j}$, subgoal
  2032  $i$ is replaced by $k@j$ new subgoals which  contain the additional
  2033  assumption $u = C^j@{i'}~x@1~\dots~x@{m^j@{i'}}$ for  $i'=1$, $\dots$,~$k@j$.
  2034 \end{ttdescription}
  2035 
  2036 Note that induction is only allowed on free variables that should not occur
  2037 among the premises of the subgoal. Case distinction applies to arbitrary terms.
  2038 
  2039 \bigskip
  2040 
  2041 
  2042 For the technically minded, we exhibit some more details.  Processing the
  2043 theory file produces an \ML\ structure which, in addition to the usual
  2044 components, contains a structure named $t$ for each datatype $t$ defined in
  2045 the file.  Each structure $t$ contains the following elements:
  2046 \begin{ttbox}
  2047 val distinct : thm list
  2048 val inject : thm list
  2049 val induct : thm
  2050 val exhaust : thm
  2051 val cases : thm list
  2052 val split : thm
  2053 val split_asm : thm
  2054 val recs : thm list
  2055 val size : thm list
  2056 val simps : thm list
  2057 \end{ttbox}
  2058 \texttt{distinct}, \texttt{inject}, \texttt{induct}, \texttt{size}
  2059 and \texttt{split} contain the theorems
  2060 described above.  For user convenience, \texttt{distinct} contains
  2061 inequalities in both directions.  The reduction rules of the {\tt
  2062   case}-construct are in \texttt{cases}.  All theorems from {\tt
  2063   distinct}, \texttt{inject} and \texttt{cases} are combined in \texttt{simps}.
  2064 In case of mutually recursive datatypes, \texttt{recs}, \texttt{size}, \texttt{induct}
  2065 and \texttt{simps} are contained in a separate structure named $t@1_\ldots_t@n$.
  2066 
  2067 
  2068 \subsection{Representing existing types as datatypes}\label{subsec:datatype:rep_datatype}
  2069 
  2070 For foundational reasons, some basic types such as \texttt{nat}, \texttt{*}, {\tt
  2071   +}, \texttt{bool} and \texttt{unit} are not defined in a \texttt{datatype} section,
  2072 but by more primitive means using \texttt{typedef}. To be able to use the
  2073 tactics \texttt{induct_tac} and \texttt{case_tac} and to define functions by
  2074 primitive recursion on these types, such types may be represented as actual
  2075 datatypes.  This is done by specifying the constructors of the desired type,
  2076 plus a proof of the  induction rule, as well as theorems
  2077 stating the distinctness and injectivity of constructors in a {\tt
  2078 rep_datatype} section.  For the sum type this works as follows:
  2079 \begin{ttbox}
  2080 rep_datatype (sum) Inl Inr
  2081 proof -
  2082   fix P
  2083   fix s :: "'a + 'b"
  2084   assume x: "!!x::'a. P (Inl x)" and y: "!!y::'b. P (Inr y)"
  2085   then show "P s" by (auto intro: sumE [of s])
  2086 qed simp_all
  2087 \end{ttbox}
  2088 The datatype package automatically derives additional theorems for recursion
  2089 and case combinators from these rules.  Any of the basic HOL types mentioned
  2090 above are represented as datatypes.  Try an induction on \texttt{bool}
  2091 today.
  2092 
  2093 
  2094 \subsection{Examples}
  2095 
  2096 \subsubsection{The datatype $\alpha~mylist$}
  2097 
  2098 We want to define a type $\alpha~mylist$. To do this we have to build a new
  2099 theory that contains the type definition.  We start from the theory
  2100 \texttt{Datatype} instead of \texttt{Main} in order to avoid clashes with the
  2101 \texttt{List} theory of Isabelle/HOL.
  2102 \begin{ttbox}
  2103 MyList = Datatype +
  2104   datatype 'a mylist = Nil | Cons 'a ('a mylist)
  2105 end
  2106 \end{ttbox}
  2107 After loading the theory, we can prove $Cons~x~xs\neq xs$, for example.  To
  2108 ease the induction applied below, we state the goal with $x$ quantified at the
  2109 object-level.  This will be stripped later using \ttindex{qed_spec_mp}.
  2110 \begin{ttbox}
  2111 Goal "!x. Cons x xs ~= xs";
  2112 {\out Level 0}
  2113 {\out ! x. Cons x xs ~= xs}
  2114 {\out  1. ! x. Cons x xs ~= xs}
  2115 \end{ttbox}
  2116 This can be proved by the structural induction tactic:
  2117 \begin{ttbox}
  2118 by (induct_tac "xs" 1);
  2119 {\out Level 1}
  2120 {\out ! x. Cons x xs ~= xs}
  2121 {\out  1. ! x. Cons x Nil ~= Nil}
  2122 {\out  2. !!a mylist.}
  2123 {\out        ! x. Cons x mylist ~= mylist ==>}
  2124 {\out        ! x. Cons x (Cons a mylist) ~= Cons a mylist}
  2125 \end{ttbox}
  2126 The first subgoal can be proved using the simplifier.  Isabelle/HOL has
  2127 already added the freeness properties of lists to the default simplification
  2128 set.
  2129 \begin{ttbox}
  2130 by (Simp_tac 1);
  2131 {\out Level 2}
  2132 {\out ! x. Cons x xs ~= xs}
  2133 {\out  1. !!a mylist.}
  2134 {\out        ! x. Cons x mylist ~= mylist ==>}
  2135 {\out        ! x. Cons x (Cons a mylist) ~= Cons a mylist}
  2136 \end{ttbox}
  2137 Similarly, we prove the remaining goal.
  2138 \begin{ttbox}
  2139 by (Asm_simp_tac 1);
  2140 {\out Level 3}
  2141 {\out ! x. Cons x xs ~= xs}
  2142 {\out No subgoals!}
  2143 \ttbreak
  2144 qed_spec_mp "not_Cons_self";
  2145 {\out val not_Cons_self = "Cons x xs ~= xs" : thm}
  2146 \end{ttbox}
  2147 Because both subgoals could have been proved by \texttt{Asm_simp_tac}
  2148 we could have done that in one step:
  2149 \begin{ttbox}
  2150 by (ALLGOALS Asm_simp_tac);
  2151 \end{ttbox}
  2152 
  2153 
  2154 \subsubsection{The datatype $\alpha~mylist$ with mixfix syntax}
  2155 
  2156 In this example we define the type $\alpha~mylist$ again but this time
  2157 we want to write \texttt{[]} for \texttt{Nil} and we want to use infix
  2158 notation \verb|#| for \texttt{Cons}.  To do this we simply add mixfix
  2159 annotations after the constructor declarations as follows:
  2160 \begin{ttbox}
  2161 MyList = Datatype +
  2162   datatype 'a mylist =
  2163     Nil ("[]")  |
  2164     Cons 'a ('a mylist)  (infixr "#" 70)
  2165 end
  2166 \end{ttbox}
  2167 Now the theorem in the previous example can be written \verb|x#xs ~= xs|.
  2168 
  2169 
  2170 \subsubsection{A datatype for weekdays}
  2171 
  2172 This example shows a datatype that consists of 7 constructors:
  2173 \begin{ttbox}
  2174 Days = Main +
  2175   datatype days = Mon | Tue | Wed | Thu | Fri | Sat | Sun
  2176 end
  2177 \end{ttbox}
  2178 Because there are more than 6 constructors, inequality is expressed via a function
  2179 \verb|days_ord|.  The theorem \verb|Mon ~= Tue| is not directly
  2180 contained among the distinctness theorems, but the simplifier can
  2181 prove it thanks to rewrite rules inherited from theory \texttt{NatArith}:
  2182 \begin{ttbox}
  2183 Goal "Mon ~= Tue";
  2184 by (Simp_tac 1);
  2185 \end{ttbox}
  2186 You need not derive such inequalities explicitly: the simplifier will dispose
  2187 of them automatically.
  2188 \index{*datatype|)}
  2189 
  2190 
  2191 \section{Recursive function definitions}\label{sec:HOL:recursive}
  2192 \index{recursive functions|see{recursion}}
  2193 
  2194 Isabelle/HOL provides two main mechanisms of defining recursive functions.
  2195 \begin{enumerate}
  2196 \item \textbf{Primitive recursion} is available only for datatypes, and it is
  2197   somewhat restrictive.  Recursive calls are only allowed on the argument's
  2198   immediate constituents.  On the other hand, it is the form of recursion most
  2199   often wanted, and it is easy to use.
  2200   
  2201 \item \textbf{Well-founded recursion} requires that you supply a well-founded
  2202   relation that governs the recursion.  Recursive calls are only allowed if
  2203   they make the argument decrease under the relation.  Complicated recursion
  2204   forms, such as nested recursion, can be dealt with.  Termination can even be
  2205   proved at a later time, though having unsolved termination conditions around
  2206   can make work difficult.%
  2207   \footnote{This facility is based on Konrad Slind's TFL
  2208     package~\cite{slind-tfl}.  Thanks are due to Konrad for implementing TFL
  2209     and assisting with its installation.}
  2210 \end{enumerate}
  2211 
  2212 Following good HOL tradition, these declarations do not assert arbitrary
  2213 axioms.  Instead, they define the function using a recursion operator.  Both
  2214 HOL and ZF derive the theory of well-founded recursion from first
  2215 principles~\cite{paulson-set-II}.  Primitive recursion over some datatype
  2216 relies on the recursion operator provided by the datatype package.  With
  2217 either form of function definition, Isabelle proves the desired recursion
  2218 equations as theorems.
  2219 
  2220 
  2221 \subsection{Primitive recursive functions}
  2222 \label{sec:HOL:primrec}
  2223 \index{recursion!primitive|(}
  2224 \index{*primrec|(}
  2225 
  2226 Datatypes come with a uniform way of defining functions, {\bf primitive
  2227   recursion}.  In principle, one could introduce primitive recursive functions
  2228 by asserting their reduction rules as new axioms, but this is not recommended:
  2229 \begin{ttbox}\slshape
  2230 Append = Main +
  2231 consts app :: ['a list, 'a list] => 'a list
  2232 rules 
  2233    app_Nil   "app [] ys = ys"
  2234    app_Cons  "app (x#xs) ys = x#app xs ys"
  2235 end
  2236 \end{ttbox}
  2237 Asserting axioms brings the danger of accidentally asserting nonsense, as
  2238 in \verb$app [] ys = us$.
  2239 
  2240 The \ttindex{primrec} declaration is a safe means of defining primitive
  2241 recursive functions on datatypes:
  2242 \begin{ttbox}
  2243 Append = Main +
  2244 consts app :: ['a list, 'a list] => 'a list
  2245 primrec
  2246    "app [] ys = ys"
  2247    "app (x#xs) ys = x#app xs ys"
  2248 end
  2249 \end{ttbox}
  2250 Isabelle will now check that the two rules do indeed form a primitive
  2251 recursive definition.  For example
  2252 \begin{ttbox}
  2253 primrec
  2254     "app [] ys = us"
  2255 \end{ttbox}
  2256 is rejected with an error message ``\texttt{Extra variables on rhs}''.
  2257 
  2258 \bigskip
  2259 
  2260 The general form of a primitive recursive definition is
  2261 \begin{ttbox}
  2262 primrec
  2263     {\it reduction rules}
  2264 \end{ttbox}
  2265 where \textit{reduction rules} specify one or more equations of the form
  2266 \[ f \, x@1 \, \dots \, x@m \, (C \, y@1 \, \dots \, y@k) \, z@1 \,
  2267 \dots \, z@n = r \] such that $C$ is a constructor of the datatype, $r$
  2268 contains only the free variables on the left-hand side, and all recursive
  2269 calls in $r$ are of the form $f \, \dots \, y@i \, \dots$ for some $i$.  There
  2270 must be at most one reduction rule for each constructor.  The order is
  2271 immaterial.  For missing constructors, the function is defined to return a
  2272 default value.  
  2273 
  2274 If you would like to refer to some rule by name, then you must prefix
  2275 the rule with an identifier.  These identifiers, like those in the
  2276 \texttt{rules} section of a theory, will be visible at the \ML\ level.
  2277 
  2278 The primitive recursive function can have infix or mixfix syntax:
  2279 \begin{ttbox}\underscoreon
  2280 consts "@"  :: ['a list, 'a list] => 'a list  (infixr 60)
  2281 primrec
  2282    "[] @ ys = ys"
  2283    "(x#xs) @ ys = x#(xs @ ys)"
  2284 \end{ttbox}
  2285 
  2286 The reduction rules become part of the default simpset, which
  2287 leads to short proof scripts:
  2288 \begin{ttbox}\underscoreon
  2289 Goal "(xs @ ys) @ zs = xs @ (ys @ zs)";
  2290 by (induct\_tac "xs" 1);
  2291 by (ALLGOALS Asm\_simp\_tac);
  2292 \end{ttbox}
  2293 
  2294 \subsubsection{Example: Evaluation of expressions}
  2295 Using mutual primitive recursion, we can define evaluation functions \texttt{evala}
  2296 and \texttt{eval_bexp} for the datatypes of arithmetic and boolean expressions mentioned in
  2297 {\S}\ref{subsec:datatype:basics}:
  2298 \begin{ttbox}
  2299 consts
  2300   evala :: "['a => nat, 'a aexp] => nat"
  2301   evalb :: "['a => nat, 'a bexp] => bool"
  2302 
  2303 primrec
  2304   "evala env (If_then_else b a1 a2) =
  2305      (if evalb env b then evala env a1 else evala env a2)"
  2306   "evala env (Sum a1 a2) = evala env a1 + evala env a2"
  2307   "evala env (Diff a1 a2) = evala env a1 - evala env a2"
  2308   "evala env (Var v) = env v"
  2309   "evala env (Num n) = n"
  2310 
  2311   "evalb env (Less a1 a2) = (evala env a1 < evala env a2)"
  2312   "evalb env (And b1 b2) = (evalb env b1 & evalb env b2)"
  2313   "evalb env (Or b1 b2) = (evalb env b1 & evalb env b2)"
  2314 \end{ttbox}
  2315 Since the value of an expression depends on the value of its variables,
  2316 the functions \texttt{evala} and \texttt{evalb} take an additional
  2317 parameter, an {\em environment} of type \texttt{'a => nat}, which maps
  2318 variables to their values.
  2319 
  2320 Similarly, we may define substitution functions \texttt{substa}
  2321 and \texttt{substb} for expressions: The mapping \texttt{f} of type
  2322 \texttt{'a => 'a aexp} given as a parameter is lifted canonically
  2323 on the types \texttt{'a aexp} and \texttt{'a bexp}:
  2324 \begin{ttbox}
  2325 consts
  2326   substa :: "['a => 'b aexp, 'a aexp] => 'b aexp"
  2327   substb :: "['a => 'b aexp, 'a bexp] => 'b bexp"
  2328 
  2329 primrec
  2330   "substa f (If_then_else b a1 a2) =
  2331      If_then_else (substb f b) (substa f a1) (substa f a2)"
  2332   "substa f (Sum a1 a2) = Sum (substa f a1) (substa f a2)"
  2333   "substa f (Diff a1 a2) = Diff (substa f a1) (substa f a2)"
  2334   "substa f (Var v) = f v"
  2335   "substa f (Num n) = Num n"
  2336 
  2337   "substb f (Less a1 a2) = Less (substa f a1) (substa f a2)"
  2338   "substb f (And b1 b2) = And (substb f b1) (substb f b2)"
  2339   "substb f (Or b1 b2) = Or (substb f b1) (substb f b2)"
  2340 \end{ttbox}
  2341 In textbooks about semantics one often finds {\em substitution theorems},
  2342 which express the relationship between substitution and evaluation. For
  2343 \texttt{'a aexp} and \texttt{'a bexp}, we can prove such a theorem by mutual
  2344 induction, followed by simplification:
  2345 \begin{ttbox}
  2346 Goal
  2347   "evala env (substa (Var(v := a')) a) =
  2348      evala (env(v := evala env a')) a &
  2349    evalb env (substb (Var(v := a')) b) =
  2350      evalb (env(v := evala env a')) b";
  2351 by (induct_tac "a b" 1);
  2352 by (ALLGOALS Asm_full_simp_tac);
  2353 \end{ttbox}
  2354 
  2355 \subsubsection{Example: A substitution function for terms}
  2356 Functions on datatypes with nested recursion, such as the type
  2357 \texttt{term} mentioned in {\S}\ref{subsec:datatype:basics}, are
  2358 also defined by mutual primitive recursion. A substitution
  2359 function \texttt{subst_term} on type \texttt{term}, similar to the functions
  2360 \texttt{substa} and \texttt{substb} described above, can
  2361 be defined as follows:
  2362 \begin{ttbox}
  2363 consts
  2364   subst_term :: "['a => ('a,'b) term, ('a,'b) term] => ('a,'b) term"
  2365   subst_term_list ::
  2366     "['a => ('a,'b) term, ('a,'b) term list] => ('a,'b) term list"
  2367 
  2368 primrec
  2369   "subst_term f (Var a) = f a"
  2370   "subst_term f (App b ts) = App b (subst_term_list f ts)"
  2371 
  2372   "subst_term_list f [] = []"
  2373   "subst_term_list f (t # ts) =
  2374      subst_term f t # subst_term_list f ts"
  2375 \end{ttbox}
  2376 The recursion scheme follows the structure of the unfolded definition of type
  2377 \texttt{term} shown in {\S}\ref{subsec:datatype:basics}. To prove properties of
  2378 this substitution function, mutual induction is needed:
  2379 \begin{ttbox}
  2380 Goal
  2381   "(subst_term ((subst_term f1) o f2) t) =
  2382      (subst_term f1 (subst_term f2 t)) &
  2383    (subst_term_list ((subst_term f1) o f2) ts) =
  2384      (subst_term_list f1 (subst_term_list f2 ts))";
  2385 by (induct_tac "t ts" 1);
  2386 by (ALLGOALS Asm_full_simp_tac);
  2387 \end{ttbox}
  2388 
  2389 \subsubsection{Example: A map function for infinitely branching trees}
  2390 Defining functions on infinitely branching datatypes by primitive
  2391 recursion is just as easy. For example, we can define a function
  2392 \texttt{map_tree} on \texttt{'a tree} as follows:
  2393 \begin{ttbox}
  2394 consts
  2395   map_tree :: "('a => 'b) => 'a tree => 'b tree"
  2396 
  2397 primrec
  2398   "map_tree f (Atom a) = Atom (f a)"
  2399   "map_tree f (Branch ts) = Branch (\%x. map_tree f (ts x))"
  2400 \end{ttbox}
  2401 Note that all occurrences of functions such as \texttt{ts} in the
  2402 \texttt{primrec} clauses must be applied to an argument. In particular,
  2403 \texttt{map_tree f o ts} is not allowed.
  2404 
  2405 \index{recursion!primitive|)}
  2406 \index{*primrec|)}
  2407 
  2408 
  2409 \subsection{General recursive functions}
  2410 \label{sec:HOL:recdef}
  2411 \index{recursion!general|(}
  2412 \index{*recdef|(}
  2413 
  2414 Using \texttt{recdef}, you can declare functions involving nested recursion
  2415 and pattern-matching.  Recursion need not involve datatypes and there are few
  2416 syntactic restrictions.  Termination is proved by showing that each recursive
  2417 call makes the argument smaller in a suitable sense, which you specify by
  2418 supplying a well-founded relation.
  2419 
  2420 Here is a simple example, the Fibonacci function.  The first line declares
  2421 \texttt{fib} to be a constant.  The well-founded relation is simply~$<$ (on
  2422 the natural numbers).  Pattern-matching is used here: \texttt{1} is a
  2423 macro for \texttt{Suc~0}.
  2424 \begin{ttbox}
  2425 consts fib  :: "nat => nat"
  2426 recdef fib "less_than"
  2427     "fib 0 = 0"
  2428     "fib 1 = 1"
  2429     "fib (Suc(Suc x)) = (fib x + fib (Suc x))"
  2430 \end{ttbox}
  2431 
  2432 With \texttt{recdef}, function definitions may be incomplete, and patterns may
  2433 overlap, as in functional programming.  The \texttt{recdef} package
  2434 disambiguates overlapping patterns by taking the order of rules into account.
  2435 For missing patterns, the function is defined to return a default value.
  2436 
  2437 %For example, here is a declaration of the list function \cdx{hd}:
  2438 %\begin{ttbox}
  2439 %consts hd :: 'a list => 'a
  2440 %recdef hd "\{\}"
  2441 %    "hd (x#l) = x"
  2442 %\end{ttbox}
  2443 %Because this function is not recursive, we may supply the empty well-founded
  2444 %relation, $\{\}$.
  2445 
  2446 The well-founded relation defines a notion of ``smaller'' for the function's
  2447 argument type.  The relation $\prec$ is \textbf{well-founded} provided it
  2448 admits no infinitely decreasing chains
  2449 \[ \cdots\prec x@n\prec\cdots\prec x@1. \]
  2450 If the function's argument has type~$\tau$, then $\prec$ has to be a relation
  2451 over~$\tau$: it must have type $(\tau\times\tau)set$.
  2452 
  2453 Proving well-foundedness can be tricky, so Isabelle/HOL provides a collection
  2454 of operators for building well-founded relations.  The package recognises
  2455 these operators and automatically proves that the constructed relation is
  2456 well-founded.  Here are those operators, in order of importance:
  2457 \begin{itemize}
  2458 \item \texttt{less_than} is ``less than'' on the natural numbers.
  2459   (It has type $(nat\times nat)set$, while $<$ has type $[nat,nat]\To bool$.
  2460   
  2461 \item $\mathop{\mathtt{measure}} f$, where $f$ has type $\tau\To nat$, is the
  2462   relation~$\prec$ on type~$\tau$ such that $x\prec y$ if and only if
  2463   $f(x)<f(y)$.  
  2464   Typically, $f$ takes the recursive function's arguments (as a tuple) and
  2465   returns a result expressed in terms of the function \texttt{size}.  It is
  2466   called a \textbf{measure function}.  Recall that \texttt{size} is overloaded
  2467   and is defined on all datatypes (see {\S}\ref{sec:HOL:size}).
  2468                                                     
  2469 \item $\mathop{\mathtt{inv_image}} R\;f$ is a generalisation of
  2470   \texttt{measure}.  It specifies a relation such that $x\prec y$ if and only
  2471   if $f(x)$ 
  2472   is less than $f(y)$ according to~$R$, which must itself be a well-founded
  2473   relation.
  2474 
  2475 \item $R@1\texttt{<*lex*>}R@2$ is the lexicographic product of two relations.
  2476   It 
  2477   is a relation on pairs and satisfies $(x@1,x@2)\prec(y@1,y@2)$ if and only
  2478   if $x@1$ 
  2479   is less than $y@1$ according to~$R@1$ or $x@1=y@1$ and $x@2$
  2480   is less than $y@2$ according to~$R@2$.
  2481 
  2482 \item \texttt{finite_psubset} is the proper subset relation on finite sets.
  2483 \end{itemize}
  2484 
  2485 We can use \texttt{measure} to declare Euclid's algorithm for the greatest
  2486 common divisor.  The measure function, $\lambda(m,n). n$, specifies that the
  2487 recursion terminates because argument~$n$ decreases.
  2488 \begin{ttbox}
  2489 recdef gcd "measure ((\%(m,n). n) ::nat*nat=>nat)"
  2490     "gcd (m, n) = (if n=0 then m else gcd(n, m mod n))"
  2491 \end{ttbox}
  2492 
  2493 The general form of a well-founded recursive definition is
  2494 \begin{ttbox}
  2495 recdef {\it function} {\it rel}
  2496     congs   {\it congruence rules}      {\bf(optional)}
  2497     simpset {\it simplification set}      {\bf(optional)}
  2498    {\it reduction rules}
  2499 \end{ttbox}
  2500 where
  2501 \begin{itemize}
  2502 \item \textit{function} is the name of the function, either as an \textit{id}
  2503   or a \textit{string}.  
  2504   
  2505 \item \textit{rel} is a HOL expression for the well-founded termination
  2506   relation.
  2507   
  2508 \item \textit{congruence rules} are required only in highly exceptional
  2509   circumstances.
  2510   
  2511 \item The \textit{simplification set} is used to prove that the supplied
  2512   relation is well-founded.  It is also used to prove the \textbf{termination
  2513     conditions}: assertions that arguments of recursive calls decrease under
  2514   \textit{rel}.  By default, simplification uses \texttt{simpset()}, which
  2515   is sufficient to prove well-foundedness for the built-in relations listed
  2516   above. 
  2517   
  2518 \item \textit{reduction rules} specify one or more recursion equations.  Each
  2519   left-hand side must have the form $f\,t$, where $f$ is the function and $t$
  2520   is a tuple of distinct variables.  If more than one equation is present then
  2521   $f$ is defined by pattern-matching on components of its argument whose type
  2522   is a \texttt{datatype}.  
  2523 
  2524   The \ML\ identifier $f$\texttt{.simps} contains the reduction rules as
  2525   a list of theorems.
  2526 \end{itemize}
  2527 
  2528 With the definition of \texttt{gcd} shown above, Isabelle/HOL is unable to
  2529 prove one termination condition.  It remains as a precondition of the
  2530 recursion theorems:
  2531 \begin{ttbox}
  2532 gcd.simps;
  2533 {\out ["! m n. n ~= 0 --> m mod n < n}
  2534 {\out   ==> gcd (?m,?n) = (if ?n=0 then ?m else gcd (?n, ?m mod ?n))"] }
  2535 {\out : thm list}
  2536 \end{ttbox}
  2537 The theory \texttt{HOL/ex/Primes} illustrates how to prove termination
  2538 conditions afterwards.  The function \texttt{Tfl.tgoalw} is like the standard
  2539 function \texttt{goalw}, which sets up a goal to prove, but its argument
  2540 should be the identifier $f$\texttt{.simps} and its effect is to set up a
  2541 proof of the termination conditions:
  2542 \begin{ttbox}
  2543 Tfl.tgoalw thy [] gcd.simps;
  2544 {\out Level 0}
  2545 {\out ! m n. n ~= 0 --> m mod n < n}
  2546 {\out  1. ! m n. n ~= 0 --> m mod n < n}
  2547 \end{ttbox}
  2548 This subgoal has a one-step proof using \texttt{simp_tac}.  Once the theorem
  2549 is proved, it can be used to eliminate the termination conditions from
  2550 elements of \texttt{gcd.simps}.  Theory \texttt{HOL/Subst/Unify} is a much
  2551 more complicated example of this process, where the termination conditions can
  2552 only be proved by complicated reasoning involving the recursive function
  2553 itself.
  2554 
  2555 Isabelle/HOL can prove the \texttt{gcd} function's termination condition
  2556 automatically if supplied with the right simpset.
  2557 \begin{ttbox}
  2558 recdef gcd "measure ((\%(m,n). n) ::nat*nat=>nat)"
  2559   simpset "simpset() addsimps [mod_less_divisor, zero_less_eq]"
  2560     "gcd (m, n) = (if n=0 then m else gcd(n, m mod n))"
  2561 \end{ttbox}
  2562 
  2563 If all termination conditions were proved automatically, $f$\texttt{.simps}
  2564 is added to the simpset automatically, just as in \texttt{primrec}. 
  2565 The simplification rules corresponding to clause $i$ (where counting starts
  2566 at 0) are called $f$\texttt{.}$i$ and can be accessed as \texttt{thms
  2567   "$f$.$i$"},
  2568 which returns a list of theorems. Thus you can, for example, remove specific
  2569 clauses from the simpset. Note that a single clause may give rise to a set of
  2570 simplification rules in order to capture the fact that if clauses overlap,
  2571 their order disambiguates them.
  2572 
  2573 A \texttt{recdef} definition also returns an induction rule specialised for
  2574 the recursive function.  For the \texttt{gcd} function above, the induction
  2575 rule is
  2576 \begin{ttbox}
  2577 gcd.induct;
  2578 {\out "(!!m n. n ~= 0 --> ?P n (m mod n) ==> ?P m n) ==> ?P ?u ?v" : thm}
  2579 \end{ttbox}
  2580 This rule should be used to reason inductively about the \texttt{gcd}
  2581 function.  It usually makes the induction hypothesis available at all
  2582 recursive calls, leading to very direct proofs.  If any termination conditions
  2583 remain unproved, they will become additional premises of this rule.
  2584 
  2585 \index{recursion!general|)}
  2586 \index{*recdef|)}
  2587 
  2588 
  2589 \section{Inductive and coinductive definitions}
  2590 \index{*inductive|(}
  2591 \index{*coinductive|(}
  2592 
  2593 An {\bf inductive definition} specifies the least set~$R$ closed under given
  2594 rules.  (Applying a rule to elements of~$R$ yields a result within~$R$.)  For
  2595 example, a structural operational semantics is an inductive definition of an
  2596 evaluation relation.  Dually, a {\bf coinductive definition} specifies the
  2597 greatest set~$R$ consistent with given rules.  (Every element of~$R$ can be
  2598 seen as arising by applying a rule to elements of~$R$.)  An important example
  2599 is using bisimulation relations to formalise equivalence of processes and
  2600 infinite data structures.
  2601 
  2602 A theory file may contain any number of inductive and coinductive
  2603 definitions.  They may be intermixed with other declarations; in
  2604 particular, the (co)inductive sets {\bf must} be declared separately as
  2605 constants, and may have mixfix syntax or be subject to syntax translations.
  2606 
  2607 Each (co)inductive definition adds definitions to the theory and also
  2608 proves some theorems.  Each definition creates an \ML\ structure, which is a
  2609 substructure of the main theory structure.
  2610 
  2611 This package is related to the ZF one, described in a separate
  2612 paper,%
  2613 \footnote{It appeared in CADE~\cite{paulson-CADE}; a longer version is
  2614   distributed with Isabelle.}  %
  2615 which you should refer to in case of difficulties.  The package is simpler
  2616 than ZF's thanks to HOL's extra-logical automatic type-checking.  The types of
  2617 the (co)inductive sets determine the domain of the fixedpoint definition, and
  2618 the package does not have to use inference rules for type-checking.
  2619 
  2620 
  2621 \subsection{The result structure}
  2622 Many of the result structure's components have been discussed in the paper;
  2623 others are self-explanatory.
  2624 \begin{description}
  2625 \item[\tt defs] is the list of definitions of the recursive sets.
  2626 
  2627 \item[\tt mono] is a monotonicity theorem for the fixedpoint operator.
  2628 
  2629 \item[\tt unfold] is a fixedpoint equation for the recursive set (the union of
  2630 the recursive sets, in the case of mutual recursion).
  2631 
  2632 \item[\tt intrs] is the list of introduction rules, now proved as theorems, for
  2633 the recursive sets.  The rules are also available individually, using the
  2634 names given them in the theory file. 
  2635 
  2636 \item[\tt elims] is the list of elimination rule.  This is for compatibility
  2637   with ML scripts; within the theory the name is \texttt{cases}.
  2638   
  2639 \item[\tt elim] is the head of the list \texttt{elims}.  This is for
  2640   compatibility only.
  2641   
  2642 \item[\tt mk_cases] is a function to create simplified instances of {\tt
  2643 elim} using freeness reasoning on underlying datatypes.
  2644 \end{description}
  2645 
  2646 For an inductive definition, the result structure contains the
  2647 rule \texttt{induct}.  For a
  2648 coinductive definition, it contains the rule \verb|coinduct|.
  2649 
  2650 Figure~\ref{def-result-fig} summarises the two result signatures,
  2651 specifying the types of all these components.
  2652 
  2653 \begin{figure}
  2654 \begin{ttbox}
  2655 sig
  2656 val defs         : thm list
  2657 val mono         : thm
  2658 val unfold       : thm
  2659 val intrs        : thm list
  2660 val elims        : thm list
  2661 val elim         : thm
  2662 val mk_cases     : string -> thm
  2663 {\it(Inductive definitions only)} 
  2664 val induct       : thm
  2665 {\it(coinductive definitions only)}
  2666 val coinduct     : thm
  2667 end
  2668 \end{ttbox}
  2669 \hrule
  2670 \caption{The {\ML} result of a (co)inductive definition} \label{def-result-fig}
  2671 \end{figure}
  2672 
  2673 \subsection{The syntax of a (co)inductive definition}
  2674 An inductive definition has the form
  2675 \begin{ttbox}
  2676 inductive    {\it inductive sets}
  2677   intrs      {\it introduction rules}
  2678   monos      {\it monotonicity theorems}
  2679 \end{ttbox}
  2680 A coinductive definition is identical, except that it starts with the keyword
  2681 \texttt{coinductive}.  
  2682 
  2683 The \texttt{monos} section is optional; if present it is specified by a list
  2684 of identifiers.
  2685 
  2686 \begin{itemize}
  2687 \item The \textit{inductive sets} are specified by one or more strings.
  2688 
  2689 \item The \textit{introduction rules} specify one or more introduction rules in
  2690   the form \textit{ident\/}~\textit{string}, where the identifier gives the name of
  2691   the rule in the result structure.
  2692 
  2693 \item The \textit{monotonicity theorems} are required for each operator
  2694   applied to a recursive set in the introduction rules.  There {\bf must}
  2695   be a theorem of the form $A\subseteq B\Imp M(A)\subseteq M(B)$, for each
  2696   premise $t\in M(R@i)$ in an introduction rule!
  2697 
  2698 \item The \textit{constructor definitions} contain definitions of constants
  2699   appearing in the introduction rules.  In most cases it can be omitted.
  2700 \end{itemize}
  2701 
  2702 
  2703 \subsection{*Monotonicity theorems}
  2704 
  2705 Each theory contains a default set of theorems that are used in monotonicity
  2706 proofs. New rules can be added to this set via the \texttt{mono} attribute.
  2707 Theory \texttt{Inductive} shows how this is done. In general, the following
  2708 monotonicity theorems may be added:
  2709 \begin{itemize}
  2710 \item Theorems of the form $A\subseteq B\Imp M(A)\subseteq M(B)$, for proving
  2711   monotonicity of inductive definitions whose introduction rules have premises
  2712   involving terms such as $t\in M(R@i)$.
  2713 \item Monotonicity theorems for logical operators, which are of the general form
  2714   $\List{\cdots \to \cdots;~\ldots;~\cdots \to \cdots} \Imp
  2715     \cdots \to \cdots$.
  2716   For example, in the case of the operator $\lor$, the corresponding theorem is
  2717   \[
  2718   \infer{P@1 \lor P@2 \to Q@1 \lor Q@2}
  2719     {P@1 \to Q@1 & P@2 \to Q@2}
  2720   \]
  2721 \item De Morgan style equations for reasoning about the ``polarity'' of expressions, e.g.
  2722   \[
  2723   (\lnot \lnot P) ~=~ P \qquad\qquad
  2724   (\lnot (P \land Q)) ~=~ (\lnot P \lor \lnot Q)
  2725   \]
  2726 \item Equations for reducing complex operators to more primitive ones whose
  2727   monotonicity can easily be proved, e.g.
  2728   \[
  2729   (P \to Q) ~=~ (\lnot P \lor Q) \qquad\qquad
  2730   \mathtt{Ball}~A~P ~\equiv~ \forall x.~x \in A \to P~x
  2731   \]
  2732 \end{itemize}
  2733 
  2734 \subsection{Example of an inductive definition}
  2735 Two declarations, included in a theory file, define the finite powerset
  2736 operator.  First we declare the constant~\texttt{Fin}.  Then we declare it
  2737 inductively, with two introduction rules:
  2738 \begin{ttbox}
  2739 consts Fin :: 'a set => 'a set set
  2740 inductive "Fin A"
  2741   intrs
  2742     emptyI  "{\ttlbrace}{\ttrbrace} : Fin A"
  2743     insertI "[| a: A;  b: Fin A |] ==> insert a b : Fin A"
  2744 \end{ttbox}
  2745 The resulting theory structure contains a substructure, called~\texttt{Fin}.
  2746 It contains the \texttt{Fin}$~A$ introduction rules as the list \texttt{Fin.intrs},
  2747 and also individually as \texttt{Fin.emptyI} and \texttt{Fin.consI}.  The induction
  2748 rule is \texttt{Fin.induct}.
  2749 
  2750 For another example, here is a theory file defining the accessible part of a
  2751 relation.  The paper \cite{paulson-CADE} discusses a ZF version of this
  2752 example in more detail.
  2753 \begin{ttbox}
  2754 Acc = WF + Inductive +
  2755 
  2756 consts acc :: "('a * 'a)set => 'a set"   (* accessible part *)
  2757 
  2758 inductive "acc r"
  2759   intrs
  2760     accI "ALL y. (y, x) : r --> y : acc r ==> x : acc r"
  2761 
  2762 end
  2763 \end{ttbox}
  2764 The Isabelle distribution contains many other inductive definitions.  Simple
  2765 examples are collected on subdirectory \texttt{HOL/Induct}.  The theory
  2766 \texttt{HOL/Induct/LList} contains coinductive definitions.  Larger examples
  2767 may be found on other subdirectories of \texttt{HOL}, such as \texttt{IMP},
  2768 \texttt{Lambda} and \texttt{Auth}.
  2769 
  2770 \index{*coinductive|)} \index{*inductive|)}
  2771 
  2772 
  2773 \section{Executable specifications}
  2774 \index{code generator}
  2775 
  2776 For validation purposes, it is often useful to {\em execute} specifications.
  2777 In principle, specifications could be ``executed'' using Isabelle's
  2778 inference kernel, i.e. by a combination of resolution and simplification.
  2779 Unfortunately, this approach is rather inefficient. A more efficient way
  2780 of executing specifications is to translate them into a functional
  2781 programming language such as ML. Isabelle's built-in code generator
  2782 supports this.
  2783 
  2784 \railalias{verblbrace}{\texttt{\ttlbrace*}}
  2785 \railalias{verbrbrace}{\texttt{*\ttrbrace}}
  2786 \railterm{verblbrace}
  2787 \railterm{verbrbrace}
  2788 
  2789 \begin{figure}
  2790 \begin{rail}
  2791 codegen : ( 'code_module' | 'code_library' ) modespec ? name ? \\
  2792   ( 'file' name ) ? ( 'imports' ( name + ) ) ? \\
  2793   'contains' ( ( name '=' term ) + | term + );
  2794 
  2795 modespec : '(' ( name * ) ')';
  2796 \end{rail}
  2797 \caption{Code generator invocation syntax}
  2798 \label{fig:HOL:codegen-invocation}
  2799 \end{figure}
  2800 
  2801 \begin{figure}
  2802 \begin{rail}
  2803 constscode : 'consts_code' (codespec +);
  2804 
  2805 codespec : const template attachment ?;
  2806 
  2807 typescode : 'types_code' (tycodespec +);
  2808 
  2809 tycodespec : name template attachment ?;
  2810 
  2811 const : term;
  2812 
  2813 template: '(' string ')';
  2814 
  2815 attachment: 'attach' modespec ? verblbrace text verbrbrace;
  2816 \end{rail}
  2817 \caption{Code generator configuration syntax}
  2818 \label{fig:HOL:codegen-configuration}
  2819 \end{figure}
  2820 
  2821 \subsection{Invoking the code generator}
  2822 
  2823 The code generator is invoked via the \ttindex{code_module} and
  2824 \ttindex{code_library} commands (see Fig.~\ref{fig:HOL:codegen-invocation}),
  2825 which correspond to {\em incremental} and {\em modular} code generation,
  2826 respectively.
  2827 \begin{description}
  2828 \item[Modular] For each theory, an ML structure is generated, containing the
  2829   code generated from the constants defined in this theory.
  2830 \item[Incremental] All the generated code is emitted into the same structure.
  2831   This structure may import code from previously generated structures, which
  2832   can be specified via \texttt{imports}.
  2833   Moreover, the generated structure may also be referred to in later invocations
  2834   of the code generator.
  2835 \end{description}
  2836 After the \texttt{code_module} and \texttt{code_library} keywords, the user
  2837 may specify an optional list of ``modes'' in parentheses. These can be used
  2838 to instruct the code generator to emit additional code for special purposes,
  2839 e.g.\ functions for converting elements of generated datatypes to Isabelle terms,
  2840 or test data generators. The list of modes is followed by a module name.
  2841 The module name is optional for modular code generation, but must be specified
  2842 for incremental code generation.
  2843 The code can either be written to a file,
  2844 in which case a file name has to be specified after the \texttt{file} keyword, or be
  2845 loaded directly into Isabelle's ML environment. In the latter case,
  2846 the \texttt{ML} theory command can be used to inspect the results
  2847 interactively.
  2848 The terms from which to generate code can be specified after the
  2849 \texttt{contains} keyword, either as a list of bindings, or just as
  2850 a list of terms. In the latter case, the code generator just produces
  2851 code for all constants and types occuring in the term, but does not
  2852 bind the compiled terms to ML identifiers.
  2853 For example,
  2854 \begin{ttbox}
  2855 code_module Test
  2856 contains
  2857   test = "foldl op + (0::int) [1,2,3,4,5]"
  2858 \end{ttbox}
  2859 binds the result of compiling the term
  2860 \texttt{foldl op + (0::int) [1,2,3,4,5]}
  2861 (i.e.~\texttt{15}) to the ML identifier \texttt{Test.test}.
  2862 
  2863 \subsection{Configuring the code generator}
  2864 
  2865 When generating code for a complex term, the code generator recursively
  2866 calls itself for all subterms.
  2867 When it arrives at a constant, the default strategy of the code
  2868 generator is to look up its definition and try to generate code for it.
  2869 Constants which have no definitions that
  2870 are immediately executable, may be associated with a piece of ML
  2871 code manually using the \ttindex{consts_code} command
  2872 (see Fig.~\ref{fig:HOL:codegen-configuration}).
  2873 It takes a list whose elements consist of a constant (given in usual term syntax
  2874 -- an explicit type constraint accounts for overloading), and a
  2875 mixfix template describing the ML code. The latter is very much the
  2876 same as the mixfix templates used when declaring new constants.
  2877 The most notable difference is that terms may be included in the ML
  2878 template using antiquotation brackets \verb|{*|~$\ldots$~\verb|*}|.
  2879 A similar mechanism is available for
  2880 types: \ttindex{types_code} associates type constructors with
  2881 specific ML code. For example, the declaration
  2882 \begin{ttbox}
  2883 types_code
  2884   "*"     ("(_ */ _)")
  2885 
  2886 consts_code
  2887   "Pair"    ("(_,/ _)")
  2888 \end{ttbox}
  2889 in theory \texttt{Product_Type} describes how the product type of Isabelle/HOL
  2890 should be compiled to ML. Sometimes, the code associated with a
  2891 constant or type may need to refer to auxiliary functions, which
  2892 have to be emitted when the constant is used. Code for such auxiliary
  2893 functions can be declared using \texttt{attach}. For example, the
  2894 \texttt{wfrec} function from theory \texttt{Wellfounded_Recursion}
  2895 is implemented as follows:
  2896 \begin{ttbox}
  2897 consts_code
  2898   "wfrec"   ("\bs<module>wfrec?")
  2899 attach \{*
  2900 fun wfrec f x = f (wfrec f) x;
  2901 *\}
  2902 \end{ttbox}
  2903 If the code containing a call to \texttt{wfrec} resides in an ML structure
  2904 different from the one containing the function definition attached to
  2905 \texttt{wfrec}, the name of the ML structure (followed by a ``\texttt{.}'')
  2906 is inserted in place of ``\texttt{\bs<module>}'' in the above template.
  2907 The ``\texttt{?}'' means that the code generator should ignore the first
  2908 argument of \texttt{wfrec}, i.e.\ the termination relation, which is
  2909 usually not executable.
  2910 
  2911 Another possibility of configuring the code generator is to register
  2912 theorems to be used for code generation. Theorems can be registered
  2913 via the \ttindex{code} attribute. It takes an optional name as
  2914 an argument, which indicates the format of the theorem. Currently
  2915 supported formats are equations (this is the default when no name
  2916 is specified) and horn clauses (this is indicated by the name
  2917 \texttt{ind}). The left-hand sides of equations may only contain
  2918 constructors and distinct variables, whereas horn clauses must have
  2919 the same format as introduction rules of inductive definitions.
  2920 For example, the declaration
  2921 \begin{ttbox}
  2922 lemma Suc_less_eq [iff, code]: "(Suc m < Suc n) = (m < n)" \(\langle\ldots\rangle\)
  2923 lemma [code]: "((n::nat) < 0) = False" by simp
  2924 lemma [code]: "(0 < Suc n) = True" by simp
  2925 \end{ttbox}
  2926 in theory \texttt{Nat} specifies three equations from which to generate
  2927 code for \texttt{<} on natural numbers.
  2928 
  2929 \subsection{Specific HOL code generators}
  2930 
  2931 The basic code generator framework offered by Isabelle/Pure has
  2932 already been extended with additional code generators for specific
  2933 HOL constructs. These include datatypes, recursive functions and
  2934 inductive relations. The code generator for inductive relations
  2935 can handle expressions of the form $(t@1,\ldots,t@n) \in r$, where
  2936 $r$ is an inductively defined relation. If at least one of the
  2937 $t@i$ is a dummy pattern ``$_$'', the above expression evaluates to a
  2938 sequence of possible answers. If all of the $t@i$ are proper
  2939 terms, the expression evaluates to a boolean value.
  2940 \begin{small}
  2941 \begin{alltt}
  2942   theory Test = Lambda:
  2943 
  2944   code_module Test
  2945   contains
  2946     test1 = "Abs (Var 0) \(\circ\) Var 0 -> Var 0"
  2947     test2 = "Abs (Abs (Var 0 \(\circ\) Var 0) \(\circ\) (Abs (Var 0) \(\circ\) Var 0)) -> _"
  2948 \end{alltt}
  2949 \end{small}
  2950 In the above example, \texttt{Test.test1} evaluates to the boolean
  2951 value \texttt{true}, whereas \texttt{Test.test2} is a sequence whose
  2952 elements can be inspected using \texttt{Seq.pull} or similar functions.
  2953 \begin{ttbox}
  2954 ML \{* Seq.pull Test.test2 *\}  -- \{* This is the first answer *\}
  2955 ML \{* Seq.pull (snd (the it)) *\}  -- \{* This is the second answer *\}
  2956 \end{ttbox}
  2957 The theory
  2958 underlying the HOL code generator is described more detailed in
  2959 \cite{Berghofer-Nipkow:2002}. More examples that illustrate the usage
  2960 of the code generator can be found e.g.~in theories
  2961 \texttt{MicroJava/J/JListExample} and \texttt{MicroJava/JVM/JVMListExample}.
  2962 
  2963 \section{The examples directories}
  2964 
  2965 Directory \texttt{HOL/Auth} contains theories for proving the correctness of
  2966 cryptographic protocols~\cite{paulson-jcs}.  The approach is based upon
  2967 operational semantics rather than the more usual belief logics.  On the same
  2968 directory are proofs for some standard examples, such as the Needham-Schroeder
  2969 public-key authentication protocol and the Otway-Rees
  2970 protocol.
  2971 
  2972 Directory \texttt{HOL/IMP} contains a formalization of various denotational,
  2973 operational and axiomatic semantics of a simple while-language, the necessary
  2974 equivalence proofs, soundness and completeness of the Hoare rules with
  2975 respect to the denotational semantics, and soundness and completeness of a
  2976 verification condition generator.  Much of development is taken from
  2977 Winskel~\cite{winskel93}.  For details see~\cite{nipkow-IMP}.
  2978 
  2979 Directory \texttt{HOL/Hoare} contains a user friendly surface syntax for Hoare
  2980 logic, including a tactic for generating verification-conditions.
  2981 
  2982 Directory \texttt{HOL/MiniML} contains a formalization of the type system of
  2983 the core functional language Mini-ML and a correctness proof for its type
  2984 inference algorithm $\cal W$~\cite{milner78,nipkow-W}.
  2985 
  2986 Directory \texttt{HOL/Lambda} contains a formalization of untyped
  2987 $\lambda$-calculus in de~Bruijn notation and Church-Rosser proofs for $\beta$
  2988 and $\eta$ reduction~\cite{Nipkow-CR}.
  2989 
  2990 Directory \texttt{HOL/Subst} contains Martin Coen's mechanization of a theory
  2991 of substitutions and unifiers.  It is based on Paulson's previous
  2992 mechanisation in LCF~\cite{paulson85} of Manna and Waldinger's
  2993 theory~\cite{mw81}.  It demonstrates a complicated use of \texttt{recdef},
  2994 with nested recursion.
  2995 
  2996 Directory \texttt{HOL/Induct} presents simple examples of (co)inductive
  2997 definitions and datatypes.
  2998 \begin{itemize}
  2999 \item Theory \texttt{PropLog} proves the soundness and completeness of
  3000   classical propositional logic, given a truth table semantics.  The only
  3001   connective is $\imp$.  A Hilbert-style axiom system is specified, and its
  3002   set of theorems defined inductively.  A similar proof in ZF is described
  3003   elsewhere~\cite{paulson-set-II}.
  3004 
  3005 \item Theory \texttt{Term} defines the datatype \texttt{term}.
  3006 
  3007 \item Theory \texttt{ABexp} defines arithmetic and boolean expressions
  3008  as mutually recursive datatypes.
  3009 
  3010 \item The definition of lazy lists demonstrates methods for handling
  3011   infinite data structures and coinduction in higher-order
  3012   logic~\cite{paulson-coind}.%
  3013 \footnote{To be precise, these lists are \emph{potentially infinite} rather
  3014   than lazy.  Lazy implies a particular operational semantics.}
  3015   Theory \thydx{LList} defines an operator for
  3016   corecursion on lazy lists, which is used to define a few simple functions
  3017   such as map and append.   A coinduction principle is defined
  3018   for proving equations on lazy lists.
  3019   
  3020 \item Theory \thydx{LFilter} defines the filter functional for lazy lists.
  3021   This functional is notoriously difficult to define because finding the next
  3022   element meeting the predicate requires possibly unlimited search.  It is not
  3023   computable, but can be expressed using a combination of induction and
  3024   corecursion.  
  3025 
  3026 \item Theory \thydx{Exp} illustrates the use of iterated inductive definitions
  3027   to express a programming language semantics that appears to require mutual
  3028   induction.  Iterated induction allows greater modularity.
  3029 \end{itemize}
  3030 
  3031 Directory \texttt{HOL/ex} contains other examples and experimental proofs in
  3032 HOL.
  3033 \begin{itemize}
  3034 \item Theory \texttt{Recdef} presents many examples of using \texttt{recdef}
  3035   to define recursive functions.  Another example is \texttt{Fib}, which
  3036   defines the Fibonacci function.
  3037 
  3038 \item Theory \texttt{Primes} defines the Greatest Common Divisor of two
  3039   natural numbers and proves a key lemma of the Fundamental Theorem of
  3040   Arithmetic: if $p$ is prime and $p$ divides $m\times n$ then $p$ divides~$m$
  3041   or $p$ divides~$n$.
  3042 
  3043 \item Theory \texttt{Primrec} develops some computation theory.  It
  3044   inductively defines the set of primitive recursive functions and presents a
  3045   proof that Ackermann's function is not primitive recursive.
  3046 
  3047 \item File \texttt{cla.ML} demonstrates the classical reasoner on over sixty
  3048   predicate calculus theorems, ranging from simple tautologies to
  3049   moderately difficult problems involving equality and quantifiers.
  3050 
  3051 \item File \texttt{meson.ML} contains an experimental implementation of the {\sc
  3052     meson} proof procedure, inspired by Plaisted~\cite{plaisted90}.  It is
  3053   much more powerful than Isabelle's classical reasoner.  But it is less
  3054   useful in practice because it works only for pure logic; it does not
  3055   accept derived rules for the set theory primitives, for example.
  3056 
  3057 \item File \texttt{mesontest.ML} contains test data for the {\sc meson} proof
  3058   procedure.  These are mostly taken from Pelletier \cite{pelletier86}.
  3059 
  3060 \item File \texttt{set.ML} proves Cantor's Theorem, which is presented in
  3061   {\S}\ref{sec:hol-cantor} below, and the Schr{\"o}der-Bernstein Theorem.
  3062 
  3063 \item Theory \texttt{MT} contains Jacob Frost's formalization~\cite{frost93} of
  3064   Milner and Tofte's coinduction example~\cite{milner-coind}.  This
  3065   substantial proof concerns the soundness of a type system for a simple
  3066   functional language.  The semantics of recursion is given by a cyclic
  3067   environment, which makes a coinductive argument appropriate.
  3068 \end{itemize}
  3069 
  3070 
  3071 \goodbreak
  3072 \section{Example: Cantor's Theorem}\label{sec:hol-cantor}
  3073 Cantor's Theorem states that every set has more subsets than it has
  3074 elements.  It has become a favourite example in higher-order logic since
  3075 it is so easily expressed:
  3076 \[  \forall f::\alpha \To \alpha \To bool. \exists S::\alpha\To bool.
  3077     \forall x::\alpha. f~x \not= S 
  3078 \] 
  3079 %
  3080 Viewing types as sets, $\alpha\To bool$ represents the powerset
  3081 of~$\alpha$.  This version states that for every function from $\alpha$ to
  3082 its powerset, some subset is outside its range.  
  3083 
  3084 The Isabelle proof uses HOL's set theory, with the type $\alpha\,set$ and
  3085 the operator \cdx{range}.
  3086 \begin{ttbox}
  3087 context Set.thy;
  3088 \end{ttbox}
  3089 The set~$S$ is given as an unknown instead of a
  3090 quantified variable so that we may inspect the subset found by the proof.
  3091 \begin{ttbox}
  3092 Goal "?S ~: range\thinspace(f :: 'a=>'a set)";
  3093 {\out Level 0}
  3094 {\out ?S ~: range f}
  3095 {\out  1. ?S ~: range f}
  3096 \end{ttbox}
  3097 The first two steps are routine.  The rule \tdx{rangeE} replaces
  3098 $\Var{S}\in \texttt{range} \, f$ by $\Var{S}=f~x$ for some~$x$.
  3099 \begin{ttbox}
  3100 by (resolve_tac [notI] 1);
  3101 {\out Level 1}
  3102 {\out ?S ~: range f}
  3103 {\out  1. ?S : range f ==> False}
  3104 \ttbreak
  3105 by (eresolve_tac [rangeE] 1);
  3106 {\out Level 2}
  3107 {\out ?S ~: range f}
  3108 {\out  1. !!x. ?S = f x ==> False}
  3109 \end{ttbox}
  3110 Next, we apply \tdx{equalityCE}, reasoning that since $\Var{S}=f~x$,
  3111 we have $\Var{c}\in \Var{S}$ if and only if $\Var{c}\in f~x$ for
  3112 any~$\Var{c}$.
  3113 \begin{ttbox}
  3114 by (eresolve_tac [equalityCE] 1);
  3115 {\out Level 3}
  3116 {\out ?S ~: range f}
  3117 {\out  1. !!x. [| ?c3 x : ?S; ?c3 x : f x |] ==> False}
  3118 {\out  2. !!x. [| ?c3 x ~: ?S; ?c3 x ~: f x |] ==> False}
  3119 \end{ttbox}
  3120 Now we use a bit of creativity.  Suppose that~$\Var{S}$ has the form of a
  3121 comprehension.  Then $\Var{c}\in\{x.\Var{P}~x\}$ implies
  3122 $\Var{P}~\Var{c}$.   Destruct-resolution using \tdx{CollectD}
  3123 instantiates~$\Var{S}$ and creates the new assumption.
  3124 \begin{ttbox}
  3125 by (dresolve_tac [CollectD] 1);
  3126 {\out Level 4}
  3127 {\out {\ttlbrace}x. ?P7 x{\ttrbrace} ~: range f}
  3128 {\out  1. !!x. [| ?c3 x : f x; ?P7(?c3 x) |] ==> False}
  3129 {\out  2. !!x. [| ?c3 x ~: {\ttlbrace}x. ?P7 x{\ttrbrace}; ?c3 x ~: f x |] ==> False}
  3130 \end{ttbox}
  3131 Forcing a contradiction between the two assumptions of subgoal~1
  3132 completes the instantiation of~$S$.  It is now the set $\{x. x\not\in
  3133 f~x\}$, which is the standard diagonal construction.
  3134 \begin{ttbox}
  3135 by (contr_tac 1);
  3136 {\out Level 5}
  3137 {\out {\ttlbrace}x. x ~: f x{\ttrbrace} ~: range f}
  3138 {\out  1. !!x. [| x ~: {\ttlbrace}x. x ~: f x{\ttrbrace}; x ~: f x |] ==> False}
  3139 \end{ttbox}
  3140 The rest should be easy.  To apply \tdx{CollectI} to the negated
  3141 assumption, we employ \ttindex{swap_res_tac}:
  3142 \begin{ttbox}
  3143 by (swap_res_tac [CollectI] 1);
  3144 {\out Level 6}
  3145 {\out {\ttlbrace}x. x ~: f x{\ttrbrace} ~: range f}
  3146 {\out  1. !!x. [| x ~: f x; ~ False |] ==> x ~: f x}
  3147 \ttbreak
  3148 by (assume_tac 1);
  3149 {\out Level 7}
  3150 {\out {\ttlbrace}x. x ~: f x{\ttrbrace} ~: range f}
  3151 {\out No subgoals!}
  3152 \end{ttbox}
  3153 How much creativity is required?  As it happens, Isabelle can prove this
  3154 theorem automatically.  The default classical set \texttt{claset()} contains
  3155 rules for most of the constructs of HOL's set theory.  We must augment it with
  3156 \tdx{equalityCE} to break up set equalities, and then apply best-first search.
  3157 Depth-first search would diverge, but best-first search successfully navigates
  3158 through the large search space.  \index{search!best-first}
  3159 \begin{ttbox}
  3160 choplev 0;
  3161 {\out Level 0}
  3162 {\out ?S ~: range f}
  3163 {\out  1. ?S ~: range f}
  3164 \ttbreak
  3165 by (best_tac (claset() addSEs [equalityCE]) 1);
  3166 {\out Level 1}
  3167 {\out {\ttlbrace}x. x ~: f x{\ttrbrace} ~: range f}
  3168 {\out No subgoals!}
  3169 \end{ttbox}
  3170 If you run this example interactively, make sure your current theory contains
  3171 theory \texttt{Set}, for example by executing \ttindex{context}~{\tt Set.thy}.
  3172 Otherwise the default claset may not contain the rules for set theory.
  3173 \index{higher-order logic|)}
  3174 
  3175 %%% Local Variables: 
  3176 %%% mode: latex
  3177 %%% TeX-master: "logics-HOL"
  3178 %%% End: