diff --git "a/data_all_eng_slimpj/shuffled/split2/finalzzkzvs" "b/data_all_eng_slimpj/shuffled/split2/finalzzkzvs" new file mode 100644--- /dev/null +++ "b/data_all_eng_slimpj/shuffled/split2/finalzzkzvs" @@ -0,0 +1,5 @@ +{"text":"\\section*{\\centering\\LARGE{Acknowledgment}}\n\\thispagestyle{empty}\nI would like to acknowledge, and thank, the \\textit{Coordena\\c{c}\\~ao de Aperfei\\c{c}oamento de Pessoal de N\\'ivel Superior} (CAPES - Coordination for Improvement of Higher Education Personnel) for the PhD scholarship with process number $88882.329848\/2019-01$, given through the \\textit{Programa de Excel\\^encia Acad\\^emica} (PROEX - Academic Excellence Program), without which this work would not have been possible.\n\\vspace{5mm}\n\nI thank Professor Marcelo E. Coniglio for his advisement, for which I am forever grateful. While at times he only gently steered things in a better direction, be it a definition to be clearer or a theorem to be broader, at others he delved with me into research, when doing so seemed necessary. I firmly believe that, despite the thesis being my own, some developments that it lead to are as his as they are mine, probably more.\n\\vspace{5mm}\n\nOf utmost importance where the contributions by Professors Walter A. Carnielli and Hugo Mariano, who saw an earlier version of this work in my Ph.D. qualifying exam and suggested several improvements, as well as directions the ongoing research could take. If this thesis has not followed all of these directions, and it hasn't, it was only for lack of more time.\n\\vspace{5mm}\n\nMany other ameliorations and corrections, all of which I tried to incorporate into the final text, were brought forward in my thesis defense by Professors Walter Carnielli, Itala L. D'Ottaviano, H{\\'e}rcules A. Feitosa, Hitoshi Omori, Darllan C. Pinto and Anna Zamansky, who all in addition suggested invaluable ideas for future work.\n\\vspace{5mm}\n\nAnd to all those who have otherwise influenced this work, including all students, professors and staff of the \\textit{Centro de L\\'ogica, Epistemologia e Hist\\'oria da Ci\\^encia} (CLE - Center for Logic, Epistemology and History of Science), the \\textit{Instituto de Filosofia e Ci\\^encias Humanas} (IFCH - Institute of Philosophy and Human Sciences), and the \\textit{Instituto de Matem\\'atica e Estat\\'istica da Universidade de S\\~ao Paulo} (IME-USP - Institute of Mathematics and Statistics of the University of S\\~ao Paulo), I am also deeply grateful.\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\\newpage\n\\thispagestyle{empty}\n\\vspace*{\\fill}\n\\begin{flushright}\n\\textit{``But man is not made for defeat.\\\\\nA man can be destroyed but not defeated.``\\\\\n(\\textit{The Old Man and the Sea}, by Ernest Hemingway)}\n\\end{flushright}\n\\newpage\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\\section*{\\centering\\LARGE{Abstract}}\n\\thispagestyle{empty}\n\nThis work is divided between two main areas: in the theory of multialgebras, we focus mostly on a new definition of what a freely generated object should be in their category, and on how this category is equivalent to another with partially ordered algebras as objects; we then use non-deterministic semantics, specially those we have named restricted Nmatrices, on paraconsistent logics and some systems dealing with a new presentation of the natural concept of incompatibility, which generalizes inconsistency.\n\nIn algebra, we will focus on the non-deterministic ones, also known as multialgebras, whose operations return non-empty subsets of their universes. While the category of algebras over a signature has freely generated objects, which in a sense permit for the unique extension of functions to homomorphisms, the category of multialgebras over a given signature does not have elements with comparable properties. To circumvent this problem, we widen our understanding of algebras of formulas: if a multialgebra generalizes an algebra by having multiple results for a given operation, a multialgebra of formulas should generalize an algebra of formulas by having multiple possibilities for applying a connective to given formulas. Concerning the category of multialgebras itself, we offer an equivalence between it and a category avoiding non-determinism altogether, relying instead on ordered Boolean-like algebras as objects.\n\nOn the part devoted to logic, our goals are again roughly twofold: firstly, some logics of formal inconsistency, \\textit{exempli gratia} those found in da Costa's hierarchy, cannot be characterized by finite Nmatrices. In what is a very natural development, a restricted Nmatrix (or RNmatrix) restricts those homomorphisms to be taken into consideration when evaluating the validity of a deduction according to an Nmatrix. We show how this distinction gives far greater expressiveness to finite RNmatrices, enough to adequately characterize da Costa's systems and provide decision methods for those logics, both based on row-branching, row-eliminating truth tables, and tableau semantics. In another direction, we generalize logics of formal inconsistency to new systems built around the notion of incompatibility: the \\textit{Leitmotiv} being that having two incompatible formulas to simultaneously hold trivializes a deduction, and as a special case, a formula is consistent when it is incompatible with its negation. We show how this notion extends that of inconsistency in a non-trivial way, presenting conservative translations for many simple inconsistent systems into logics of incompatibility; we also provide semantics built on RNmatrices for these new logics, and prove that they can not be characterized by more standard methods.\n\n\\vspace*{\\fill}\n\\textbf{Keywords}: Non-classical mathematical logic; Paraconsistent logic; Algebraic logic; Inconsistency (Logic); Universal algebra.\n\n\\newpage\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\\section*{\\centering\\LARGE{Resumo}}\n\\thispagestyle{empty}\n\n\\foreignlanguage{portuguese}{\nEste trabalho est\\'a dividido entre duas grandes \\'areas: na teoria de multi\\'algebras, focamos majoritariamente em uma nova defini\\c{c}\\~ao do que um objeto livremente gerado deveria ser em sua categoria e em como esta categoria \\'e equivalente a outra com \\'algebras parcialmente ordenadas como objetos; ent\\~ao usamos sem\\^anticas n\\~ao-determin\\'isticas, especialmente aquela que nomeamos Nmatrizes restritas, nas l\\'ogicas paraconsistentes e em alguns sistemas lidando com uma nova apresenta\\c{c}\\~ao do conceito natural de incompatibilidade, que generaliza o conceito de inconsist\\^encia.}\n\n\\foreignlanguage{portuguese}{\nEm \\'algebra, nos focaremos nas n\\~ao-determin\\'isticas, tamb\\'em conhecidas como multi\\'algebras, cujas opera\\c{c}\\~oes retornam subconjuntos n\\~ao vazios de seus universos. Enquanto a categoria de \\'algebras sobre uma assinatura possui objetos livremente gerados, os quais permitem em certo sentido a extens\\~ao \\'unica de fun\\c{c}\\~oes a homomorfismos, a categoria de multi\\'algebras sobre uma assinatura dada n\\~ao possui elementos com propriedades compar\\'aveis. Para contornar este problema, estendemos o significado de uma \\'algebra de f\\'ormulas: se uma multi\\'algebra generaliza uma \\'algebra ao ter m\\'ultiplos resultados para uma dada opera\\c{c}\\~ao, uma multi\\'algebra de f\\'ormulas generaliza uma \\'algebra de f\\'ormulas ao ter m\\'ultiplas possibilidade para a aplica\\c{c}\\~ao de um conectivo a f\\'ormulas dadas. Quanto \\`a categoria de multi\\'algebras propriamente dita, oferecemos uma equival\\^encia entre ela e uma categoria livre de n\\~ao-determinismo, que alternativamente possui \\'algebras ordenadas, semelhantes a \\'algebras de Boole, como objetos.}\n\n\\foreignlanguage{portuguese}{\nNa parte dedicada \\`a l\\'ogica, nossos objetivos s\\~ao novamente dois: primeiramente, algumas l\\'ogicas de inconsist\\^encia formal, \\textit{exempli gratia} aquelas da hierarquia de da Costa, n\\~ao podem ser caracterizadas por Nmatrizes finitas. No que \\'e um desenvolvimento muito natural, uma Nmatriz restrita, ou RNmatriz, restringe aquelas homomorfismos que devem ser considerados quando testamos a validade de uma dedu\\c{c}\\~ao por uma Nmatriz. Mostramos como esta distin\\c{c}\\~ao prov\\^e as RNmatrizes finitas com poder expressivo muito superior, suficientente para adequadamente caracterizar os sistemas de da Costa e dar a eles m\\'etodos de decis\\~ao, tanto baseados em tabelas de verdade quanto em sem\\^anticas de tableaux. Em outra dire\\c{c}\\~ao, generalizamos as l\\'ogicas de inconsist\\^encia formal a sistemas constru\\'idos em torno da no\\c{c}\\~ao de incompatibilidade: o \\textit{Leitmotiv} sendo que duas f\\'ormulas incompat\\'iveis simultaneamente verdadeiras trivializam uma dedu\\c{c}\\~ao, e como um caso especial, uma f\\'ormula \\'e consistente quando \\'e incompat\\'ivel com sua nega\\c{c}\\~ao. Mostramos como essa no\\c{c}\\~ao estende aquela de inconsist\\^encia de maneira n\\~ao-trivial, apresentando tradu\\c{c}\\~oes conservativas para muitos dos sistemas inconsistentes mais simples em l\\'ogicas de incompatilidade, apresentamos sem\\^anticas constru\\'idas com RNmatrizes para essas novas l\\'ogicas e mostramos que elas n\\~ao podem ser caracterizadas por m\\'etodos mais usuais.}\n\n\\vspace*{\\fill}\n\\foreignlanguage{portuguese}{\n\\textbf{Palavras-chave}: L\\'ogica matem\\'atica n\\~ao cl\\'assica; L\\'ogica paraconsistente; L\\'ogica alg\\'ebrica; Inconsist\\^encia (L\\'ogica); \\'Algebra universal.}\n\n\\pagestyle{empty}\n\\addtocontents{toc}{\\protect\\thispagestyle{empty}}\n\\shorttoc{Abridged Table of Contents}{1}\n\\thispagestyle{empty}\n\\newpage\n\n\n\\shorttoc{Extended Table of Contents}{3}\n\\thispagestyle{empty}\n\\newpage\n\n\n\\tableofcontents\n\\cleardoublepage\n\n\n\n\\pagenumbering{arabic}\n\\setcounter{page}{16}\n\n\n\\newpage\n\\thispagestyle{plain}\n\\hspace{0pt}\n\\vfill\n\\begin{figure}[h]\n\\centering\n\\includegraphics[width=\\textwidth]{.\/Image1}\n\\caption*{Specimen of \\textit{Eupetomena macroura} resting, Monte Alegre do Sul, Brazil.\\\\Photographed by Guilherme Vicentin de Toledo, all rights reserved.}\n\\end{figure}\n\\vfill\n\\hspace{0pt}\n\n\n\n\n\n\n\\pagestyle{fancy}\n\n\\begin{refsegment}\n\\defbibfilter{notother}{not segment=\\therefsegment}\n\\chapter*{Introduction}\\label{Chapter0}\n\\chaptermark{Introduction}\n\\addcontentsline{toc}{chapter}{Introduction} \n\n \nThe research that makes up the bulk of this work may be divided in two grand areas, what explains our division of the thesis in two parts: algebra, corresponding to Part \\ref{Part1}, and logic, to Part \\ref{Part2}. \n\nIn algebra, we focus on the subject of multialgebras: we present a new concept of \\textit{weakly free multialgebras}, which we designed in order to offer a generalization of free algebras; furthermore, we show how the category of multialgebras, as usually defined in the context of non-deterministic semantics, can be presented alternatively as a category of algebras equipped with orders that are compatible with the underlying operations.\n\nIn logic, we will delve into paraconsistent logics, together with their generalizations, and non-deterministic semantics: we generalize non-deterministic matrices to \\textit{restricted non-deterministic matrices} (also known as RNmatrices), a semantical tool that uses both multialgebras and restrictions over the set of valuations to be considered, and apply this methodology to give an extend analysis of da Costa's hierarchy that includes new decision methods for its logics; we also give a new formalization of a natural concept, that of incompatibility, and treat the resulting logical systems again with RNmatrices.\n\n\\subsection*{Multialgebras}\n\nIn Part \\ref{Part1} we will deal mainly with multialgebras, also known as hyperalgebras or\\\\ non-deterministic algebras. Chapter \\ref{Chapter1} gives a brief introduction to the subject, broaching the definition of multialgebras, as well as those of homomorphisms between multialgebras, submultialgebras, the interpretation of formulas, and so on. We spend a few pages over alternative definitions for a homomorphism of multialgebras that allow for stronger representation theorems, and discuss briefly how these, sadly, are not very practical to use in non-deterministic semantics. Given its frequent use in our work, we also use this chapter to define, for completeness sake, lattices, Boolean algebras and Heyting algebras.\n\nChapter \\ref{Chapter2} offers an alternative solution to a classical problem: the category of algebras, on a given signature, possesses objects satisfying the universal mapping property, namely free algebras; meanwhile, the category of multialgebras over a signature $\\Sigma$, denoted by $\\textbf{MAlg}(\\Sigma)$, does not. Algebras of formulas over a signature $\\Sigma$ and variables $\\mathcal{V}$, that we denote by $\\textbf{F}(\\Sigma, \\mathcal{V})$, are then extended to structures that admit non-determinism: after all, if a multialgebra generalizes an algebra by allowing multiple results to an operation, a multialgebra of formulas should allow many formulas to be obtained from the application of a connective to given formulas. This new concept is shown to have many desirable characterizations, some inspired by linear algebra, others graph-theoretic in nature, motivating us to coin the nomenclature of weakly free multialgebras for them. They are shown to greatly simplify the proof that $\\textbf{MAlg}(\\Sigma)$ does not have free objects, and some other categorical considerations are then developed by the end of the chapter.\n\nConcerning the category $\\textbf{MAlg}(\\Sigma)$ itself, Chapter \\ref{Chapter3} proves this category is equivalent to another that avoids non-determinism altogether by considering ordered, Boolean-like algebras. Intuitively, one wants to capture both the operations and order of the natural algebra over the powerset of the universe $A$ of a multialgebra $\\mathcal{A}$, where: the order is the usual order for a powerset; and, given non-empty subsets $A_{1}$ trough $A_{n}$ of $A$, an $n$-ary operation $\\sigma$ on $(A_{1}, \\dotsc , A_{n})$ is given by the union of $\\sigma(a_{1}, \\dotsc , a_{n})$ for $(a_{1}, \\dotsc , a_{n})$ in $A_{1}\\times\\cdots\\times A_{n}$. This presents an alternative for those logicians wishing to stick to deterministic semantics without any loss of expressiveness, and can be generalized to offer equivalences for other categories with non-deterministic algebras, such as partial multialgebras.\n\n\n\n\\subsection*{Paraconsistent Logic}\n\nPart \\ref{Part2} involves paraconsistent logics, as well as non-deterministic semantics and broader systems. It starts with Chapter \\ref{Chapter4}, which provides a brief introduction to formal logic, and semantics of logical matrices, non-deterministic matrices (also known as Nmatrices) and restricted matrices. The high point of the chapter, however, is the definition of restricted non-deterministic matrices, also known as restricted Nmatrices or RNmatrices, which very naturally combine restricted and non-deterministic matrices: some theoretical considerations are made about these semantics, as well as a brief analysis of its previous, unrecognized uses in the literature. In essence, an RNmatrix is a triple $(\\mathcal{A}, D, \\mathcal{F})$, with $\\mathcal{A}$ a $\\Sigma$-multialgebra, $D$ a subset of its universe and $\\mathcal{F}$ a set of homomorphisms $\\nu:\\textbf{F}(\\Sigma, \\mathcal{V})\\rightarrow\\mathcal{A}$. Their motivation is to give finite semantics to logical systems that are not characterizable by finite Nmatrices, such as the logics of formal inconsistency ($\\textbf{LFI}'s$) between $\\textbf{mbCcl}$ and $\\textbf{Cila}$, and the whole hierarchy of da Costa.\n\nThe $C$-systems of da Costa, due to their complexity, are treated separately in Chapter \\ref{Chapter5}: in it, we start by defining these logics first devised in order to formalize the notion of inconsistency, or paraconsistency. We then proceed to provide RNmatrices of $n+2$ elements capable of characterizing each and every $C_{n}$ of the hierarchy, starting from $C_{2}$ to fix ideas: the intuition is that an RNmatrix for $C_{n}$ must have two classical values, standing from true and false, as well as $n$ inconsistent values, each standing for a different degree of inconsistency achieved in the logic; of course, this suggests that the $n$-th logic in the hierarchy could be regarded as an $n+2$-valued, non-deterministic logic. Furthermore, we show how these finite RNmatrices can be made into decision methods for da Costa's systems through both row-branching, row-eliminating truth tables, and tableau semantics.\n\nChapter \\ref{Chapter6} extends the RNmatrices for $C_{n}$ from the previous chapter to allow for model-theoretic considerations: while we started with $n+2$-valued restricted non-deterministic matrices, these can be shown to be constructed from the two-valued Boolean algebra as swap structures; the next logical step is to construct these swap structures over any Boolean algebras, leading to a class of RNmatrices capable itself of characterizing $C_{n}$. Further results include a brief combinatorial description of the snapshots found in these swap structures, as well as the construction of a category of swap structures for $C_{n}$, which is then proven to be isomorphic to the category of non-trivial Boolean algebras; this has important applications to the model theory of da Costa's hierarchy.\n\n\nChapter \\ref{Chapter7} introduces logics of incompatibility: a first attempt of formalizing the notion of incompatibility from natural language would be to declare two formulas incompatible if, and only if, together they trivialize a deduction. As it is done in logics of formal inconsistency, we weaken that condition to state that two incompatible formulas, a concept which may be primitive, are capable of trivializing an argument; we will denote by $\\alpha\\uparrow\\beta$ the fact that $\\alpha$ and $\\beta$ are incompatible. We start by defining some very basic systems of incompatibility, as of yet without negation, characterize then with RNmatrices and provide decision methods: of $\\bI^{-}$ we ask nothing, while $\\bI$ must have a commutative incompatibility connective and $\\bIpr$ propagates incompatibility under some conditions. We finish the chapter by discussing axioms that collapse $\\uparrow$ to its classical interpretation, and finally comparing our approach to a preexisting formalization of incompatibility, that of Brandom.\n\nIt is natural, once we have a connective for incompatibility, to consider its interplay with negation: the systems of Chapter \\ref{Chapter7} are devoid of negation, by design, but Chapter \\ref{Chapter8} consider those logics now equipped with negation and some axioms governing its interaction with incompatibility, most of them heavily inspired by the most common axioms for paraconsistent systems: while $\\nbI$ only adds a negation to $\\bI$ satisfying \\textit{tertium non datur}, $\\nbIciw$, $\\nbIci$ and $\\nbIcl$ generalize the logics of paraconsistency $\\mbCciw$, $\\mbCci$ and $\\mbCcl$, respectively. Of course, we then provide these systems with characterizing RNmatrices, as well as decision methods through row-branching, row-eliminating truth tables and tableau calculi, and explain why these semantics, instead of more classical ones, are necessary: not only our basic incompatible systems are not algebraizable by Blok and Pigozzi, they are also not characterizable by either finite Nmatrices or finite restricted matrices. We finish by analyzing the generalizations of logical matrices we have broached here, how do they relate to each other and to other possible generalizations.\n\n\n\nFinally, Chapter \\ref{Chapter9} studies an equivalence merely implied in previous chapters: in logics of formal inconsistency, a formula $\\alpha$ being consistent is recurrently expressed by $\\circ\\alpha$, where the connective $\\circ$ stands precisely for consistency. When dealing with incompatibility, the fact consistency is adequately reintroduced as incompatibility with negation becomes apparent: that is, $\\circ\\alpha$ may be viewed as $\\alpha\\uparrow\\neg\\alpha$. Accordingly, we define a function from the logics of formal inconsistency into the logics of incompatibility that is not only a translation, but a conservative one nonetheless. This seems to imply that our interpretation of incompatibility in logic strictly extends the notion of inconsistency in a non-trivial way.\n\n\n\\subsection*{Publications}\n\nA considerable portion of this thesis was made available as preprints in\n\n\\begin{enumerate}\n\n\\item \\fullcite{AbsFreeHyp};\n\n\\item \\fullcite{CostaRNmatrix};\n\n\\item \\fullcite{RestrictedSwap};\n\n\\item \\fullcite{Frominconsistency},\n\n\\end{enumerate}\n\nand published in\n\n\\begin{enumerate}\n\n\\item \\fullcite{WeaklyFreeMultialgebras};\n\n\\item \\fullcite{TwoDecisionProcedures}.\n\n\\end{enumerate}\n\n\n\n\\subsection*{Presentations}\n\nThe work found in this thesis has also lead to the following presentations.\n\n\\begin{enumerate}\n\n\\item ``\\textbf{RNmatrices for da Costa's hierarchy}'', in \\textit{I Enc(ue-o)ntro de L\\'ogica Brasil\\\\ Col(o-\\^o)mbia} (1st Meeting Brazil-Colombia in Logic), December of 2021.\n\n\\item ``\\textbf{Uma categoria de \\'algebras ordenadas equivalente a uma categoria de hiper\\'algebras}'', in \\textit{I Encontro Brasileiro em Teoria das Categorias}, January of 2021.\n\n\\item ``\\textbf{Sem\\^anticas n\\~ao-determin\\'isticas para l\\'ogicas n\\~ao-cl\\'assicas: Uma abordagem da\\\\ perspectiva de Teoria de Modelos e de \\'Algebra Universal}'', in \\textit{Encontro Brasileiro de L\\'ogica 2019} (19th Brazilian Logic Conference), May of 2019.\n\n\\item ``\\textbf{Sem\\^anticas n\\~ao-determin\\'isticas para l\\'ogicas n\\~ao-cl\\'assicas: Uma abordagem da\\\\ perspectiva de Teoria de Modelos e de \\'Algebra Universal}'', in \\textit{XVIII Encontro Nacional da ANPOF}, October of 2018.\n\n\\item ``\\textbf{Nondeterministic Semantics for Nonclassical Logics: An Approach from the\\\\ Perspective of Model Theory and Universal Algebra}'', in \\textit{Fourth Workshop CLE-Buenos Aires Logic Group}, April of 2018.\n\\end{enumerate}\n\n\n\\end{refsegment}\n\n\\part{Multialgebras}\\label{Part1}\n\n\n\n\n\\begin{refsegment}\n\\defbibfilter{notother}{not segment=\\therefsegment}\n\\setcounter{chapter}{0}\n\\chapter{$\\Sigma$-Multialgebras and lattices}\\label{Chapter1}\\label{Chapter 1}\n\nA multialgebra, or hyperalgebra\\index{Hyperalgebra}, is a generalization of the notion of algebra usually found in the context of universal algebras, see \\cite{Burris} for a standard approach to universal algebra. The main objective of such a generalization is to address the possibility that the outcome of an operation may be diffuse, that is, non-deterministic: we may have an idea of what the outcome should be, but not be certain about it. The first known appearance of multialgebras in literature may be found in \\cite{Marty}.\n\nA collection of disjoint sets $\\Sigma=\\{\\Sigma_{n}\\}_{n\\in\\mathbb{N}}$ indexed by $\\mathbb{N}$ will be called a signature\\index{Signature}\\label{signature}; the elements of the sets $\\Sigma_{n}$ will be called functional symbols of arity $n$, or $n$-ary functional symbols. For simplicity, we will denote $\\bigcup_{n\\in\\mathbb{N}}\\Sigma_{n}$ also by $\\Sigma$.\n\nA pair $\\mathcal{A}=(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma})$ is said to be, where $\\mathcal{P}(A)$\\label{powerset} denotes the powerset (also known as power set) of $A$:\n\\begin{enumerate}\n\\item a $\\Sigma$-algebra\\index{Algebra} if, for every $\\sigma\\in\\Sigma_{n}$, $\\sigma_{\\mathcal{A}}$ is a function of the form\n\\[\\sigma_{\\mathcal{A}}:A^{n}\\rightarrow A;\\]\n\\item a $\\Sigma$-multialgebra\\index{Multialgebra} if $A\\neq\\emptyset$ and, for every $\\sigma\\in\\Sigma_{n}$, $\\sigma_{\\mathcal{A}}$ is a function of the form\n\\[\\sigma_{\\mathcal{A}}:A^{n}\\rightarrow \\mathcal{P}(A)\\setminus\\{\\emptyset\\};\\]\n\\end{enumerate}\n\nThe set $A$ is called the universe\\index{Universe} of $\\mathcal{A}$. \n\nGiven a $\\Sigma$-algebra $\\mathcal{A}=(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma})$, one can always define the $\\Sigma$-multialgebra $\\overline{\\mathcal{A}}=(A, \\{\\sigma_{\\overline{\\mathcal{A}}}\\}_{\\sigma\\in\\Sigma})$ such that, for $\\sigma\\in\\Sigma_{n}$ and $a_{1}, \\dotsc , a_{n}\\in A$, \n\\[\\sigma_{\\overline{\\mathcal{A}}}(a_{1}, \\dotsc , a_{n})=\\{\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})\\};\\]\nit is clear that $\\overline{\\mathcal{A}}$ carries the same information that $\\mathcal{A}$, and so one can see $\\Sigma$-algebras as $\\Sigma$-multialgebras. For most of our studies here we will focus mainly on multialgebras.\n\n\n\n\n\n\\section{Homomorphisms}\n\n\n\\subsection{Single-valued homomorphisms}\n\nGiven $\\Sigma$-algebras $\\mathcal{A}=(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma})$ and $\\mathcal{B}=(B, \\{\\sigma_{\\mathcal{B}}\\}_{\\sigma\\in\\Sigma})$, a function $\\varphi:A\\rightarrow B$ is said to be a homomorphism from $\\mathcal{A}$ to $\\mathcal{B}$ if, for every $\\sigma\\in\\Sigma_{n}$ and $a_{1}, \\dotsc , a_{n}\\in A$ we have that\n\\[\\varphi(\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n}))=\\sigma_{\\mathcal{B}}(\\varphi(a_{1}), \\dotsc , \\varphi(a_{n})).\\]\n\nGiven $\\Sigma$-multialgebras $\\mathcal{A}=(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma})$ and $\\mathcal{B}=(B, \\{\\sigma_{\\mathcal{B}}\\}_{\\sigma\\in\\Sigma})$, we would like to define once again a morphism between the two of them. But in the case of multialgebras, one can come up with many possible straightforward definitions of homomorphisms, all of them starting with a function $\\varphi:A\\rightarrow B$. The following two definitions are the most useful to our purposes:\n\\begin{enumerate}\n\\item if, for all $\\sigma\\in\\Sigma_{n}$ and $a_{1}, \\dotsc , a_{n}\\in A$, \n\\[\\{\\varphi(a)\\ :\\ a\\in\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})\\}\\subseteq \\sigma_{\\mathcal{B}}(\\varphi(a_{1}), \\dotsc , \\varphi(a_{n})),\\]\nwe call $\\varphi$ a homomorphism\\index{Homomorphism}, or $\\Sigma$-homomorphism;\n\\item we call $\\varphi$ a full homomorphism\\index{Homomorphism, Full} if the above condition is replaced by \n\\[\\{\\varphi(a)\\ :\\ a\\in\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})\\}=\\sigma_{\\mathcal{B}}(\\varphi(a_{1}), \\dotsc , \\varphi(a_{n})).\\]\n\\end{enumerate}\nIf the function $\\varphi:A\\rightarrow B$ is a homomorphism from $\\mathcal{A}$ to $\\mathcal{B}$, we will simply write $\\varphi:\\mathcal{A}\\rightarrow\\mathcal{B}$\\label{homomorphism}.\n\n\\begin{theorem}\nThe class of all $\\Sigma$-multialgebras becomes a category $\\textbf{MAlg}(\\Sigma)$\\label{MAlg} or $\\textbf{MAlg}_{=}(\\Sigma)$\\label{MAlg=} when the set of morphisms between two $\\Sigma$-multialgebras $\\mathcal{A}$ and $\\mathcal{B}$ is, respectively:\n\\begin{enumerate}\n\\item the set of all homomorphisms between $\\mathcal{A}$ and $\\mathcal{B}$;\n\\item the set of all full homomorphisms between $\\mathcal{A}$ and $\\mathcal{B}$.\n\\end{enumerate}\nIn both cases, the composition of morphisms is the usual composition of functions.\n\\end{theorem}\n\n\\begin{proof}\nWe must show that all two of those alleged categories have identity morphisms and that their compositions are well-defined, meaning that composing two morphisms returns again a morphism; clearly there is no need to show the associativity of composition, since it is know that the composition of functions is indeed associative.\n\nFor every multialgebra $\\mathcal{A}$ we consider the morphism $Id_{\\mathcal{A}}:\\mathcal{A}\\rightarrow \\mathcal{A}$ given by, for every $a\\in A$, $Id_{\\mathcal{A}}(a)=a$. This morphism is the desired identity morphism for $\\mathcal{A}$ in both categories, since: it is, in fact, a morphism in the two of them, given that it is a full homomorphism and therefore also a homomorphism; it is the identity for the composition of functions and therefore, for any morphisms $\\varphi:\\mathcal{A}\\rightarrow \\mathcal{B}$ and $\\psi:\\mathcal{C}\\rightarrow\\mathcal{A}$, \n\\[\\varphi\\circ Id_{\\mathcal{A}}=\\varphi\\quad\\text{and}\\quad Id_{\\mathcal{A}}\\circ\\psi=\\psi.\\]\n\nNow to prove that the composition is well-defined, fix $\\sigma\\in\\Sigma_{n}$ and $a_{1}, \\dotsc , a_{n}\\in A$.\n\\begin{enumerate}\n\\item If $\\varphi:\\mathcal{A}\\rightarrow\\mathcal{B}$ and $\\psi:\\mathcal{B}\\rightarrow\\mathcal{C}$ are simply homomorphisms,\n\\[\\{\\varphi(a)\\ :\\ a\\in\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})\\}\\subseteq\\sigma_{\\mathcal{B}}(\\varphi(a_{1}), \\dotsc , \\varphi(a_{n}))\\] \nand therefore \n\\[\\{\\psi\\circ\\varphi(a)\\ :\\ a\\in\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})\\}\\subseteq\\{\\psi(b)\\ :\\ b\\in \\sigma_{\\mathcal{B}}(\\varphi(a_{1}), \\dotsc , \\varphi(a_{n}))\\};\\]\nsince $\\psi$ is a homomorphism, we have that\n\\[\\{\\psi(b)\\ :\\ b\\in \\sigma_{\\mathcal{B}}(\\varphi(a_{1}), \\dotsc , \\varphi(a_{n}))\\}\\subseteq \\sigma_{\\mathcal{C}}(\\psi\\circ\\varphi(a_{1}), \\dotsc , \\psi\\circ\\varphi(a_{n})),\\]\nand from that $\\psi\\circ\\varphi$ is also a homomorphism.\n\n\\item If $\\varphi:\\mathcal{A}\\rightarrow\\mathcal{B}$ and $\\psi:\\mathcal{B}\\rightarrow\\mathcal{C}$ are full homomorphisms, is enough to replace all \"$\\subseteq$\" on the proof above by equalities to obtain a proof that $\\psi\\circ\\varphi$ is also a full homomorphism.\n\\end{enumerate}\n\\end{proof}\n\n\\begin{definition}\nA full homomorphism $\\varphi:\\mathcal{A}\\rightarrow\\mathcal{B}$ is said to be an isomorphism\\index{Isomorphism} if $\\varphi:A\\rightarrow B$ is a bijection.\\footnote{We do not attempt to define isomorphisms for homomorphisms that are not full since the latter class is not closed under inverses: the inverse of a non-full homomorphism is an antihomomorphism, which satisfies $\\sigma_{\\mathcal{B}}(\\varphi(a_{1}), \\dotsc, \\varphi(a_{n}))\\subseteq \\{\\varphi(a)\\ :\\ a\\in\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc a_{n})\\}$; considering antihomomorphisms does not make the underlying theory uninteresting, but it does make the theory much harder.}\n\\end{definition}\n\n\\begin{proposition}\nLet $\\mathcal{A}$ and $\\mathcal{B}$ be $\\Sigma$-multialgebras and $\\varphi:A\\rightarrow B$ be a bijective function with inverse $\\psi:B\\rightarrow A$: if $\\varphi$ is a full homomorphism, so is $\\psi$.\n\\end{proposition}\n\n\\begin{proof}\nLet $\\sigma\\in\\Sigma_{n}$, $b_{1}, \\dotsc , b_{n}\\in B$ and $a_{1}=\\psi(b_{1}), \\dotsc , a_{n}=\\psi(b_{n})$, so that $b_{1}=\\varphi(a_{1}), \\dotsc , b_{n}=\\varphi(a_{n})$. We have that\n\\[ \\{\\psi(b)\\ :\\ b\\in \\sigma_{\\mathcal{B}}(b_{1}, \\dotsc , b_{n})\\}=\\{\\psi(b)\\ :\\ b\\in\\sigma_{\\mathcal{B}}(\\varphi(a_{1}), \\dotsc , \\varphi(a_{n}))\\}=\\]\n\\[\\{\\psi(b)\\ :\\ b\\in\\{\\varphi(a)\\ :\\ a\\in\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})\\}\\}=\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})=\\sigma_{\\mathcal{A}}(\\psi(b_{1}), \\dotsc , \\psi(b_{n})),\\]\nand therefore $\\psi$ is indeed a full homomorphism.\n\\end{proof}\n\n\n\n\n\n\\subsection{Multi-valued homomorphisms}\\label{Multi-valued homomorphisms}\n\nOne natural consideration when defining morphisms between multialgebras is that, if the operations are of a non-deterministic nature, perhaps the morphisms should be as well.\n\nAlthough there exist a plethora of possible definitions in this case, we will mention only two of them, even avoiding considerations about isomorphisms for they are out of scope for this text. Given a signature $\\Sigma$ and $\\Sigma$-multialgebras $\\mathcal{A}=(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma})$ and $\\mathcal{B}=(B, \\{\\sigma_{\\mathcal{B}}\\}_{\\sigma\\in\\Sigma})$, a function $\\varphi:A\\rightarrow\\mathcal{P}(B)\\setminus\\{\\emptyset\\}$ is said to be a multihomomorphism\\index{Multihomomorphism} if, for all $\\sigma\\in\\Sigma_{n}$ and $a_{1}, \\dotsc , a_{n}\\in A$,\n\\[\\bigcup_{a\\in \\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})}\\varphi(a)\\subseteq\\bigcup_{(b_{1}, \\dotsc , b_{n})\\in \\varphi(a_{1})\\times\\cdots\\times\\varphi(a_{n})}\\sigma_{\\mathcal{B}}(b_{1}, \\dotsc , b_{n});\\]\nthe same function is said to be a full multihomomorphism\\index{Multihomomorphism, Full} if this conditions is replaced by \n\\[\\bigcup_{a\\in \\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})}\\varphi(a)=\\bigcup_{(b_{1}, \\dotsc , b_{n})\\in \\varphi(a_{1})\\times\\cdots\\times\\varphi(a_{n})}\\sigma_{\\mathcal{B}}(b_{1}, \\dotsc , b_{n}).\\]\nWe will denote a multihomomorphism $\\varphi$ between $\\mathcal{A}$ and $\\mathcal{B}$ simply by $\\varphi:\\mathcal{A}\\rightarrow\\mathcal{B}$.\n\n\\begin{lemma}\\label{lemma about unions}\nFor an $n\\in\\mathbb{N}$, take sets $X_{1}, \\dotsc , X_{n}$; for $X=\\bigcup_{i=1}^{n}X_{i}$, let $\\{Y_{x}\\}_{x\\in X}$ be a family indexed by $X$; then\n\\[\\bigcup_{(x_{1}, \\dotsc , x_{n})\\in X_{1}\\times\\cdots\\times X_{n}} Y_{x_{1}}\\times\\cdots\\times Y_{x_{n}}\\subseteq\\bigcup_{x_{1}\\in X_{1}}Y_{x_{1}}\\times\\cdots\\times\\bigcup_{x_{n}\\in X_{n}}Y_{x_{n}}.\\]\nIf $n=1$, we have instead an equality.\n\\end{lemma}\n\n\\begin{proof}\nSuppose $(y_{1}, \\dotsc , y_{n})$ belongs to the left side of the inequality, and there must exist\\\\ $(x_{1}, \\dotsc , x_{n})$ in $X_{1}\\times\\cdots\\times X_{n}$ such that $(y_{1}, \\dotsc , y_{n})\\in Y_{x_{1}}\\times\\cdots\\times Y_{x_{n}}$.\n\nNow, for any $i\\in\\{1, \\dotsc , n\\}$, since $y_{i}\\in Y_{x_{i}}$ for an $x_{i}\\in X_{i}$, $y_{i}\\in \\bigcup_{x\\in X_{i}}Y_{x}$. It follows that $(y_{1}, \\dotsc , y_{n})\\in\\bigcup_{x_{1}\\in X_{1}}Y_{x_{1}}\\times\\cdots\\times\\bigcup_{x_{n}\\in X_{n}}Y_{x_{n}}$.\n\nThe equality if $n=1$ is trivial.\n\\end{proof}\n\n\n\n\\begin{theorem}\nThe class of all $\\Sigma$-multialgebras becomes a category $\\textbf{MMAlg}(\\Sigma)$\\label{MMAlg} or $\\textbf{MMAlg}_{=}(\\Sigma)$\\label{MMAlg=} when the set of morphisms between two $\\Sigma$-multialgebras $\\mathcal{A}$ and $\\mathcal{B}$ is, respectively:\n\\begin{enumerate}\n\\item the set of all multihomomorphisms between $\\mathcal{A}$ and $\\mathcal{B}$;\n\\item the set of all full multihomomorphisms between $\\mathcal{A}$ and $\\mathcal{B}$.\n\\end{enumerate}\nIn both cases, the composition $\\psi\\circ\\varphi$ of multihomomorphisms $\\varphi:\\mathcal{A}\\rightarrow\\mathcal{B}$ and $\\psi:\\mathcal{B}\\rightarrow\\mathcal{C}$ is given by, on an element $a\\in A$, $\\psi\\circ\\varphi(a)=\\bigcup_{b\\in\\varphi(a)}\\psi(b)$.\n\\end{theorem}\n\n\\begin{proof}\nWe must show the existence of identity morphisms and that the composition of morphisms is well-defined and associative.\n\nFor every $\\Sigma$-multialgebra $\\mathcal{A}$ we consider the morphism $Id_{\\mathcal{A}}:\\mathcal{A}\\rightarrow\\mathcal{A}$ given by, for every $a\\in A$, $Id_{\\mathcal{A}}(a)=\\{a\\}$. It is a full multihomomorphism, and therefore also a multihomomorphism, given that, for $\\sigma\\in\\Sigma_{n}$ and $a_{1}, \\dotsc , a_{n}\\in A$,\n\\[\\bigcup_{a\\in \\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})}Id_{\\mathcal{A}}(a)=\\bigcup_{a\\in\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})}\\{a\\}=\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})=\\bigcup_{(b_{1}, \\dotsc , b_{n})\\in \\{a_{1}\\}\\times\\cdots\\times\\{a_{n}\\}}\\sigma_{\\mathcal{A}}(b_{1}, \\dotsc , b_{n})=\\]\n\\[\\bigcup_{(b_{1}, \\dotsc , b_{n})\\in Id_{\\mathcal{A}}(a_{1})\\times\\cdots\\times Id_{\\mathcal{A}}(a_{n})}\\sigma_{\\mathcal{A}}(b_{1}, \\dotsc , b_{n}).\\]\nFor any multihomomorphisms $\\varphi:\\mathcal{A}\\rightarrow\\mathcal{B}$ and $\\psi:\\mathcal{C}\\rightarrow\\mathcal{A}$ and elements $a\\in A$ and $c\\in C$, we have that\n\\[\\varphi\\circ Id_{\\mathcal{A}}(a)=\\bigcup_{d\\in Id_{\\mathcal{A}}(a)}\\varphi(d)=\\bigcup_{d\\in \\{a\\}}\\varphi(d)=\\varphi(a)\\]\nand\n\\[Id_{\\mathcal{A}}\\circ\\psi(c)=\\bigcup_{e\\in\\psi(c)}Id_{\\mathcal{A}}(e)=\\bigcup_{e\\in\\psi(c)}\\{e\\}=\\psi(c),\\]\nso that $\\varphi\\circ Id_{\\mathcal{A}}=\\varphi$ and $Id_{\\mathcal{A}}\\circ\\psi=\\psi$, and $Id_{\\mathcal{A}}$ is indeed an identity.\n\nTo see that the composition is well-defined, fix $\\sigma\\in\\Sigma$ and $a_{1}, \\dotsc , a_{n}\\in A$.\n\n\\begin{enumerate}\n\\item If $\\varphi:\\mathcal{A}\\rightarrow\\mathcal{B}$ and $\\psi:\\mathcal{B}\\rightarrow\\mathcal{C}$ are multihomomorphisms, \n\\[\\bigcup_{a\\in \\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})}\\psi\\circ\\varphi(a)=\\bigcup_{a\\in \\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})}\\bigcup_{b\\in \\varphi(a)}\\psi(b);\\]\nsince \n\\[\\bigcup_{a\\in \\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})}\\varphi(a)\\subseteq \\bigcup_{(b_{1}, \\dotsc , b_{n})\\in\\varphi(a_{1})\\times\\cdots\\times\\varphi(a_{n})}\\sigma_{\\mathcal{B}}(b_{1}, \\dotsc , b_{n}),\\]\nwe have that the last set on the equality is contained in \n\\[\\bigcup_{(b_{1}, \\dotsc , b_{n})\\in \\varphi(a_{1})\\times\\cdots\\times\\varphi(a_{n})}\\bigcup_{b\\in \\sigma_{\\mathcal{B}}(b_{1}, \\dotsc , b_{n})}\\psi(b)\\subseteq\\bigcup_{(b_{1}, \\dotsc , b_{n})\\in \\varphi(a_{1})\\times\\cdots\\times\\varphi(a_{n})}\\bigcup_{(c_{1}, \\dotsc , c_{n})\\in \\psi(b_{1})\\times\\cdots\\times\\psi(b_{n})}\\sigma_{\\mathcal{C}}(c_{1}, \\dotsc , c_{n});\\]\nby Lemma \\ref{lemma about unions}, for $X_{i}=\\varphi(a_{i})$ and $Y_{x}=\\psi(x)$, this last set is contained in\n\\[\\bigcup_{(c_{1}, \\dotsc , c_{n})\\in \\bigcup_{b_{1}\\in \\varphi(a_{1})}\\psi(b_{1})\\times\\cdots\\times\\bigcup_{b_{n}\\in \\varphi(a_{n})}\\psi(b_{n})}\\sigma_{\\mathcal{C}}(c_{1}, \\dotsc , c_{n})=\\bigcup_{(c_{1}, \\dotsc , c_{n})\\in \\psi\\circ\\varphi(a_{1})\\times\\cdots\\times\\psi\\circ\\varphi(a_{n})}\\sigma_{\\mathcal{C}}(c_{1}, \\dotsc , c_{n}),\\]\nand from that $\\psi\\circ\\varphi$ is also a multihomomorphism.\n\n\\item If $\\varphi:\\mathcal{A}\\rightarrow\\mathcal{B}$ and $\\psi:\\mathcal{B}\\rightarrow\\mathcal{C}$ are full multihomomorphisms, is enough to replace all occurrences of \"$\\subseteq$\" in the proof above by \"$=$\" to obtain a proof that $\\psi\\circ\\varphi$ is also a full multihomomorphism.\n\\end{enumerate}\n\nFinally, it remains to be proved that such a notion of composition is associative: let $\\varphi:\\mathcal{A}\\rightarrow\\mathcal{B}$, $\\psi:\\mathcal{B}\\rightarrow\\mathcal{C}$ and $\\theta:\\mathcal{C}\\rightarrow\\mathcal{D}$ be multihomomorphisms and $a\\in A$; then, using Lemma \\ref{lemma about unions} again, we have that\n\\[[\\theta\\circ(\\psi\\circ\\varphi)](a)=\\bigcup_{c\\in \\psi\\circ\\varphi(a)}\\theta(c)=\\bigcup_{c\\in \\bigcup_{b\\in\\varphi(a)}\\psi(b)}\\theta(c)=\\bigcup_{b\\in\\varphi(a)}\\bigcup_{c\\in\\psi(b)}\\theta(c)=\\]\n\\[\\bigcup_{b\\in\\varphi(a)}\\theta\\circ\\psi(b)=[(\\theta\\circ\\psi)\\circ\\varphi](a)\\]\n\\end{proof}\n\nWe can arrange the categories that have so far appeared in our study of multialgebras in the following helpful diagram.\n\n\\[ \\begin{tikzcd}\n\\textbf{MAlg}_{=}(\\Sigma)\\arrow[hook]{dd}\\arrow{rr}{J_{=}} & & \\textbf{MMAlg}_{=}(\\Sigma)\\arrow[hook]{dd}\\\\\n&&\\\\\n\\textbf{MAlg}(\\Sigma) \\arrow{rr}{J} & & \\textbf{MMAlg}(\\Sigma)\n\\end{tikzcd}\n\\]\n\nIt is clear that $\\textbf{MAlg}_{=}(\\Sigma)$ is a subcategory of $\\textbf{MAlg}(\\Sigma)$, and that $\\textbf{MMAlg}_{=}(\\Sigma)$ is a subcategory of $\\textbf{MMAlg}(\\Sigma)$, all of these four categories having as objects the class of all $\\Sigma$-multialgebras. And, although $\\textbf{MAlg}_{=}(\\Sigma)$ is not a subcategory of $\\textbf{MMAlg}_{=}(\\Sigma)$, nor is $\\textbf{MAlg}(\\Sigma)$ a subcategory of $\\textbf{MMAlg}(\\Sigma)$, we can easily define functors \n\\[J_{=}:\\textbf{MAlg}_{=}(\\Sigma)\\rightarrow \\textbf{MMAlg}_{=}(\\Sigma)\\quad\\text{and}\\quad J:\\textbf{MAlg}(\\Sigma)\\rightarrow \\textbf{MMAlg}(\\Sigma)\\]\nsuch that the image of $J$ (respectively $J_{=}$) is a subcategory of $\\textbf{MMAlg}(\\Sigma)$ ($\\textbf{MMAlg}_{=}(\\Sigma)$) isomorphic to $\\textbf{MAlg}(\\Sigma)$ ($\\textbf{MAlg}_{=}(\\Sigma)$); we define $J$ ($J_{=}$) as the identity on objects, and for a (full) homomorphism $\\varphi:\\mathcal{A}\\rightarrow\\mathcal{B}$, the (full) multihomomorphism $J\\varphi$ (respectively $J_{=}\\varphi$) from $\\mathcal{A}$ to $\\mathcal{B}$, on an element $a$ of $\\mathcal{A}$, is just $\\{\\varphi(a)\\}$.\n\n\n\n\\section{Formulas and how to interpret them}\\label{Formulas and how to interpret them}\n\n\\begin{definition}\nGiven a set $\\mathcal{V}$, whose elements we will call propositional variables, we define the formulas\\index{Formula} over the signature $\\Sigma$ on the variables $\\mathcal{V}$ by recursion, and only by the following rules:\n\\begin{enumerate}\n\\item all elements of $\\mathcal{V}$ are formulas;\n\\item all elements of $\\Sigma_{0}$ are formulas;\n\\item if $\\sigma\\in\\Sigma_{n}$ and $\\alpha_{1}, \\dotsc , \\alpha_{n}$ are formulas, $\\sigma(\\alpha_{1}, \\dotsc , \\alpha_{n})$ is a formula.\n\\end{enumerate}\n\\end{definition}\n\nHowever, the expression $\\sigma(\\alpha_{1}, \\dotsc , \\alpha_{n})$ is not entirely formally defined: in what would be the correct, formal definition, yet not as clear, a formula over the signature $\\Sigma$ on the variables $\\mathcal{V}$ is any function $f:I_{k}\\rightarrow\\mathcal{V}\\cup\\bigcup_{n\\in\\mathbb{N}}\\Sigma_{n}$, where $I_{k}=\\{0, \\dotsc , k\\}$ is a initial segment of $\\mathbb{N}$, such that:\n\\begin{enumerate}\n\\item either $k=0$ and $f(0)\\in\\mathcal{V}\\cup\\Sigma_{0}$;\n\\item or there exist $m\\in\\mathbb{N}\\setminus\\{0\\}$, $\\sigma\\in\\Sigma_{m}$ and formulas \n\\[f_{1}:I_{k_{1}}\\rightarrow \\mathcal{V}\\cup\\bigcup_{n\\in\\mathbb{N}}\\Sigma_{n}\\quad\\text{through}\\quad f_{m}:I_{k_{m}}\\rightarrow\\mathcal{V}\\cup\\bigcup_{n\\in\\mathbb{N}}\\Sigma_{n}\\]\nsuch that $k=1+\\sum_{j=1}^{m}k_{j}$, $f(0)=\\sigma$ and\n\\[\\text{$f(i-k_{p}+\\sum_{j=1}^{p}k_{j})=f_{p}(i)$, for $p\\in\\{1, \\dotsc , m\\}$ and $i\\in I_{k_{p}}$.}\\]\n\\end{enumerate}\n\nThe set of all formulas over the signature $\\Sigma$ on the variables $\\mathcal{V}$ is denoted by $F(\\Sigma, \\mathcal{V})$\\label{FSigmaV}. If we define, for a $\\sigma\\in\\Sigma_{n}$, the function $\\sigma_{\\textbf{F}(\\Sigma, \\mathcal{V})}:F(\\Sigma, \\mathcal{V})^{n}\\rightarrow F(\\Sigma, \\mathcal{V})$ to be, for formulas $\\alpha_{1}, \\dotsc , \\alpha_{n}$,\n\\[\\sigma_{\\textbf{F}(\\Sigma, \\mathcal{V})}(\\alpha_{1}, \\dotsc , \\alpha_{n})=\\sigma(\\alpha_{1}, \\dotsc , \\alpha_{n}),\\]\nwe obtain a $\\Sigma$-algebra \n\\[\\textbf{F}(\\Sigma, \\mathcal{V})=(F(\\Sigma, \\mathcal{V}), \\{\\sigma_{\\textbf{F}(\\Sigma, \\mathcal{V})}\\}_{\\sigma\\in\\Sigma}),\\]\n\\label{tFSigmaV}said to be the free $\\Sigma$-algebra on the variables $\\mathcal{V}$, or the $\\Sigma$-algebra of formulas on the variables $\\mathcal{V}$.\n\nWe will also use $\\textbf{F}(\\Sigma, \\mathcal{V})$ to denote the corresponding $\\Sigma$-multialgebra.\n\n\\begin{definition}\nWe define the order\\index{Order of a formula}\\label{order}, or complexity\\index{Complexity of a formula}, $|\\alpha|$ of a formula $\\alpha$ in $F(\\Sigma, \\mathcal{V})$ as:\n\\begin{enumerate}\n\\item $0$, if $\\alpha$ is an element of $\\mathcal{V}$;\n\\item $0$, if $\\alpha$ is an element of $\\Sigma_{0}$;\n\\item if $\\alpha$ is of the form $\\sigma(\\alpha_{1}, \\dotsc , \\alpha_{n})$, \n\\[|\\alpha|=1+\\max_{1\\leq i\\leq n}|\\alpha_{i}|.\\]\n\\end{enumerate}\n\\end{definition}\n\nNow, we have given the simplest definition of formula: when we approach choice-dependent freely generated multialgebras, we will find what one can identify as a broader definition of formula, but with which we shall not deal with until later. \n\n\n\\subsection{Different notions of interpretation}\n\nPerhaps more important than defining what is a formula is interpreting this formula, and here the non-determinism of multialgebras once again gives us an array of different notions.\n\n\\begin{definition}\nGiven a $\\Sigma$-multialgebra $\\mathcal{A}$, a homomorphism $\\nu:\\textbf{F}(\\Sigma, \\mathcal{V})\\rightarrow \\mathcal{A}$ is called a legal valuation.\\index{Valuation, Legal}\n\\end{definition}\n\nThe notion of legal valuation is one attempt to give an interpretation of a formula $\\alpha$: in this case, the formula $\\alpha$, under the legal valuation $\\nu$, takes the value $\\nu(\\alpha)$.\n\nA serious problem that arises from the notion of legal valuation is that one interpretation of the variables can lead to several legal valuations, and therefore different interpretations of a formula. This does not occur to $\\Sigma$-algebras, where one interpretation of the variables implies an unique interpretation of the formulas.\n\n\\begin{example}\nTake the signature $\\Sigma$ with $\\Sigma_{1}=\\{s\\}$ and $\\Sigma_{n}=\\emptyset$ for $n\\neq 1$; take the $\\Sigma$-multialgebras $\\textbf{F}(\\Sigma, \\mathcal{V})$, for $\\mathcal{V}=\\{x\\}$, and $\\mathcal{A}=(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma})$ such that $A=\\{0,1\\}$ and $s_{\\mathcal{A}}(0)=s_{\\mathcal{A}}(1)=\\{0,1\\}$.\n\\begin{figure}[H]\n\\centering\n\\begin{minipage}[t]{4cm}\n\\centering\n\\begin{tikzcd}[row sep=1cm]\n\\vdots\\\\\ns^{2}(x)\\arrow{u}{s_{\\textbf{F}(\\Sigma, \\mathcal{V})}}\\\\\ns(x)\\arrow{u}{s_{\\textbf{F}(\\Sigma, \\mathcal{V})}}\\\\\nx\\arrow{u}{s_{\\textbf{F}(\\Sigma, \\mathcal{V})}}\n\\end{tikzcd}\n\\caption*{$\\textbf{F}(\\Sigma, \\mathcal{V})$}\n\\end{minipage}\n\\hspace{3cm}\n\\centering\n\\begin{minipage}[t]{4cm}\n\\centering\n\\begin{tikzcd}[column sep=1cm]\n0\\arrow[loop left, out=150, in=-150, distance=3em, swap]{}{s_{\\mathcal{A}}}\\arrow[shift right, swap]{r}{s_{\\mathcal{A}}} & 1\\arrow[loop right, out=30, in=-30, distance=3em]{}{s_{\\mathcal{A}}}\\arrow[shift right, swap]{l}{s_{\\mathcal{A}}}\n\\end{tikzcd}\n\\caption*{$\\mathcal{A}$}\n\\end{minipage}\n\\end{figure}\nWe state that $\\nu_{1}:F(\\Sigma, \\mathcal{V})\\rightarrow A$, given by $\\nu_{1}(\\alpha)=0$ for every $\\alpha\\in F(\\Sigma, \\mathcal{V})$, and $\\nu_{2}:F(\\Sigma, \\mathcal{V})\\rightarrow A$, given by $\\nu_{2}(\\alpha)=1$ for every $\\alpha\\in F(\\Sigma, \\mathcal{V})\\setminus\\{x\\}$ and $\\nu_{2}(x)=0$, are homomorphisms. In fact, for every $\\alpha\\in F(\\Sigma, \\mathcal{V})$, we have that\n\\[\\{\\nu_{i}(\\beta)\\ :\\ \\beta\\in s_{\\textbf{F}(\\Sigma, \\mathcal{V})}(\\alpha)\\}=\\{\\nu_{i}(s(\\alpha))\\}\\subseteq \\{0,1\\}=s_{\\mathcal{A}}(\\nu_{i}(\\alpha)),\\]\nwhere $i\\in \\{1,2\\}$. So $\\nu_{1}$ and $\\nu_{2}$ are different legal valuations, despite being the same over the variables.\n\\end{example}\n\nWe will say that a legal valuation $\\nu:\\textbf{F}(\\Sigma, \\mathcal{V})\\rightarrow\\mathcal{A}$ is associated\\index{Valuation, Associated legal} to the function $\\chi:\\mathcal{V}\\rightarrow A$ given by $\\chi=\\nu|_{\\mathcal{V}}$,\\footnote{Here, it is important to clarify the notation: for a function $f:X\\rightarrow Y$ and a set $Z\\subset X$, we will denote by $f|_{Z}$ the restriction of $f$ to $Z$.} which we shall call a valuation or interpretation of the variables; more generally, we say $\\nu$ is associated to the function $\\chi:\\mathcal{V}\\rightarrow\\mathcal{P}(A)\\setminus\\{\\emptyset\\}$ if $\\nu(x)\\in \\chi(x)$ for every $x\\in\\mathcal{V}$. What we have shown in the previous example is that two distinct legal valuations may be associated to the same interpretation of the variables.\n\nSo, if legal valuations are problematic, what other definition should we take? Many attempts to correctly interpret a formula in a multialgebra have been made, each with its own advantages and drawbacks. When studying choice-dependent freely generated multialgebras, we will show that maybe one needs to specify both an interpretation of the variables and what we shall call a collection of choices. For now, we shall offer a few other possibilities.\n\n\\begin{definition}\nGiven a $\\Sigma$-multialgebra $\\mathcal{A}$, a full multihomomorphism $\\nu:\\textbf{F}(\\Sigma, \\mathcal{V})\\rightarrow\\mathcal{A}$ is called a full valuation\\index{Valuation, Full}.\n\\end{definition}\n\n\\begin{proposition}\nTwo full valuations $\\nu_{1}, \\nu_{2}:\\textbf{F}(\\Sigma, \\mathcal{V})\\rightarrow\\mathcal{A}$ such that $\\nu_{1}|_{\\mathcal{V}}=\\nu_{2}|_{\\mathcal{V}}$ are the same.\n\\end{proposition}\n\n\\begin{proof}\nWe will prove that, for any formula $\\alpha$, $\\nu_{1}(\\alpha)=\\nu_{2}(\\alpha)$ by induction on the order of $\\alpha$. If $|\\alpha|=0$, either $\\alpha$ is an element $x\\in\\mathcal{V}$ and then $\\nu_{1}(x)=\\nu_{2}(x)$ since $\\nu_{1}|_{\\mathcal{V}}=\\nu_{2}|_{\\mathcal{V}}$, or $\\alpha$ is a $\\sigma\\in\\Sigma_{0}$, and then \n\\[\\nu_{1}(\\alpha)=\\sigma_{\\mathcal{A}}=\\nu_{2}(\\alpha).\\]\n\nAssuming the proposition is true for formulas of order at most $m$, if $\\alpha=\\sigma(\\alpha_{1}, \\dotsc , \\alpha_{n})$ is a formula of order $m+1$ then $\\alpha_{1}$ through $\\alpha_{n}$ are formulas of order at most $m$, and therefore $\\nu_{1}(\\alpha_{1})=\\nu_{2}(\\alpha_{2}), \\dotsc , \\nu_{1}(\\alpha_{n})=\\nu_{2}(\\alpha_{n})$. It follows that \n\\[\\nu_{1}(\\alpha)=\\bigcup_{(a_{1}, \\dotsc , a_{n})\\in \\nu_{1}(\\alpha_{1})\\times\\cdots\\nu_{1}(\\alpha_{n})}\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})=\\bigcup_{(a_{1}, \\dotsc , a_{n})\\in \\nu_{2}(\\alpha_{1})\\times\\cdots\\nu_{2}(\\alpha_{n})}\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})=\\nu_{2}(\\alpha).\\]\n\\end{proof}\n\nSo, to every function $\\chi:\\mathcal{V}\\rightarrow \\mathcal{P}(A)\\setminus\\{\\emptyset\\}$, there corresponds at most one full valuation: but we can also show that there is at least one full valuation associated to $\\chi$, which we denote by $\\overline{\\chi}$. We define it by induction on the order of a formula $\\alpha$:\n\\begin{enumerate}\n\\item if $|\\alpha|=0$, either $\\alpha$ is a $x\\in\\mathcal{V}$, when we define $\\overline{\\chi}(\\alpha)=\\chi(x)$, or $\\alpha$ is a $\\sigma\\in\\Sigma_{0}$, when $\\overline{\\chi}(\\alpha)=\\sigma_{\\mathcal{A}}$;\n\\item supposing $\\overline{\\chi}$ is defined for formulas of degree at most $m$, if $\\alpha=\\sigma(\\alpha_{1}, \\dotsc , \\alpha_{n})$ is of degree $m+1$, and so $\\alpha_{1}, \\dotsc , \\alpha_{n}$ are of degree at most $m$, we define\n\\[\\overline{\\chi}(\\alpha)=\\bigcup_{(a_{1}, \\dotsc , a_{n})\\in\\overline{\\chi}(\\alpha_{1})\\times\\cdots\\times\\overline{\\chi}(\\alpha_{n})}\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n}).\\]\n\\end{enumerate}\n\nWe can generalize both legal and full valuations through what we will call simply valuations: any multihomomorphism $\\nu:\\textbf{F}(\\Sigma, \\mathcal{V})\\rightarrow \\mathcal{A}$ will be said to be a valuation on $\\mathcal{A}$, associated to any of the functions $\\chi:\\mathcal{V}\\rightarrow \\mathcal{P}(A)\\setminus\\{\\emptyset\\}$ such that $\\nu(x)\\subseteq\\chi(x)$ for every $x\\in\\mathcal{V}$.\n\nIt is quite clear how this is a generalization of a full valuation, but one can see this is not a direct generalization of a legal valuation, since those are function from $F(\\Sigma, \\mathcal{V})$ to $A$, instead of $\\mathcal{P}(A)\\setminus\\{\\emptyset\\}$; however, when one considers for the legal valuation $\\nu$ the function $\\nu^{*}:F(\\Sigma, \\mathcal{V})\\rightarrow \\mathcal{P}(A)\\setminus\\{\\emptyset\\}$ such that $\\nu^{*}(\\alpha)=\\{\\nu(\\alpha)\\}$ for every $\\alpha\\in F(\\Sigma, \\mathcal{V})$, one sees that $\\nu^{*}$ is a valuation.\n\nWithout much further ado, we will identify $\\nu^{*}$ with $\\nu$, and characterize some valuations as legal ones.\n\n\\begin{proposition}\nFixed $\\textbf{F}(\\Sigma,\\mathcal{V})$, the set $V(\\mathcal{A})$ of all valuations from it to the $\\Sigma$-multialgebra $\\mathcal{A}$ is a partially ordered set, when for valuations $\\nu_{1}, \\nu_{2}\\in V(\\mathcal{A})$ one considers $\\nu_{1}\\leq \\nu_{2}$ if and only if \n\\[\\nu_{1}(\\alpha)\\subseteq \\nu_{2}(\\alpha),\\quad \\forall \\alpha\\in F(\\Sigma, \\mathcal{V}),\\]\nclosed under suprema of non-empty sets, with as minimal elements the set of all legal valuations $LV(\\mathcal{A})$ on $\\mathcal{A}$.\n\\end{proposition}\n\n\\begin{proof}\nFirst, we prove that the mentioned order is in fact an order.\n\\begin{enumerate}\n\\item For any $\\nu\\in V(\\mathcal{A})$ and all $\\alpha\\in F(\\Sigma, \\mathcal{V})$, it is true that $\\nu(\\alpha)\\subseteq\\nu(\\alpha)$, and therefore $\\nu\\leq \\nu$.\n\\item For any two $\\nu_{1}, \\nu_{2}\\in V(\\mathcal{A})$, if $\\nu_{1}\\leq\\nu_{2}$ and $\\nu_{2}\\leq \\nu_{1}$, then for every $\\alpha\\in F(\\Sigma, \\mathcal{V})$ we have that $\\nu_{1}(\\alpha)\\subseteq\\nu_{2}(\\alpha)$ and $\\nu_{2}(\\alpha)\\subseteq\\nu_{1}(\\alpha)$, and therefore $\\nu_{1}(\\alpha)=\\nu_{2}(\\alpha)$; it follows that $\\nu_{1}=\\nu_{2}$.\n\\item For any three $\\nu_{1}, \\nu_{2}, \\nu_{3}\\in V(\\mathcal{A})$, if $\\nu_{1}\\leq\\nu_{2}$ and $\\nu_{2}\\leq\\nu_{3}$, then for every $\\alpha\\in F(\\Sigma, \\mathcal{V})$ we see that $\\nu_{1}(\\alpha)\\subseteq\\nu_{2}(\\alpha)$ and $\\nu_{2}(\\alpha)\\subseteq\\nu_{3}(\\alpha)$, and therefore $\\nu_{1}(\\alpha)\\subseteq\\nu_{3}(\\alpha)$; it follows that $\\nu_{1}\\leq\\nu_{3}$.\n\\end{enumerate}\n\nNow, we state that the supremum of a non-empty set $\\Lambda\\subseteq V(\\mathcal{A})$ is the valuation such that, for every $\\alpha\\in F(\\Sigma, \\mathcal{V})$, \n\\[\\nu(\\alpha)=\\bigcup_{\\lambda\\in \\Lambda}\\lambda(\\alpha);\\]\nit is, in fact, a valuation since it is a multihomomorphism: if $\\sigma\\in\\Sigma_{n}$ and $\\alpha_{1}, \\dotsc , \\alpha_{n}$ are formulas in $F(\\Sigma, \\mathcal{V})$, we have that\n\\[\\nu(\\sigma(\\alpha_{1}, \\dotsc , \\alpha_{n}))=\\bigcup_{\\lambda\\in\\Lambda}\\lambda(\\sigma(\\alpha_{1}, \\dotsc , \\alpha_{n}))\\subseteq\\bigcup_{\\lambda\\in\\Lambda}\\bigcup_{(a_{1}, \\dotsc , a_{n})\\in \\lambda(\\alpha_{1})\\times\\cdots\\times\\lambda(\\alpha_{n})}\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})\\subseteq\\]\n\\[\\bigcup_{\\lambda\\in\\Lambda}\\bigcup_{(a_{1}, \\dotsc , a_{n})\\in \\nu(\\alpha_{1})\\times\\cdots\\times\\nu(\\alpha_{n})}\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})=\\bigcup_{(a_{1}, \\dotsc , a_{n})\\in \\nu(\\alpha_{1})\\times\\cdots\\times\\nu(\\alpha_{n})}\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n}).\\]\n\nClearly $\\nu$ is an upper bound of $\\Lambda$, so suppose $\\nu^{*}$ is another upper bound: for every formula $\\alpha$ and $\\lambda\\in \\Lambda$ we have that $\\lambda(\\alpha)\\subseteq \\nu^{*}(\\alpha)$, and from that $\\nu(\\alpha)=\\bigcup_{\\lambda\\in\\Lambda}\\lambda(\\alpha)\\subseteq \\nu^{*}(\\alpha)$, that is, $\\nu\\leq\\nu^{*}$ and so $\\nu=\\sup\\Lambda$.\n\nSince for any legal valuation $\\nu\\in LV(\\mathcal{A})$ and formula $\\alpha$ the set $\\nu(\\alpha)$ is always of cardinality $1$, legal valuations are clearly minimal elements.\n\\end{proof}\n\nFor a function $\\chi:\\mathcal{V}\\rightarrow\\mathcal{P}(A)\\setminus\\{\\emptyset\\}$, we would like to consider the set \n\\[V(\\chi)=\\{\\nu\\in V(\\mathcal{A})\\ :\\ \\nu\\leq\\overline{\\chi}\\}.\\]\nIt is clearly a partially ordered set, containing $\\overline{\\chi}$ and all valuations, legal or not, associated to $\\chi$; it has as minimal elements $LV(\\chi)=LV(\\mathcal{A})\\cap V(\\chi)$ and maximum $\\overline{\\chi}$; it is closed under suprema since for any $\\Lambda\\subseteq V(\\chi)$ we have that $\\overline{\\chi}$ is an upper bound for $\\Lambda$, and therefore $\\sup\\Lambda\\leq \\overline{\\chi}$, meaning that $\\sup\\Lambda\\in V(\\chi)$.\n\nNotice that, alternatively, $\\nu\\in V(\\chi)$ if, and only if, $\\nu$ is a valuation for which, for every $x\\in\\mathcal{V}$, $\\nu(x)\\subseteq \\chi(x)$.\n\nThe join-semilattice $V(\\chi)$ is then a very interesting object: if one is in doubt of how to interpret a formula given the interpretation of variables $\\chi$, using valuations, legal valuations or full valuations, $V(\\chi)$ captures the information offered by all of these attempts, despite being considerably more complex.\n\n \n\n\\subsection{Full valuations and adjointedness}\n\nTake the class of all sets and, as morphisms between any two sets $X$ and $Y$, take the functions $f:X\\rightarrow\\mathcal{P}(Y)\\setminus\\{\\emptyset\\}$, with the composition of two such functions $f:X\\rightarrow\\mathcal{P}(Y)\\setminus\\{\\emptyset\\}$ and $g:Y\\rightarrow\\mathcal{P}(Z)\\setminus\\{\\emptyset\\}$ being defined as the function $g\\circ f:X\\rightarrow\\mathcal{P}(Z)\\setminus\\{\\emptyset\\}$ such that, for any $x\\in X$, \n\\[g\\circ f(x)=\\bigcup_{y\\in f(x)}g(y).\\]\n\nAs we saw before, such a composition is associative, and has as identity, for a given set $X$, the function $Id_{X}:X\\rightarrow\\mathcal{P}(X)\\setminus\\{\\emptyset\\}$ such that, for every $x\\in X$, $Id_{X}(x)=\\{x\\}$. Therefore, this object is a category, which we will call the category of sets with multifunctions, and denote by $\\textbf{MSet}$\\label{MSet}; its morphisms will be called multifunctions.\n\nNow, fixed a signature $\\Sigma$, take for every set $X$ the $\\Sigma$-multialgebra $FX=\\textbf{F}(\\Sigma, X)$ and, for every morphism $f$ from $X$ to $Y$, that is, function $f:X\\rightarrow \\mathcal{P}(Y)\\setminus\\{\\emptyset\\}$, take the full multihomomorphism $Ff=\\overline{f}:\\textbf{F}(\\Sigma, X)\\rightarrow \\textbf{F}(\\Sigma, Y)$.\n\nGiven the identity $Id_{X}:X\\rightarrow X$ on $\\textbf{MSet}$, we state that $\\overline{Id_{X}}:\\textbf{F}(\\Sigma, X)\\rightarrow \\textbf{F}(\\Sigma, X)$ is exactly the identity $Id_{\\textbf{F}(\\Sigma, X)}$ of this multialgebra in $\\textbf{MMAlg}_{=}(\\Sigma)$: one can see this by induction over the order of a formula $\\alpha$.\n\\begin{enumerate}\n\\item If $\\alpha$ is of order $0$, either $\\alpha$ is an $x\\in X$ and then $\\overline{Id_{X}}(x)=Id_{X}(x)=\\{x\\}$, or $\\alpha$ is a $\\sigma\\in\\Sigma_{0}$, when $\\overline{Id_{X}}(\\sigma)=\\sigma_{\\textbf{F}(\\Sigma, X)}=\\{\\sigma\\}$.\n\\item If $\\alpha=\\sigma(\\alpha_{1}, \\dotsc , \\alpha_{n})$, for $\\sigma\\in\\Sigma_{n}$ and $\\alpha_{1}$ through $\\alpha_{n}$ of order less than that of $\\alpha$, we have that\n\\[\\overline{Id_{X}}(\\alpha)=\\bigcup_{(\\beta_{1}, \\dotsc , \\beta_{n})\\in \\overline{Id_{X}}(\\alpha_{1})\\times\\cdots\\times\\overline{Id_{X}}(\\alpha_{n})}\\sigma_{\\textbf{F}(\\Sigma, X)}(\\beta_{1}, \\dotsc , \\beta_{n})=\\bigcup_{(\\beta_{1}, \\dotsc , \\beta_{n})\\in \\{\\alpha_{1}\\}\\times\\cdots\\times\\{\\alpha_{n}\\}}\\{\\sigma(\\beta_{1}, \\dotsc , \\beta_{n})\\}=\\]\n\\[\\{\\sigma(\\alpha_{1}, \\dotsc , \\alpha_{n})\\}=\\{\\alpha\\}.\\]\n\\end{enumerate}\n\nFurthermore, given morphisms $f$ from $X$ to $Y$ and $g$ from $Y$ to $Z$ in $\\textbf{MSet}$, we see that for every $x\\in X$ \n\\[\\overline{g}\\circ\\overline{f}(x)=\\bigcup_{y\\in \\overline{f}(x)}\\overline{g}(y)=\\bigcup_{y\\in f(x)}\\overline{g}(y),\\]\nand since $f(x)\\subseteq Y$, for any $y\\in f(x)$ we have that $\\overline{g}(y)=g(y)$ and therefore\n\\[\\bigcup_{y\\in f(x)}\\overline{g}(y)=\\bigcup_{y\\in f(x)}g(y)=g\\circ f(x)=\\overline{g\\circ f}(x),\\]\nsince $g\\circ f$ goes from $X$ to $Z$; therefore both $\\overline{g}\\circ\\overline{f}$ and $\\overline{g\\circ f}$ are full multihomomorphisms equal over the variables, and are therefore equal. \n\nSince $FId_{X}=Id_{FX}$ and $Fg\\circ Ff=F(g\\circ f)$, we have proved\n\\[F:\\textbf{MSet}\\rightarrow\\textbf{MMAlg}_{=}(\\Sigma)\\]\nis a functor. Now, we can also consider the forgetful functor\n\\[\\mathcal{U}:\\textbf{MMAlg}_{=}(\\Sigma)\\rightarrow \\textbf{MSet},\\]\nsince full multihomomorphisms are multifunctions; what we shall prove now is that $F$ and $\\mathcal{U}$ are actually adjoint. So, we consider the functions, for a set $X$ and a $\\Sigma$-multialgebra $\\mathcal{A}$,\n\\[\\Phi_{\\mathcal{A}, X}:Hom_{\\textbf{MSet}}(X, \\mathcal{U}\\mathcal{A})\\rightarrow Hom_{\\textbf{MMAlg}_{=}(\\Sigma)}(FX, \\mathcal{A})\\]\nassociating a map $f:X\\rightarrow\\mathcal{P}(A)\\setminus\\{\\emptyset\\}$ to the full multihomomorphism $\\overline{f}:\\textbf{F}(\\Sigma, X)\\rightarrow\\mathcal{A}$ extending $f$. Since, for every $f$, there corresponds one and only one $\\overline{f}$, the $\\Phi_{\\mathcal{A}, X}$ are bijections.\n\nNow, given sets $X$ and $Y$, $\\Sigma$-multialgebras $\\mathcal{A}$ and $\\mathcal{B}$, a morphism $f$ from $Y$ to $X$ in $\\textbf{MSet}$ and a full multihomomorphism $\\varphi:\\mathcal{A}\\rightarrow\\mathcal{B}$, we must only prove that the following diagram commutes.\n\n\\[ \\begin{tikzcd}[row sep=5em, column sep=3em]\nHom_{\\textbf{MSet}}(X, \\mathcal{U}\\mathcal{A}) \\arrow{r}{\\Phi_{\\mathcal{A}, X}} \\arrow{d}{Hom(f, \\mathcal{U}\\varphi)}& Hom_{\\textbf{MMAlg}_{=}(\\Sigma)}(FX, \\mathcal{A}) \\arrow{d}{Hom(Ff, \\varphi)} \\\\%\nHom_{\\textbf{MSet}}(Y, \\mathcal{U}\\mathcal{B}) \\arrow{r}{\\Phi_{\\mathcal{B}, Y}}& Hom_{\\textbf{MMAlg}_{=}(\\Sigma)}(FY, \\mathcal{B})\n\\end{tikzcd}\n\\]\n\nSo, denoting the universe of $\\mathcal{A}$ by $A$ as usual, we take a function $g:X\\rightarrow\\mathcal{P}(A)\\setminus\\{\\emptyset\\}$ in $Hom_{\\textbf{MSet}}(X, \\mathcal{U}\\mathcal{A})$: on the top edge of the diagram we obtain the full multihomomorphism $\\Phi_{\\mathcal{A}, X}(g)=\\overline{g}$ from $\\textbf{F}(\\Sigma, X)$ to $\\mathcal{A}$; and on the right edge we obtain the once again full multihomomorphism $\\varphi\\circ \\overline{g}\\circ\\overline{f}$, since $Ff=\\overline{f}$, from $\\textbf{F}(\\Sigma, Y)$ to $\\mathcal{B}$.\n\nOn the left edge, we obtain the multifunction $\\varphi\\circ g\\circ f$ from $Y$ to the universe $B$ of $\\mathcal{B}$, since $\\mathcal{U}\\varphi=\\varphi$; applying $\\Phi_{\\mathcal{B}, Y}$ on the bottom edge of the diagram we obtain $\\overline{\\varphi\\circ g\\circ f}$, also a full multihomomorphism from $\\textbf{F}(\\Sigma, Y)$ to $\\mathcal{B}$.\n\nNow, to prove that the diagram commutes or, what is equivalent, that $\\varphi\\circ\\overline{g}\\circ\\overline{f}=\\overline{\\varphi\\circ g\\circ f}$, since both are full valuations on $\\textbf{F}(\\Sigma, Y)$ is enough to prove that, when restricted to $Y$, both are the same. So, for $y\\in Y$,\n\\[\\varphi\\circ\\overline{g}\\circ\\overline{f}(y)=\\bigcup_{x\\in \\overline{f}(x)}\\varphi\\circ\\overline{g}(x)=\\bigcup_{x\\in\\overline{f}(x)}\\bigcup_{a\\in \\overline{g}(x)}\\varphi(a)=\\bigcup_{x\\in f(x)}\\bigcup_{a\\in\\overline{g}(x)}\\varphi(a)=\\bigcup_{x\\in f(y)}\\bigcup_{a\\in g(x)}\\varphi(a)=\\]\n\\[\\bigcup_{x\\in f(y)}\\varphi\\circ g(x)=\\varphi\\circ g\\circ f(y)=\\overline{\\varphi\\circ g\\circ f}(y),\\]\nwhat ends the proof that $F$ and $\\mathcal{U}$ are adjoint.\n\n\n\\begin{proposition}\nLet $\\chi:\\mathcal{V}\\rightarrow\\mathcal{P}(A)\\setminus\\{\\emptyset\\}$ be a function and $\\varphi:\\textbf{F}(\\Sigma, \\mathcal{V})\\rightarrow\\mathcal{A}$ a multihomomorphism such that, for every $x\\in\\mathcal{V}$, $\\varphi(x)\\subseteq\\chi(x)$: then $\\varphi\\leq \\overline{\\chi}$.\n\\end{proposition}\n\n\\begin{proof}\nIt is enough to proceed by induction over the order of a formula $\\alpha$.\n\nIf it is $0$, either $\\alpha$ is an element $x\\in\\mathcal{V}$, when the result is true by hypothesis; or $\\alpha$ is a $\\sigma\\in\\Sigma_{0}$, and since $\\varphi$ is a multihomomorphism we have that $\\varphi(\\sigma)\\subseteq \\sigma_{\\mathcal{A}}=\\overline{\\chi}(\\sigma)$.\n\nNow, suppose the result is true for formulas of order smaller than that of $\\alpha$, and suppose $\\alpha$ is of the form $\\sigma(\\alpha_{1}, \\dotsc , \\alpha_{n})$: then \n\\[\\varphi(\\alpha)=\\bigcup_{\\beta\\in\\sigma_{\\textbf{F}(\\Sigma, \\mathcal{V})}(\\alpha_{1}, \\dotsc , \\alpha_{n})}\\varphi(\\beta)\\subseteq\\bigcup_{(\\beta_{1}, \\dotsc , \\beta_{n})\\in \\varphi(\\alpha_{1})\\times\\cdots\\times\\varphi(\\alpha_{n})}\\sigma_{\\mathcal{A}}(\\beta_{1}, \\dotsc , \\beta_{n})\\subseteq\\]\n\\[\\bigcup_{(\\beta_{1}, \\dotsc , \\beta_{n})\\in\\overline{\\chi}(\\alpha_{1})\\times\\cdots\\times\\overline{\\chi}(\\alpha_{n})}\\sigma_{\\mathcal{A}}(\\beta_{1}, \\dotsc , \\beta_{n})=\\bigcup_{\\beta\\in\\sigma_{\\textbf{F}(\\Sigma, \\mathcal{V})}(\\alpha_{1}, \\dotsc , \\alpha_{n})}\\overline{\\chi}(\\beta)=\\overline{\\chi}(\\alpha).\\]\n\\end{proof}\n\nThis way, we see that full valuations have many important properties: they allow, together with the algebras of formulas, to build a functor adjoint to a forgetful functor on a rather natural category ($\\textbf{MSet}$); furthermore, they are maximal among valuations associated to a given evaluation of the variables.\n\nAnd, at the same time, they are not very useful to most applications in logic, one reason being that, from the point of view of full valuations, the free objects of $\\textbf{MMAlg}_{=}(\\Sigma)$ are the algebras of formulas $\\textbf{F}(\\Sigma, \\mathcal{V})$, meaning we gain no new structures to analyze. A second reason for or disinterest on full valuations is that they are not at all very precise: legal valuations, that are only functions instead of multifunctions, are much more useful and \"precise\". Unfortunately, to an evaluation of the variables there are many legal valuations associated: one approach in the direction of fixing this problem is to consider collections of choices (Definition \\ref{Collection of Choices}), which will allow us at the same time to consider generalizations of the algebras of formulas (Definition \\ref{Multialgebra of formulas}).\n\nWe will work those objects in Chapter \\ref{Chapter2}. Their algebraic structure is rather rich, regardless of how difficult their treatment is trough category theory. And, speaking of category theory, the categories shown up until now can be arranged as in the diagram below. The functor $J_{\\textbf{Set}}$ is the identity on sets and takes functions to multifunctions in the obvious way, like $J$ and $J_{=}$; the question of whether the forgetful functor from $\\textbf{MAlg}_{=}(\\Sigma)$ to $\\textbf{Set}$, here represented by a dashed arrow, has a left adjoint is answered negatively at the end of Section \\ref{Freely generated multialgebra}.\n\n\\[ \\begin{tikzcd}\n\\textbf{MAlg}(\\Sigma) \\arrow{dd}{J} && \\textbf{MAlg}_{=}(\\Sigma)\\arrow[hook]{ll}\\arrow[dashed]{rr} \\arrow{dd}{J_{=}} &&\\textbf{Set}\\arrow{dd}{J_{\\textbf{Set}}} \\\\\n&&&&\\\\\n \\textbf{MMAlg}(\\Sigma) && \\textbf{MMAlg}_{=}(\\Sigma)\\arrow[hook]{ll}\\arrow[transform canvas={yshift=.7ex}]{rr}{\\mathcal{U}} && \\textbf{MSet}\\arrow[transform canvas={yshift=-.7ex}]{ll}{F}\n\\end{tikzcd}\n\\]\n\n\n\n\n\\section{Lattices, and Boolean and Heyting algebras}\\label{Lattices, and Boolean... }\n\nConsider the signature $\\Sigma_{\\textbf{Lat}}$\\label{SigmaLat} with two binary connectives, $\\vee$ and $\\wedge$: that is $(\\Sigma_{\\textbf{Lat}})_{2}=\\{\\vee, \\wedge\\}$ and $(\\Sigma_{\\textbf{Lat}})_{n}=\\emptyset$, for $n\\neq 2$. \n\n\\begin{definition}\nA lattice\\index{Lattice} is a $\\Sigma_{\\textbf{Lat}}$-algebra $\\mathcal{L}=(L,\\{\\sigma_{\\mathcal{L}}\\}_{\\sigma\\in\\Sigma})$ satisfying, for every $x,y,z\\in L$,\n\\begin{enumerate}\n\\item $x\\vee y=y\\vee x$ and $x\\wedge y=y\\wedge x$ (commutative laws);\n\\item $x\\vee(y\\vee z)=(x\\vee y)\\vee z$ and $x\\wedge(y\\wedge z)=(x\\wedge y)\\wedge z$ (associative laws);\n\\item $x\\vee x=x$ and $x\\wedge x=x$ (idempotent laws);\n\\item $x\\vee(x\\wedge y)=x$ and $x\\wedge(x\\vee y)=x$ (absorption laws).\n\\end{enumerate}\n\\end{definition}\nIn this definition, for simplicity, we have denoted $\\vee_{\\mathcal{L}}(x,y)$ and $\\wedge_{\\mathcal{L}}(x,y)$ by, respectively, $x\\vee y$ and $x\\wedge y$.\n\nThis is often regarded as the algebraic definition of a lattice, an order-theoretic one existing as well. A partial order in a set $X$ is a binary relation $\\leq$ on $X$, meaning a subset of $X\\times X$, such that, for all $x, y, z\\in X$,\n\\begin{enumerate}\n\\item $x\\leq x$ (reflexivity);\n\\item $x\\leq y$ and $y\\leq x$ imply $x=y$ (antisymmetry);\n\\item $x\\leq y$ and $y\\leq z$ imply $x\\leq z$ (transitivity),\n\\end{enumerate}\nwhere we will denote the fact that $(x,y)\\in\\leq$ by $x\\leq y$. A set $X$, together with a partial order $\\leq$ in it, will be named a partially ordered set, or poset. \n\nIn a poset $(X, \\leq)$, we say $x$ is an upper bound of $S\\subseteq X$ if $y\\leq x$ for every $y\\in S$; $x$ is a lower bound for $S$ if $x\\leq y$ for every $y\\in S$. Then, the supremum of $S\\subseteq X$, if it exists, is its least upper bound, meaning an upper bound $\\sup S\\in X$ for $S$ such that, for any other upper bound $x$ for $S$, $\\sup S\\leq x$; the infimum of $S$, if it exists, is its greatest lower bound, meaning a lower bound $\\inf S$ for $S$ such that, for any other lower bound $x$ for $S$, $x\\leq \\inf S$.\n\n\\begin{definition}\nA lattice is a poset $\\mathcal{L}=(L,\\leq)$ such that, for any elements $x, y\\in L$, $\\sup\\{x,y\\}$ and $\\inf\\{x,y\\}$ both exist.\n\\end{definition}\n\nOne can translate between the two definitions: given a lattice $\\mathcal{L}$ presented as a $\\Sigma_{\\textbf{Lat}}$-algebra $(L,\\{\\sigma_{\\mathcal{L}}\\}_{\\sigma\\in\\Sigma_{\\textbf{Lat}}})$, we may define an order $\\leq$ in $L$ by \n\\[x\\leq y\\quad\\text{if and only if}\\quad x=x\\wedge y,\\]\nand it is easy to prove that in this case $(L,\\leq)$ is a lattice, where $x\\wedge y$ and $x\\vee y$ are, respectively, the infimum and supremum of $x$ and $y$. Reciprocally, given a lattice $\\mathcal{L}$ presented as a poset $(L,\\leq)$, we may define operations $\\vee$ and $\\wedge$ on $L$ by\n\\[x\\wedge y=\\inf\\{x,y\\}\\quad\\text{and}\\quad x\\vee y=\\sup\\{x,y\\},\\]\nfrom what we obtain a $\\Sigma_{\\textbf{Lat}}$-algebra on which $x\\leq y$ if, and only if, $x=x\\wedge y$.\n\nAn element $0$ of a lattice $\\mathcal{L}$ is a minimum, or bottom element, when either: $x\\vee 0=x$, for every $x\\in L$; $x\\wedge 0=0$, for every $x\\in L$; or $0\\leq x$, for every $x\\in X$. Notice that all three conditions are equivalent. An element $1$ of $\\mathcal{L}$ is a maximum, or top element, when either: $x\\vee 1=1$, for every $x\\in L$; $x\\wedge 1=x$, for every $x\\in L$; or $x\\leq 1$, for every $x\\in L$. Again, all three conditions are equivalent. \n\nTo accommodate bottom and top elements, we extend the signature $\\Sigma_{\\textbf{Lat}}$ to $\\Sigma_{\\textbf{Lat}}^{0}$\\label{SigmaLat0}, $\\Sigma_{\\textbf{Lat}}^{1}$\\label{SigmaLat1} and $\\Sigma_{\\textbf{Lat}}^{0,1}$\\label{SigmaLat01} by adding, respectively, a $0$-ary symbol $0$, a $0$-ary symbol $1$ and $0$-ary symbols $0$ and $1$. For simplicity, we will drop the indexes $\\mathcal{L}$ from $0_{\\mathcal{L}}$ and $1_{\\mathcal{L}}$.\n\n\\begin{definition}\n\\begin{enumerate}\n\\item A lattice with bottom (respectively top) element is a $\\Sigma_{\\textbf{Lat}}^{0}$-algebra ($\\Sigma_{\\textbf{Lat}}^{1}$-algebra) $\\mathcal{L}=(L,\\{\\sigma_{\\mathcal{L}}\\}_{\\sigma\\in\\Sigma_{\\textbf{Lat}}^{0}})$ such that $(L,\\{\\sigma_{\\mathcal{L}}\\}_{\\sigma\\in\\Sigma_{\\textbf{Lat}}})$ is a lattice and $0$ is a bottom element ($1$ is a top element).\n\\item A bounded lattice\\index{Lattice, Bounded} is a $\\Sigma_{\\textbf{Lat}}^{0,1}$-algebra $\\mathcal{L}=(L,\\{\\sigma_{\\mathcal{L}}\\}_{\\sigma\\in\\Sigma_{\\textbf{Lat}}^{0,1}})$ such that $(L,\\{\\sigma_{\\mathcal{L}}\\}_{\\sigma\\in\\Sigma_{\\textbf{Lat}}})$ is a lattice, $0$ is a bottom element and $1$ is a top element.\n\\end{enumerate}\n\\end{definition}\n\nIn a lattice, \"$x$ implies $y$\" is defined as the element\n\\[\\sup\\{z\\in L: x\\wedge z\\leq y\\},\\]\nif it exists, and denoted by $x\\rightarrow y$. We add to the signatures $\\Sigma_{\\textbf{Lat}}^{1}$ and $\\Sigma_{\\textbf{Lat}}^{0,1}$ the binary symbol $\\rightarrow$, obtaining respectively the signatures $\\Sigma_{\\textbf{Imp}}$\\label{SigmaImp} and $\\Sigma_{\\textbf{Hey}}$\\label{SigmaHey}.\n\n\\begin{definition}\n\\begin{enumerate}\n\\item An implicative lattice\\index{Lattice, Implicative} is a $\\Sigma_{\\textbf{Imp}}$-algebra $\\mathcal{L}=(L,\\{\\sigma_{\\mathcal{L}}\\}_{\\sigma\\in\\Sigma_{\\textbf{Imp}}})$ such that $(L, \\{\\sigma_{\\mathcal{L}}\\}_{\\sigma\\in\\Sigma_{\\textbf{Lat}}^{1}})$ is a lattice with top element, $\\sup\\{z\\in L: x\\wedge z\\leq y\\}$ exists for any two $x,y\\in L$ and \n\\[x\\rightarrow y=\\sup\\{z\\in L: x\\wedge z\\leq y\\}.\\]\n\n\\item A Heyting algebra\\index{Heyting algebra} is a $\\Sigma_{\\textbf{Hey}}$-algebra such that $(L,\\{\\sigma_{\\mathcal{L}}\\}_{\\sigma\\in\\Sigma_{\\textbf{Imp}}})$ is an implicative lattice and $0$ is a bottom element.\n\\end{enumerate}\n\\end{definition}\n\nIt is important here to point out that various different definitions of implicative lattices and Heyting algebras exist in the literature, most, if not all, equivalent to each other; we chose definitions most befitting of our purposes, taken from \\cite{ParLog}.\n\nAnother relevant point to make is that Heyting algebras are no strange beasts: classical propositional logic is to Boolean algebras, which we shall soon formally define, as intuitionistic logic is to Heyting algebras, meaning that not only are Heyting algebras models of intuitionistic logic, but their class characterizes intuitionistic logic (unfortunately, the parallel ends there: although the two-valued Boolean algebra is, by itself, capable of characterizing $\\textbf{CPL}$, no finite algebra, Heyting or not, can do the same to intuitionistic logic, see \\cite{Godel}). One could argue that while intuitionistic logic possesses a negation, Heyting algebras do not, but this can be easily dealt with: it is enough to define\n\\[\\neg x=x\\rightarrow 0.\\]\n\nWith this, we can define Heyting algebras over the signature $\\Sigma_{\\textbf{Boo}}$\\label{SigmaBoo}, obtained from $\\Sigma_{\\textbf{Hey}}$ by addition of an unary symbol \"$\\neg$\", by merely requesting that $\\neg x=x\\rightarrow 0$ for all elements $x$; we say $\\neg x$ is the partial complement, or (intuitionistic) negation of $x$.\n\n\\begin{definition}\nA Boolean algebra\\index{Boolean algebra} is a Heyting algebra over $\\Sigma_{\\textbf{Boo}}$ such that, for every $x$ in its universe, we have the law of excluded middle, or \\textit{tertium non datur}:\n\\[x\\vee\\neg x=1.\\]\n\\end{definition}\n\nThe following lemma will need an observation about the definition of implication on a Heyting algebra: notice that, if $x\\wedge z\\leq y$, then $z\\leq x\\rightarrow y$, by the very definition of $x\\rightarrow y$ as the supremum of $\\{z\\in L: x\\wedge z\\leq y\\}$. Reciprocally, if $z\\leq x\\rightarrow y$, $x\\wedge z\\leq x\\wedge(x\\rightarrow y)\\leq y$,\\footnote{Here we are using that $x\\leq y$ implies $x\\wedge z\\leq y\\wedge z$, but this is trivial to prove: if $x\\leq y$, $x=x\\wedge y$, and so $(x\\wedge z)\\wedge(y\\wedge z)=(x\\wedge y)\\wedge z=x\\wedge z$.} meaning that $z\\leq x\\rightarrow y$ if, and only if, $x\\wedge z\\leq y$. The following proof, although quite standard, follows closely the one found in \\cite{HeytingAlgebras}.\n\n\\begin{lemma}\nIn a Heyting algebra, $\\vee$ is distributive over $\\wedge$ and vice-versa, meaning that\n\\[x\\vee(y\\wedge z)=(x\\vee y)\\wedge(x\\vee z)\\quad\\text{and}\\quad x\\wedge(y\\vee z)=(x\\wedge y)\\vee(x\\wedge z),\\]\nfor all $x$, $y$ and $z$ in the universe.\n\\end{lemma}\n\n\\begin{proof}\nWe will prove that $x\\wedge(y\\vee z)=(x\\wedge y)\\vee(x\\wedge z)$, being the other equality proved analogously. Even more, we need only to prove that $x\\wedge(y\\vee z)\\leq (x\\wedge y)\\vee(x\\wedge z)$, since in any lattice one finds that $y\\leq y\\vee z$ and $z\\leq y\\vee z$, meaning that $x\\wedge y\\leq x\\wedge (y\\vee z)$ and $x\\wedge z\\leq x\\wedge(y\\vee z)$, and therefore $(x\\wedge y)\\vee(x\\wedge z)\\leq x\\wedge (y\\vee z)$.\n\nLet $w$ denote $(x\\wedge y)\\vee(x\\wedge z)$, and since $x\\wedge y\\leq w$ and $x\\wedge z\\leq w$, we obtain $y\\leq x\\rightarrow w$ and $z\\leq x\\rightarrow w$. So $y\\vee z\\leq x\\rightarrow w$, and we get, as desired,\n\\[x\\wedge(y\\vee z)\\leq x\\wedge(x\\rightarrow w)\\leq w=(x\\wedge y)\\vee(x\\wedge z).\\]\n\\end{proof}\n\nIn the following proposition, we omit several more trivial steps, such as proving that $\\neg x\\wedge x=0$. The reader is invited to fill them in as they appear.\n\n\\begin{proposition}\nA Heyting algebra $\\mathcal{H}$, with universe $H$, is a Boolean algebra if, and only if, $x\\vee(x\\rightarrow y)=1$ for all $x, y\\in H$.\n\\end{proposition}\n\n\\begin{proof}\nStart by assuming that we have $x\\vee(x\\rightarrow y)=1$ for all $x, y\\in H$: then, since in particular $0\\in H$, $x\\vee\\neg x=x\\vee(x\\rightarrow 0)=1$ for all $x\\in H$, meaning $\\mathcal{H}$ is a Boolean algebra.\n\nReciprocally, suppose $\\mathcal{H}$ is a Boolean algebra. We state that $x\\rightarrow y=\\neg x\\vee y$, being then necessary to prove that $\\neg x\\vee y$ is the supremum of $\\{z\\in H: x\\wedge z\\leq y\\}$: since \n\\[x\\wedge(\\neg x\\vee y)=(x\\wedge\\neg x)\\vee(x\\wedge y)=0\\vee (x\\wedge y)=x\\wedge y,\\]\nwhich is smaller or equal to $y$, $\\neg x\\vee y$ is in fact an element of this set; furthermore, if $x\\wedge z\\leq y$, then \n\\[z\\leq \\neg x\\vee z=1\\wedge (\\neg x\\vee z)=(\\neg x\\vee x)\\wedge (\\neg x\\vee z)=\\neg x\\vee(x\\wedge z)\\leq \\neg x\\vee y.\\]\n\nThen, since $x\\rightarrow y=\\neg x\\vee y$,\n\\[x\\vee (x\\rightarrow y)=x\\vee(\\neg x\\vee y)=(x\\vee \\neg x)\\vee y=1\\vee y=1,\\]\nas we wished to prove.\n\\end{proof}\n\nOf course, this means we could avoid adding a negation to the signature of Heyting algebras to be able to express Boolean algebras: a Boolean algebra is a $\\Sigma_{\\textbf{Hey}}$-algebra which is a Heyting algebra and satisfies, for all $x$ and $y$ in its universe, $x\\vee(x\\rightarrow y)=1$. But we prefer to have a negation at hand when dealing with Boolean algebras, it is just more convenient that way.\n\nNow, it is important to point out that some symbols on the signature $\\Sigma_{\\textbf{Boo}}$ can be changed, usually because we are inserting Boolean algebras in a context where they are already in use. So $0$ may be replaced with \"$\\bot$\", $1$ with \"$\\top$\" and $\\neg$ with \"$\\sim$\". We will make such changes more or less freely, without much ado.\n\nAs a final remark, we would like to make clear that lattices, implicative or not, and Heyting algebras will play a very minor role in what is to come: they are, more importantly, milestones in defining Boolean algebras that also appear when dealing with some swap structures, specifically for paraconsistent logics (\\cite{ParLog}). Boolean algebras, however, will be used time and time again in our study: in Section \\ref{bottomless Boolean algebras} we will use them to search for a category equivalent to that of multialgebras, when we will offer a more order-theoretic approach to the subject; Section \\ref{B-valuations} shows a new semantics of valuations for da Costa's logics $C_{n}$ based on Boolean algebras, which is used in Section \\ref{RNmatrices RMBCn} to offer yet another semantics for those logics, proven in Section \\ref{BA and RSwap} to generate a category of models for $C_{n}$ isomorphic to the category of Boolean algebras itself.\n\n\n\n\n\n\n\n\n\\newpage\n\n\\printbibliography[segment=\\therefsegment,heading=subbibliography]\n\\end{refsegment}\n\n\n\\begin{refsegment}\n\\defbibfilter{notother}{not segment=\\therefsegment}\n\\setcounter{chapter}{1}\n\\chapter{Weakly free multialgebras}\\label{Chapter2}\\label{Chapter 2}\n\n\nIn the study of universal algebra, it is well known (\\cite{Burris}) that there exist $\\Sigma$-algebras $\\mathcal{A}$ freely generated\\index{Freely generated} (or absolutely freely generated) by subsets $X$ of their universes $A$, meaning that for any $\\Sigma$-algebra $\\mathcal{B}$, with universe $B$, and function $f:X\\rightarrow B$, there exists precisely one homomorphism $\\overline{f}: \\mathcal{A}\\rightarrow\\mathcal{B}$ extending $f$, and therefore commuting the following diagram in $\\textbf{Set}$, for $j:X\\rightarrow A$ the inclusion.\n\\[\n\\begin{tikzcd}\n A \\arrow{ddrr}{\\overline{f}} & & \\\\\n & & \\\\\nX \\arrow{uu}{j} \\arrow{rr}{f} & & B\n\\end{tikzcd}\n\\]\nMoreover, the $\\Sigma$-algebra freely generated by $X$ is isomorphic to the $\\Sigma$-algebra of formulas over $X$, that is $\\textbf{F}(\\Sigma, X)$, and therefore unique up to isomorphisms; even more, given $\\textbf{F}(\\Sigma, X)$ and $\\textbf{F}(\\Sigma, Y)$ are isomorphic whenever $X$ and $Y$ are of the same cardinality, we discover there is precisely one freely generated $\\Sigma$-algebra, up to isomorphisms, for each cardinality. Equivalently, in the language of categories, the forgetful functor \n\\[U:\\textbf{Alg}(\\Sigma)\\rightarrow\\textbf{Set},\\]\nfrom the category of $\\Sigma$-algebras (with homomorphisms between them) to the category of sets, which takes a $\\Sigma$-algebra and returns its underlying universe, has a left adjoint $F$: it associates to a set $X$ any $\\Sigma$-algebra freely generated by $X$, and to a function the only homomorphism extending it.\n\nHowever, an algebraic structure being absolutely free is a concept that does not extend well to the context of multialgebras (\\cite{Marty}), specially when one restricts oneself to the non-partial multialgebras (whose multioperations do not return the empty-set), as we often do here given our interest on non-deterministic semantics, specifically those designed for paraconsistency. It is easy to prove that, first of all, freely generated multialgebras, which generalize freely generated algebras in the most obvious way, do not exist, and second, that the forgetful functor \n\\[\\mathcal{U}:\\textbf{MAlg}(\\Sigma)\\rightarrow\\textbf{Set},\\]\nfrom the category of multialgebras over the signature $\\Sigma$ to the category of sets, does not have a left adjoint. These results are deeply folkloric, being difficult to pinpoint a proof of them in standard literature, although it seems no one is not aware of their validity.\n\nWe here expose our reasons to believe that understanding as formulas, in $\\textbf{MAlg}(\\Sigma)$, only those elements found in the universe of $\\textbf{F}(\\Sigma, \\mathcal{V})$ disregards other multialgebras with an astoundingly similar behavior, so that we generalize the algebras of formulas to multialgebras of formulas. These structures indeed share many of the properties one expects of the algebras of formulas (or, equivalently, freely generated algebras):\n\n\\begin{enumerate}\n\\item they possess the unique extension property not for functions, but rather pairs of functions and what we will call collections of choices, that ``select'' how a homomorphism will approach indeterminacies; \n\nin fact, given multialgebras $\\mathcal{A}$ and $\\mathcal{B}$ and an $n$-ary $\\sigma$, for all $n$-tuples $(a_{1},\\dotsc , a_{n})\\in A^{n}$ and $(b_{1}, \\dotsc , b_{n})\\in B^{n}$ a collection of choices determines how to map those elements of $\\sigma_{\\mathcal{A}}(a_{1},\\dotsc , a_{n})$ into those of $\\sigma_{\\mathcal{B}}(b_{1}, \\dotsc , b_{n})$;\n\n\\item they are somewhat ``free'' of identities, an intuition we formalize through disconnected multialgebras, and they are generated by a set of ``indecomposable'' elements we shall call the ground of the multialgebra, much like variables (which are formulas without proper subformulas and therefore indecomposable in some sense); \n\nmore formally, a multialgebra is disconnected whenever different operations, or the same operation performed on different elements, always return disjoint results, while the ground $G(\\mathcal{A})$ of a multialgebra $\\mathcal{A}$ is the subset of its universe of all elements $a$ for which there do not exist $\\sigma$ (of arity $n$) and elements $a_{1}, \\dotsc , a_{n}$ of $\\mathcal{A}$ such that $a\\in\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})$;\n\n\\item strengthening the previous point, they are disconnected and have a minimum generating set that behaves quite similarly to a basis, of e.g. a vector space;\n\nnotice, however, that while a basis of a vector space is a minimal generating set, we are looking here at minimum generating sets, so we use the terminology of ``strong basis'', which end up being precisely the grounds we have mentioned earlier;\n\n\\item a final pair of properties we present is that they are simultaneously disconnected and satisfy that every sequence of further simpler and simpler elements eventually ends on an indecomposable element, condition we call being ``chainless'' and that implies having a strong basis;\n\nessentially, by $a$ being simpler than $b$ we mean that there exist $\\sigma$ (of arity $n$) and elements $a_{1}, \\dotsc , a_{n}$ such that $a_{i}=a$, for some $i\\in\\{1, \\dotsc , n\\}$, and $b\\in\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})$.\n\\end{enumerate}\n\nFurthermore, we prove all these four listed items are equivalent, characterizing exactly the same multialgebras; and with these weakly free multialgebras we here define at hand, we can offer simple proofs that freely generated multialgebras (now in the naive generalization) do not exist, and that $\\mathcal{U}$ does not have a left adjoint.\n\nMost of the research presented in this chapter was submitted as a preprint in \\cite{AbsFreeHyp} and finally published in \\cite{WeaklyFreeMultialgebras}.\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\\section{Formulas}\n\nHere we briefly recall the basic notions involving formulas found in Section \\ref{Formulas and how to interpret them}. Given a set $\\mathcal{V}$ of propositional variables and a signature $\\Sigma=\\{\\Sigma_{n}\\}_{n\\in\\mathbb{N}}$, the algebra of formulas freely generated by $\\mathcal{V}$ over $\\Sigma$ will be denoted $\\textbf{F}(\\Sigma, \\mathcal{V})$, and its universe will be denoted $F(\\Sigma, \\mathcal{V})$. Intuitively, the set of formulas $F(\\Sigma, \\mathcal{V})$ is the smallest set containing:\n\\begin{enumerate}\n\\item the variables $\\mathcal{V}$;\n\\item the formula $\\sigma(\\alpha_{1}, \\dotsc , \\alpha_{n})$, given a $\\sigma\\in\\Sigma_{n}$ and already defined formulas $\\alpha_{1}, \\dotsc , \\alpha_{n}$ in $F(\\Sigma, \\mathcal{V})$.\n\\end{enumerate}\n\nMore formally, $F(\\Sigma, \\mathcal{V})$ should be smallest set containing:\n\\begin{enumerate}\n\\item for every $x\\in\\mathcal{V}$, the function $f_{x}:\\{0\\}\\rightarrow \\mathcal{V}\\cup \\Sigma$ defined by $f_{x}(0)=x$;\n\\item the function \n\\[f_{\\sigma(\\alpha_{1}, \\dotsc , \\alpha_{n})}:\\{0, \\dotsc , 1+\\sum_{i=1}^{n}m_{n}\\}\\rightarrow \\mathcal{V}\\cup\\Sigma,\\]\ngiven a $\\sigma\\in\\Sigma_{n}$ and already defined formulas $f_{\\alpha_{1}}:\\{0, \\dotsc , m_{1}\\}\\rightarrow\\mathcal{V}\\cup\\Sigma, \\dotsc , f_{\\alpha_{n}}:\\{0, \\dotsc , m_{n}\\}\\rightarrow\\mathcal{V}\\cup\\Sigma$, such that $f_{\\sigma(\\alpha_{1}, \\dotsc , \\alpha_{n})}(0)=\\sigma$ and, for every $j\\in\\{1, \\dotsc , n\\}$ and $k\\in\\{0, \\dotsc , m_{j}\\}$,\n\\[f_{\\sigma(\\alpha_{1}, \\dotsc , \\alpha_{n})}(k+\\sum_{i=0}^{j-1}m_{i})=f_{\\alpha_{j}}(k),\\]\nwhere for simplicity we define $m_{0}=1$.\n\\end{enumerate}\nOne should notice that $f_{\\sigma(\\alpha_{1}, \\dotsc , \\alpha_{n})}$ is simply the formalization of the polish notation of $\\sigma(\\alpha_{1}, \\dotsc , \\alpha_{n})$.\n\nThe set $F(\\Sigma, \\mathcal{V})$ becomes the $\\Sigma-$algebra $\\textbf{F}(\\Sigma, \\mathcal{V})$ when we define, for a $\\sigma\\in\\Sigma_{n}$ and formulas $\\alpha_{1}, \\dotsc , \\alpha_{n}$ in $F(\\Sigma, \\mathcal{V})$, \n\\[\\sigma_{\\textbf{F}(\\Sigma, \\mathcal{V})}(\\alpha_{1}, \\dotsc , \\alpha_{n})=\\sigma(\\alpha_{1}, \\dotsc , \\alpha_{n}).\\]\n\nWe define the order, or complexity, of an element of $F(\\Sigma, \\mathcal{V})$ as: $0$ if the formula is a variable or a constant, that is, a $\\sigma\\in \\Sigma_{0}$; as $1+\\max\\{p_{1}, \\dotsc , p_{n}\\}$ if the formula is of the form $\\sigma(\\alpha_{1}, \\dotsc , \\alpha_{n})$, with $\\alpha_{j}$ being of order $p_{j}$.\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\\subsection{Multialgebra of non-deterministic formulas}\n\n\\begin{definition}\nGiven a signature $\\Sigma$ and a cardinal $\\kappa>0$, the expanded signature\\index{Signature, Expanded}\\label{expandedsignature} $\\Sigma^{\\kappa}=\\{\\Sigma_{n}^{\\kappa}\\}_{n\\in\\mathbb{N}}$ is the signature such that $\\Sigma_{n}^{\\kappa}=\\Sigma_{n}\\times \\kappa$, where we will denote the pair $(\\sigma, \\beta)$ by $\\sigma^{\\beta}$, for $\\sigma\\in\\Sigma$ and $\\beta\\in\\kappa$. \n\\end{definition}\n\nWe demand that $\\kappa$ is greater than zero: hence, if $\\Sigma$ is non-empty, so is $\\Sigma^{\\kappa}$.\n\n\\begin{definition}\\label{Multialgebra of formulas}\nGiven a set of variables $\\mathcal{V}$, a signature $\\Sigma$ and a cardinal $\\kappa>0$, we define the $\\Sigma$-multialgebra of non-deterministic formulas, or simply multialgebra of formulas\\index{Formulas, Non-deterministic}\\index{Multialgebra of non-deterministic formulas}\\label{mFSigmakappaV}, as\n\\[\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa)=(F(\\Sigma^{\\kappa}, \\mathcal{V}), \\{\\sigma_{\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa)}\\}_{\\sigma\\in\\Sigma})\\]\nwith universe $F(\\Sigma^{\\kappa}, \\mathcal{V})$ and such that, for $\\sigma\\in \\Sigma_{n}$ and $\\alpha_{1}, \\dotsc , \\alpha_{n}\\in {T}(\\Sigma^{\\kappa}, \\mathcal{V})$,\n\\[\\sigma_{\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa)}(\\alpha_{1}, \\dotsc , \\alpha_{n})=\\{\\sigma^{\\beta}(\\alpha_{1}, \\dotsc , \\alpha_{n}) \\ : \\ \\beta\\in\\kappa\\}.\\]\n\\end{definition}\n\nThe intuition behind this definition is that, connecting given formulas $\\alpha_{1}$ through $\\alpha_{n}$ with a connective $\\sigma$ can, in a broader interpretation taking into account non-determinism, return many formulas with the same general shape, that is $\\sigma(\\alpha_{1}, \\dotsc ,\\alpha_{n})$, over which we maintain certain degree of control by counting them, what we achieve by using an index to our connective, $\\sigma^{\\beta}$.\n\nOne can ask why all connectives must return the exact same number of generalized formulas, that is $\\kappa$, but this will not be the case: more useful to us shall be the submultialgebras of $\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa)$, where the cardinality will vary as long as it is bounded by $\\kappa$; we have defined the multialgebras of formulas as above since defining its submultialgebras directly is substantially more difficult.\n\n Here, we will restrict ourselves to the cases where $\\Sigma_{0}\\neq\\emptyset$ or $\\mathcal{V}\\neq\\emptyset$, so that $\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa)$ is always well defined. \n\nWe will understand as the order of an element $\\alpha$ of $\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa)$ simply its order as an element of $F(\\Sigma^{\\kappa}, \\mathcal{V})$. Notice that, if \n\\[\\sigma_{\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa)}(\\alpha_{1}, \\dotsc , \\alpha_{n})\\cap \\theta_{\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa)}(\\beta_{1}, \\dotsc ,\\beta_{m})\\neq\\emptyset,\\]\nthen $\\sigma=\\theta$, $n=m$ and $\\alpha_{1}=\\beta_{1}, \\dotsc , \\alpha_{n}=\\beta_{m}$, since if the intersection is not empty there are $\\beta, \\gamma\\in\\kappa$ such that $\\sigma^{\\beta}(\\alpha_{1}, \\dotsc , \\alpha_{n})= \\theta^{\\gamma}(\\beta_{1}, \\dotsc ,\\beta_{m})$ and by the structure of $F(\\Sigma^{\\kappa}, \\mathcal{V})$ we find that $\\sigma^{\\beta}=\\theta^{\\gamma}$. \n\n\\begin{example}\nThe $\\Sigma$-algebras of formulas $\\textbf{F}(\\Sigma, \\mathcal{V})$, when considered as multialgebras such that $\\sigma_{\\textbf{F}(\\Sigma, \\mathcal{V})}(\\alpha_{1}, \\dotsc , \\alpha_{n})=\\{\\sigma(\\alpha_{1}, \\dotsc ,\\alpha_{n})\\}$, are multialgebras of formulas, with $\\kappa=1$; that is, $\\textbf{F}(\\Sigma, \\mathcal{V})$ and $\\textbf{mF}(\\Sigma, \\mathcal{V}, 1)$ are isomorphic.\n\\end{example}\n\nFrom now on, the cardinal of a set $X$ will be denoted by $|X|$.\\label{|X|}\n\n\\begin{example}\\label{s}\nA directed graph is a pair $(V, A)$, with $V$ a non-empty set of elements called vertices and $A\\subseteq V^{2}$ a set of elements called arrows, where we say that there is an arrow from $u$ to $v$, both in $V$, if $(u,v)\\in A$; we say that the $n$-tuple $(v_{1}, \\dotsc , v_{n})$ is a path between $u$ and $v$ if $u=v_{1}$, $v=v_{n}$ and $(v_{i}, v_{i+1})\\in A$ for every $i\\in\\{1, \\dotsc , n-1\\}$; we say that a vertex $u\\in V$ has a successor if there exists $v\\in V$ such that $(u,v)\\in A$, and $u$ has a predecessor if there exists $v\\in V$ such that $(v,u)\\in A$.\n\nA directed graph $F=(V, A)$ is a forest if, for any two vertices $u, v\\in V$, there exists at most one path between $u$ and $v$, and a forest is said to have height $\\omega$ if every vertex has a successor. Then, we state that forests of height $\\omega$ are in bijection with the submultialgebras of the multialgebras of formulas over the signature $\\Sigma_{s}$\\label{Sigmas} with exactly one operator $s$ of arity $1$.\n\nEssentially, take as $\\mathcal{V}$ the set of elements of $F$ that have no predecessor and define, for $u\\in V$,\n\\[s_{\\mathcal{A}}(u)=\\{v\\in V \\ : \\ (u,v)\\in A\\},\\]\nand we have that the $\\Sigma_{s}$-multialgebra $\\mathcal{A}=(V, \\{s_{\\mathcal{A}}\\})$, submultialgebra of $\\textbf{mF}(\\Sigma_{s}, \\mathcal{V}, |V|)$, carries the same information that $F$.\n\\end{example}\n\n\\begin{example}\nMore generally, a directed multi-graph\\index{Multi-graph} \\cite{ConiglioSernadas}, or directed $m$-graph, is a pair $(V, A)$ with $V$ a non-empty set of vertices and $A$ a subset of $V^{+}\\times V$, where $V^{+}=\\bigcup_{n\\in\\mathbb{N}\\setminus\\{0\\}}V^{n}$ is the set of finite, non-empty, sequences over $V$. We will say that $(v_{1}, \\dotsc , v_{n})$ is a path between $u$ and $v$ if $u=v_{1}$, $v=v_{n}$ and, for every $i\\in\\{1, \\dotsc , n-1\\}$, there exist $v_{i_{1}}, \\dotsc , v_{i_{m}}$ such that $((v_{i_{1}}, \\dotsc , v_{i_{m}}), v_{i+1})$, with $v_{i}=v_{i_{j}}$ for some $j\\in\\{1, \\dotsc , m\\}$.\n\nThen an $m$-forest is a directed $m$-graph such that any two elements are connected by at most one path; and an $m$-forest is said to have $n$-height $\\omega$, for $n\\in\\mathbb{N}\\setminus\\{0\\}$, if, for any $(u_{1}, \\dotsc , u_{n})\\in V^{n}$, there exists $v\\in V$ such that $((u_{1}, \\dotsc , u_{n}), v)\\in A$. Finally, we see that every $m$-forest $F=(V, A)$ with $n$-height $\\omega$, for every $n\\in S\\subseteq\\mathbb{N}\\setminus\\{\\emptyset\\}$, is essentially equivalent to the $\\Sigma_{S}$-multialgebra $\\mathcal{A}=(V, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma_{S}})$, with\n\\[\\sigma_{\\mathcal{A}}(u_{1}, \\dotsc , u_{m})=\\{v\\in V \\ : \\ ((u_{1}, \\dotsc , u_{m}), v)\\in A\\},\\]\nfor $\\sigma$ of arity $m$, and $\\Sigma_{S}$ the signature with exactly one operator of arity $n$, for every $n\\in S$. It is not hard to see that $\\mathcal{A}$ is a submultialgebra of $\\textbf{mF}(\\Sigma_{S}, \\mathcal{V}, |V|)$, with $\\mathcal{V}$ the set of elements $v$ of $V$ such that, for no $(u_{1}, \\dotsc , u_{n})\\in V^{+}$, $((u_{1}, \\dotsc , u_{n}), v)\\in A$.\n\\end{example}\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\\section{Equivalences for being a submultialgebra of $\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa)$}\n\\subsection{Being $\\textbf{cdf}$-generated}\n \nNow, in universal algebra, the algebras of formulas $\\textbf{F}(\\Sigma, \\mathcal{V})$ are absolutely free, also said to be freely generated, also said to be freely generated in the variety of all $\\Sigma$-algebras: this means that there exists a set, in their case the set of variables $\\mathcal{V}$, such that, for every other $\\Sigma$-algebra $\\mathcal{B}$ with universe $B$ and function $f:\\mathcal{V}\\rightarrow B$, there exists an unique homomorphism $\\overline{f}:\\textbf{F}(\\Sigma, \\mathcal{V})\\rightarrow\\mathcal{B}$ extending $f$, essentially defined as:\n\\begin{enumerate}\n\\item $\\overline{f}(p)=f(p)$, for every $p\\in \\mathcal{V}$;\n\\item $\\overline{f}(\\sigma(\\alpha_{1}, \\dotsc , \\alpha_{n}))=\\sigma_{\\mathcal{B}}(\\overline{f}(\\alpha_{1}), \\dotsc , \\overline{f}(\\alpha_{n}))$. \n\\end{enumerate}\nAs we mentioned before, this is no longer true when dealing with multialgebras, but we can define a closely related concept with the aid of what we will call collections of choices.\n\nCollections of choices are motivated by legal valuations, notion first defined in Avron and Lev's seminal paper \\cite{AvronLev} on non-deterministic logical matrices. A map $\\nu$ from $\\textbf{F}(\\Sigma, \\mathcal{V})$ to the universe of a $\\Sigma$-multialgebra $\\mathcal{A}$ is a legal valuation whenever \n\\[\\nu(\\sigma(\\alpha_{1}, \\dotsc ,\\alpha_{n}))\\in \\sigma_{\\mathcal{A}}(\\nu(\\alpha_{1}), \\dotsc , \\nu(\\alpha_{n}));\\]\nessentially, at every formula $\\sigma(\\alpha_{1}, \\dotsc ,\\alpha_{n})$, $\\nu$ ``chooses'' a value from all the possible values\\\\ $\\sigma_{\\mathcal{A}}(\\nu(\\alpha_{1}), \\dotsc , \\nu(\\alpha_{n}))$, possible values which depend themselves on previous choices $\\nu(\\alpha_{1})$ through $\\nu(\\alpha_{n})$ performed by $\\nu$. What a collection of choices does is then to automatize these choices, what justifies its name. \n\n\\begin{definition}\\label{Collection of Choices}\nGiven multialgebras $\\mathcal{A}=(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma})$ and $\\mathcal{B}=(B, \\{\\sigma_{\\mathcal{B}}\\}_{\\sigma\\in\\sigma})$ over the signature $\\Sigma$, a collection of choices\\index{Collection of choices} from $\\mathcal{A}$ to $\\mathcal{B}$ is a collection $C=\\{C_{n}\\}_{n\\in\\mathbb{N}}$ of collections of functions\n\\[C_{n}=\\{C\\sigma_{a_{1}, \\dotsc , a_{n}}^{b_{1}, \\dotsc , b_{n}} : \\sigma\\in\\Sigma_{n}, a_{1}, \\dotsc , a_{n}\\in A, b_{1}, \\dotsc ,b_{n}\\in B\\}\\]\nsuch that, for $\\sigma\\in\\Sigma_{n}$, $a_{1}, \\dotsc , a_{n}\\in A$ and $b_{1}, \\dotsc , b_{n}\\in B$, $C\\sigma_{a_{1}, \\dotsc , a_{n}}^{b_{1}, \\dotsc , b_{n}}$ is a function of the form \n\\[C\\sigma_{a_{1}, \\dotsc , a_{n}}^{b_{1}, \\dotsc , b_{n}}:\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc ,a_{n})\\rightarrow\\sigma_{\\mathcal{B}}(b_{1}, \\dotsc , b_{n}).\\]\n\\end{definition}\n\n\\begin{example}\nIf $\\mathcal{B}$ is actually an algebra (meaning all its operations return singletons), there only exists one collection of choices from any $\\mathcal{A}$ to $\\mathcal{B}$ (since there exists only one function to a set with only one element); this means that in universal algebra, collections of choices are somewhat irrelevant.\n\\end{example}\n\n\\begin{example}\nA directed tree is a directed forest where there exists exactly one element without predecessor; we say that $v$ ramifies from $u$ if there exists an arrow from $u$ to $v$. Then, given two directed trees $T_{1}=(V_{1}, A_{1})$ and $T_{2}=(V_{2}, A_{2})$ of height $\\omega$, seem as $\\Sigma_{s}$-multialgebras, and a collection of choices $C$ from $T_{1}$ to $T_{2}$, for every $v\\in V_{1}$ and $u\\in V_{2}$ the function $Cs_{v}^{u}$ chooses, for each of the elements that ramify from $v$, one element that ramifies from $u$.\n\\end{example}\n\n\\begin{definition}\nGiven a signature $\\Sigma$, a $\\Sigma$-multialgebra $\\mathcal{A}=(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma})$ is choice-dependent freely generated\\index{Choice-dependent freely generated} by $X$ if $X\\subseteq A$ and, for all $\\Sigma$-multialgebras $\\mathcal{B}=(B, \\{\\sigma_{\\mathcal{B}}\\}_{\\sigma\\in\\Sigma})$, all functions $f:X\\rightarrow B$ and all collections of choices $C$ from $\\mathcal{A}$ to $\\mathcal{B}$, there is a unique homomorphism $f_{C}:\\mathcal{A}\\rightarrow \\mathcal{B}$ such that:\n\n\\begin{enumerate} \\item $f_{C}|_{X}=f$;\n\n\\item for all $\\sigma\\in \\Sigma_{n}$ and $a_{1}, \\dotsc , a_{n}\\in A$, \n\\[f_{C}|_{\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})}=C\\sigma_{a_{1}, \\dotsc , a_{n}}^{f_{C}(a_{1}), \\dotsc , f_{C}(a_{n})}.\\]\n\\end{enumerate}\n\\end{definition}\n\nFor simplicity, when $\\mathcal{A}$ is choice-dependent freely generated by $X$, we will write that $\\mathcal{A}$ is $\\textbf{cdf}$-generated\\label{cdfgenerated} by $X$, or merely that $\\mathcal{A}$ is $\\textbf{cdf}$-generated, when the set $X$ is not important.\n\nWe now introduce the concept of ground to indicate what elements of a multialgebra are not ``achieved'', ``reached'' by its multioperations; alternatively, while thinking of formulas and their respective algebras, the ground is the set of indecomposable formulas, that is, variables.\n\n\\begin{definition}\nGiven a $\\Sigma$-multialgebra $\\mathcal{A}=(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma})$, we define its build\\index{Build}\\label{build} as\n\\[B(\\mathcal{A})=\\bigcup \\big\\{\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n}) \\ : \\ n\\in\\mathbb{N}, \\, \\sigma\\in\\Sigma_{n}, \\, a_{1}, \\dotsc , a_{n}\\in A \\big\\}.\\]\nWe define the ground\\index{Ground}\\label{ground} of $\\mathcal{A}$ as\n\\[G(\\mathcal{A})=A\\setminus B(\\mathcal{A}).\\]\n\\end{definition}\n\n\\begin{example}\\label{ground of formulas}\n$B(\\textbf{F}(\\Sigma, \\mathcal{V}))={T}(\\Sigma, \\mathcal{V})\\setminus\\mathcal{V}$ and $G(\\textbf{F}(\\Sigma, \\mathcal{V}))=\\mathcal{V}$.\n\\end{example}\n\n\\begin{example}\nIf $F=(V, A)$ is a directed forest of height $\\omega$, thought as a $\\Sigma_{s}$-multialgebra, its ground is the set of elements $v$ in $V$ without predecessors.\n\\end{example}\n\n\\begin{proposition}\nLet $\\mathcal{A}$ and $\\mathcal{B}$ be $\\Sigma$-multialgebras.\n\\begin{enumerate} \n\\item If $f:\\mathcal{A}\\rightarrow\\mathcal{B}$ is a homomorphism between $\\Sigma$-multialgebras, then \n\\[B(\\mathcal{A})\\subseteq f^{-1}(B(\\mathcal{B}))\\quad\\text{and}\\quad f^{-1}(G(\\mathcal{B}))\\subseteq G(\\mathcal{A}).\\]\n\n\\item If $\\mathcal{B}$ is a submultialgebra of $\\mathcal{A}$, $B(\\mathcal{B})\\subseteq B(\\mathcal{A})$ and $G(\\mathcal{A})\\subseteq G(\\mathcal{B})$.\n\\end{enumerate}\n\\end{proposition}\n\n\\begin{proof}\n\\begin{enumerate} \\item If $a\\in B(\\mathcal{A})$, there exist $\\sigma\\in\\Sigma_{n}$ and $a_{1}, \\dotsc , a_{n}\\in A$ such that $a\\in\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})$. Since $f(\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})) \\subseteq \\sigma_{\\mathcal{B}}(f(a_{1}), \\dotsc ,f(a_{n}))$,\nwe find that $f(a)\\in \\sigma_{\\mathcal{B}}(f(a_{1}), \\dotsc , f(a_{n}))$ and therefore $f(a)\\in B(\\mathcal{B})$, meaning that $a\\in f^{-1}(B(\\mathcal{B}))$. Using that $G(\\mathcal{A})=A\\setminus B(\\mathcal{A})$ and $G(\\mathcal{B})=B\\setminus B(\\mathcal{B})$ we obtain the second mentioned inclusion.\n\n\n\\item If $b\\in B(\\mathcal{B})$, there exist $\\sigma\\in\\Sigma_{n}$ and $b_{1}, \\dotsc , b_{n}\\in B$ such that $b\\in \\sigma_{\\mathcal{B}}(b_{1}, \\dotsc , b_{n})$, and given that $\\sigma_{\\mathcal{B}}(b_{1}, \\dotsc , b_{n})\\subseteq \\sigma_{\\mathcal{A}}(b_{1}, \\dotsc , b_{n})$ we obtain $b\\in B(\\mathcal{A})$. Using again that $G(\\mathcal{A})=A\\setminus B(\\mathcal{A})$ and $G(\\mathcal{B})=B\\setminus B(\\mathcal{B})$ we finish the proof.\n\\end{enumerate}\n\\end{proof}\n\nFrom this it also follows that if $f:\\mathcal{A}\\rightarrow\\mathcal{B}$ is a homomorphism, $G(\\mathcal{B})\\cap f(A)$ is contained in $\\{f(a)\\ : \\ a\\in G(\\mathcal{A})\\}$. Indeed, if $b$ is in $G(\\mathcal{B})\\cap f(A)$, $a\\in A$ such that $f(a)=b$ is in $f^{-1}(G(\\mathcal{B}))$ and, by the previous proposition, it is in $G(\\mathcal{A})$, and therefore $b$ is in $\\{f(a) \\ : \\ a\\in G(\\mathcal{A})\\}$.\n\nGeneralizing Example \\ref{ground of formulas}, we have that $G(\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa))=\\mathcal{V}$: we proceed by induction to show that $B(\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa))={T}(\\Sigma^{\\kappa}, \\mathcal{V})\\setminus \\mathcal{V}$, what is equivalent. If $\\alpha$ is of order $0$, either we have $\\alpha=\\sigma^{\\beta}$, for a $\\sigma\\in\\Sigma_{0}$ and $\\beta\\in\\kappa$, and therefore $\\alpha\\in B(\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa))$; or we have that $\\alpha=p\\in \\mathcal{V}$, and if there exists $\\sigma\\in\\Sigma_{m}$ and $\\alpha_{1}, \\dotsc , \\alpha_{m}\\in {T}(\\Sigma^{\\kappa}, \\mathcal{V})$ such that \n\\[p\\in \\sigma_{\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa)}(\\alpha_{1}, \\dotsc , \\alpha_{m})\\]\nwe have $p=\\sigma^{\\beta}(\\alpha_{1}, \\dotsc ,\\alpha_{m})$, for $\\beta\\in\\kappa$, which is absurd given the structure of $F(\\Sigma^{\\kappa}, \\mathcal{V})$, forcing us to conclude that $x\\notin B(\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa))$. If $\\alpha$ is of order $n+1$, we have that $\\alpha=\\sigma^{\\beta}(\\alpha_{1}, \\dotsc ,\\alpha_{m})$ for a $\\sigma\\in\\Sigma_{m}$, $\\beta\\in\\kappa$ and $\\alpha_{1}, \\dotsc , \\alpha_{m}$ of order at most $n$, and therefore we have $\\alpha$ in $\\sigma_{\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa)}(\\alpha_{1}, \\dotsc , \\alpha_{m})$, meaning that $\\alpha\\in B(\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa))$.\n\n\\begin{definition}\nGiven a $\\Sigma$-multialgebra $\\mathcal{A}=(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma})$ and a set $S\\subseteq A$, we define the sets $\\langle S\\rangle_{m}$\\label{mth generated} by induction: $\\langle S\\rangle _{0}=S\\cup\\bigcup_{\\sigma\\in \\Sigma_{0}}\\sigma_{\\mathcal{A}}$; and assuming we have defined $\\langle S\\rangle_{m}$, we make\n\\[\\langle S\\rangle_{m+1}=\\langle S\\rangle_{m}\\cup \\bigcup \\big\\{\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n}) \\ : \\ n\\in\\mathbb{N}, \\, \\sigma\\in\\Sigma_{n}, \\, a_{1}, \\dotsc , a_{n}\\in \\langle S\\rangle_{m} \\big\\}.\\]\nThe set generated\\index{Generated}\\label{generated} by $S$, denoted by $\\langle S\\rangle$, is then defined as $\\langle S\\rangle=\\bigcup_{m\\in\\mathbb{N}}\\langle S\\rangle_{m}$.\n\nWe say $\\mathcal{A}$ is generated by $S$ if $\\langle S\\rangle=A$.\n\\end{definition}\n\n\n\n\\begin{lemma}\\label{sub mF is gen ground}\nEvery submultialgebra $\\mathcal{A}$ of $\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa)$ is generated by $G(\\mathcal{A})$.\n\\end{lemma}\n\n\\begin{proof}\nSuppose $a$ is an element of $\\mathcal{A}$ not contained in $\\langle G(\\mathcal{A})\\rangle$ of minimum order: since $a$ cannot belong to $G(\\mathcal{A})\\cup\\bigcup_{\\sigma\\in\\Sigma_{0}}\\sigma_{\\mathcal{A}}=\\langle G(\\mathcal{A})\\rangle_{0}$, there exist $n>0$, $\\sigma\\in\\Sigma_{n}$ and $a_{1}, \\dotsc , a_{n}\\in A$ such that $a\\in\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})$.\n\nSince $\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})\\subseteq \\sigma_{\\textbf{mF}(\\Sigma, X, \\kappa)}(a_{1}, \\dotsc , a_{n})$ we derive that $a_{1}$ to $a_{n}$ are of order less than that of $a$: by our hypothesis, there must exist $m_{1}, \\dotsc , m_{n}$ such that $a_{j}\\in \\langle G(\\mathcal{A})\\rangle_{m_{j}}$ for all $j\\in\\{1, \\dotsc , n\\}$; taking $m=\\max\\{m_{1}, \\dotsc , m_{n}\\}$, $a_{1}, \\dotsc , a_{n}\\in \\langle G(\\mathcal{A})\\rangle_{m}$, and therefore\n\\[a\\in \\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})\\subseteq \\langle G(\\mathcal{A})\\rangle_{m+1},\\]\nwhich contradicts our assumption and proves the lemma.\n\\end{proof}\n\n\\begin{theorem}\\label{sub mF is cdf-gen ground}\nEvery submultialgebra $\\mathcal{A}$ of $\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa)$ is $\\textbf{cdf}$-generated by $G(\\mathcal{A})$.\n\\end{theorem}\n\n\\begin{proof}\nLet $\\mathcal{A}=(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma})$ be a submultialgebra of $\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa)$, let $\\mathcal{B}=(B, \\{\\sigma_{\\mathcal{B}}\\}_{\\sigma\\in\\Sigma})$ be any $\\Sigma$-multialgebra, let $f:G(\\mathcal{A})\\rightarrow B$ be a function and $C$ a collection of choices from $\\mathcal{A}$ to $\\mathcal{B}$. We define $f_{C}:\\mathcal{A}\\rightarrow\\mathcal{B}$ by induction on $\\langle G(\\mathcal{A})\\rangle_{m}$:\n\n\\begin{enumerate} \\item if $a\\in \\langle G(\\mathcal{A})\\rangle_{0}$ and $a\\in G(\\mathcal{A})$, we define $f_{C}(a)=f(a)$;\n\n\\item if $a\\in \\langle G(\\mathcal{A})\\rangle_{0}$ and $a\\in\\sigma_{\\mathcal{A}}$, for a $\\sigma\\in\\Sigma_{0}$, we define $f_{C}(a)=C\\sigma(a)$;\n\n\\item if $f_{C}$ is defined for all elements of $\\langle G(\\mathcal{A})\\rangle_{m}$, $a_{1}, \\dotsc , a_{n}\\in \\langle G(\\mathcal{A})\\rangle_{m}$ and $\\sigma\\in\\Sigma_{n}$, for every element $a\\in \\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})$ we define \n\\[f_{C}(a)=C\\sigma_{a_{1}, \\dotsc , a_{n}}^{f_{C}(a_{1}), \\dotsc , f_{C}(a_{n})}(a).\\]\n\\end{enumerate}\n\nFirst, we must prove that $f_{C}$ is well defined. There are two possibly problematic cases to consider for an element $a\\in A$: the one in which $a\\in G(\\mathcal{A})$ and there are $\\sigma\\in\\Sigma_{n}$ and $a_{1}, \\dotsc , a_{n}\\in A$ for which $a\\in\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})$, corresponding to $a$ falling simultaneously in the cases $(1)$ and $(2)$, or $(1)$ and $(3)$ of the definition; and the one where there are $\\sigma\\in\\Sigma_{n}$, $\\theta\\in\\Sigma_{m}$, $a_{1}, \\dotsc , a_{n}\\in A$ and $b_{1}, \\dotsc , b_{m}\\in A$ such that $a\\in\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})$ and $a\\in\\theta_{\\mathcal{A}}(b_{1}, \\dotsc , b_{m})$, corresponding to the cases $(2)$ and $(3)$, $(2)$ and $(2)$, or $(3)$ and $(3)$ occurring simultaneously.\n\nThe first case is not possible, since $G(\\mathcal{A})\\subseteq A\\setminus\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})$ for every $\\sigma\\in\\Sigma_{n}$ and $a_{1}, \\dotsc , a_{n}\\in A$; in the second case, we find that \n\\[a\\in \\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})\\cap\\theta_{\\mathcal{A}}(b_{1}, \\dotsc , b_{m})\\subseteq\\sigma_{\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa)}(a_{1}, \\dotsc , a_{n})\\cap \\theta_{\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa)}( b_{1}, \\dotsc , b_{m}),\\]\nso $n=m$, $\\sigma=\\theta$ and $a_{1}=b_{1}, \\dotsc , a_{n}=b_{m}$, and therefore $f_{C}(a)$ is well-defined.\n\nSecond, we must prove that $f_{C}$ is defined over all of $A$: that is simple, for $f_{C}$ is defined over all of $\\langle G(\\mathcal{A})\\rangle$ and we established in Lemma~\\ref{sub mF is gen ground} that $A=\\langle G(\\mathcal{A})\\rangle$.\n\nSo $f_{C}:A\\rightarrow B$ is a well-defined function: it remains to be shown that it is a homomorphism; given $\\sigma\\in\\Sigma_{n}$ and $a_{1}, \\dotsc , a_{n}$, we see that\n\\[f_{C}(\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n}))=\\big\\{C\\sigma_{a_{1}, \\dotsc , a_{n}}^{f_{C}(a_{1}), \\dotsc , f_{C}(a_{n})}(a)\\ : \\ a\\in\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})\\big\\}\\subseteq\\sigma_{\\mathcal{B}}(f_{C}(a_{1}), \\dotsc , f_{C}(a_{n})),\\]\nwhile we also have that $f_{C}$ clearly extends both $f$ and all $C\\sigma_{a_{1}, \\dotsc , a_{n}}^{f_{C}(a_{1}, \\dotsc , f_{C}(a_{n})}$. \n\nTo finish the proof, suppose $g:\\mathcal{A}\\rightarrow\\mathcal{B}$ is another homomorphism extending both $f$ and all $C\\sigma_{a_{1}, \\dotsc , a_{n}}^{g(a_{1}, \\dotsc , g(a_{n})}$: we will prove that $g=f_{C}$ again by induction on the $m$ of $\\langle G(\\mathcal{A})\\rangle_{m}$. For $m=0$, an element $a\\in \\langle G(\\mathcal{A})\\rangle_{0}$ is either in $G(\\mathcal{A})$, when we have $g(a)=f(a)=f_{C}(a)$, or in $\\sigma_{\\mathcal{A}}$ for a $\\sigma\\in\\Sigma_{0}$, when $g(a)=C\\sigma(a)=f_{C}(a)$.\n\nSuppose $g$ is equal to $f_{C}$ in $\\langle G(\\mathcal{A})\\rangle_{m}$ and take an $a\\in \\langle G(\\mathcal{A})\\rangle_{m+1}\\setminus \\langle G(\\mathcal{A})\\rangle_{m}$: we have that there exist $\\sigma\\in\\Sigma_{n}$ and $a_{1}, \\dotsc , a_{n}\\in \\langle G(\\mathcal{A})\\rangle_{m}$ such that $a\\in\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})$, and then \n\\[g(a)=C\\sigma_{a_{1}, \\dotsc , a_{n}}^{g(a_{1}), \\dotsc , g(a_{n})}(a)=C\\sigma_{a_{1}, \\dotsc , a_{n}}^{f_{C}(a_{1}), \\dotsc , f_{C}(a_{n})}(a)=f_{C}(a),\\]\nproving that $g=f_{C}$ and that, in fact, $f_{C}$ is unique. This means that $\\mathcal{A}$ is $\\textbf{cdf}$-generated by $G(\\mathcal{A})$.\n\\end{proof}\n\nThe following lemma may be found in section 2 of \\cite{CFG}.\n\n\\begin{lemma}\\label{image of hom is sub}\nIf $\\mathcal{A}=(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma})$ and $\\mathcal{B}=(B, \\{\\sigma_{\\mathcal{B}}\\}_{\\sigma\\in\\Sigma})$ are $\\Sigma$-multialgebras and $f:\\mathcal{A}\\rightarrow\\mathcal{B}$ is a homomorphism, $\\mathcal{C}=(f(A), \\{\\sigma_{\\mathcal{C}}\\}_{\\sigma\\in\\Sigma})$ such that\n\\[\\sigma_{\\mathcal{C}}(c_{1}, \\dotsc , c_{n})=\\bigcup\\{f(\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})) : f(a_{1})=c_{1}, \\dotsc , f(a_{n})=c_{n}\\}\\]\nis a $\\Sigma$-submultialgebra of $\\mathcal{B}$, while $f:\\mathcal{A}\\rightarrow\\mathcal{C}$ is an epimorphism.\\footnote{A epimorphism between $\\Sigma$-multialgebras $\\mathcal{A}=(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma})$ and $\\mathcal{B}=(B, \\{\\sigma_{\\mathcal{B}}\\}_{\\sigma\\in\\Sigma})$ is defined, as usual, as a homomorphism $\\varphi:A\\rightarrow B$ between $\\mathcal{A}$ and $\\mathcal{B}$ that is surjective.} The $\\Sigma$-multialgebra $\\mathcal{C}$ is known as the direct image\\index{Direct image} of $\\mathcal{A}$ trough $f$.\n\\end{lemma}\n\n\\begin{proof}\nFirst of all, $A$ is not empty, and therefore so is $f(A)$.\n\nTake $c_{1}, \\dotsc , c_{n}\\in f(A)$: there must exist $a_{1}, \\dotsc , a_{n}\\in A$ such that $f(a_{1})=c_{1}, \\dotsc ,$\\\\ $f(a_{n})=c_{n}$, and since $\\mathcal{A}$ is a multialgebra, $\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})\\neq\\emptyset$, implying that $f(\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n}))$ is not empty and, therefore, that $\\sigma_{\\mathcal{C}}(c_{1}, \\dotsc , c_{n})$ is non-empty, given it contains $f(\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n}))$. It is obvious that, as defined, $\\sigma_{\\mathcal{C}}(c_{1}, \\dotsc , c_{n})$ is a subset of $f(A)$, and so we can deduce that $\\mathcal{C}$ is a $\\Sigma$-multialgebra.\n\nNow, given $f:\\mathcal{A}\\rightarrow\\mathcal{B}$ is a homomorphism, for all $a_{1}, \\dotsc , a_{n}\\in A$ we have $f(\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n}))\\subseteq \\sigma_{\\mathcal{B}}(f(a_{1}), \\dotsc , f(a_{n}))$, meaning\n\\[\\sigma_{\\mathcal{C}}(c_{1}, \\dotsc , c_{n})=\\bigcup\\{f(\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})) : f(a_{1})=c_{1}, \\dotsc , f(a_{n})=c_{n}\\}\\subseteq\\]\n\\[\\bigcup\\{\\sigma_{\\mathcal{B}}(f(a_{1}), \\dotsc , f(a_{n})) : f(a_{1})=c_{1}, \\dotsc , f(a_{n})=c_{n}\\}=\\sigma_{\\mathcal{B}}(c_{1}, \\dotsc , c_{n}),\\]\nor what is equivalent, that $\\mathcal{C}$ is a submultialgebra of $\\mathcal{B}$.\n\nFinally, $f:A\\rightarrow f(A)$ is still a well-defined function, obviously surjective: for any $n$-ary $\\sigma$ and elements $a_{1}$ through $a_{n}$ of $A$, one has\n\\[f(\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n}))\\subseteq \\bigcup\\{f(\\sigma_{\\mathcal{A}}(a'_{1}, \\dotsc , a'_{n})) : f(a'_{1})=f(a_{1}), \\dotsc , f(a'_{n})=f(a_{n})\\}=\\]\n\\[\\sigma_{\\mathcal{C}}(f(a_{1}), \\dotsc , f(a_{n}))\\]\nand that, in conclusion, $f$ is a homomorphism.\n\\end{proof}\n\n\\begin{theorem}\\label{cdf-gen is iso to sub mF}\nIf the multialgebra $\\mathcal{A}=(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma})$ over $\\Sigma$ is $\\textbf{cdf}$-generated by $X$, then $\\mathcal{A}$ is isomorphic to a submultialgebra of $\\textbf{mF}(\\Sigma, X, |A|)$ containing $X$.\n\\end{theorem}\n\n\\begin{proof}\nTake $f:X\\rightarrow F(\\Sigma^{|A|}, X)$ to be the inclusion (such that $f(x)=x$), and take a collection of choices $C$ such that, for $\\sigma\\in\\Sigma_{n}$, $a_{1}, \\dotsc , a_{n}\\in A$ and $\\alpha_{1}, \\dotsc , \\alpha_{n}\\in F(\\Sigma^{|A|}, X)$, \n\\[C\\sigma_{a_{1}, \\dotsc , a_{n}}^{\\alpha_{1}, \\dotsc , \\alpha_{n}}:\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})\\rightarrow\\sigma_{\\textbf{mF}(\\Sigma, X, |A|)}(\\alpha_{1}, \\dotsc , \\alpha_{n})\\]\nis an injective function. Such collection of choices exist since $\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})\\subseteq A$ and\\\\ $\\sigma_{\\textbf{mF}(\\Sigma, X, |A|)}(\\alpha_{1}, \\dotsc , \\alpha_{n})$ is of cardinality $|A|$. Now, since $\\mathcal{A}$ is $\\textbf{cdf}$-generated by $X$, there exists a homomorphism $f_{C}:\\mathcal{A}\\rightarrow\\textbf{mF}(\\Sigma, X, |A|)$ extending $f$ and each $C\\sigma_{a_{1}, \\dotsc , a_{n}}^{f_{C}(a_{1}), \\dotsc , f_{C}(a_{n})}$. \n\n\nLet $\\mathcal{B}=(f_{C}(A), \\{\\sigma_{\\mathcal{B}}\\}_{\\sigma\\in\\Sigma})$ be the direct image of $\\mathcal{A}$ trough $f_{C}$, so that $f_{C}:\\mathcal{A}\\rightarrow\\mathcal{B}$ is an epimorphism, what is possible given Lemma \\ref{image of hom is sub}: notice too that \n\\[X=X\\cap f_{C}(A)=G(\\textbf{mF}(\\Sigma, X, |A|))\\cap f_{C}(A)\\subseteq G(\\mathcal{B})\\]\nbecause $\\mathcal{B}$ is a submultialgebra of $\\textbf{mF}(\\Sigma, X, |A|)$. Take any $g:G(\\mathcal{B})\\rightarrow A$ such that $g(x)=x$, for every $x\\in X$. And take a collection of choices $D$ from $\\mathcal{B}$ to $\\mathcal{A}$ such that, for any $\\sigma\\in\\Sigma_{n}$, $b_{1}, \\dotsc , b_{n}\\in f_{C}(A)$ and $a_{1}, \\dotsc , a_{n}\\in A$, the function\n\\[D\\sigma_{b_{1}, \\dotsc , b_{n}}^{a_{1}, \\dotsc , a_{n}}:\\sigma_{\\mathcal{B}}(b_{1}, \\dotsc , b_{n})\\rightarrow\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})\\]\nsatisfies that, if $a \\in \\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})$ is such that $C\\sigma_{a_{1}, \\dotsc , a_{n}}^{b_{1}, \\dotsc , b_{n}}(a) \\in \\sigma_{\\mathcal{B}}(b_{1}, \\dotsc , b_{n})$, then\\\\ $D\\sigma_{b_{1}, \\dotsc , b_{n}}^{a_{1}, \\dotsc , a_{n}}(C\\sigma_{a_{1}, \\dotsc , a_{n}}^{b_{1}, \\dotsc , b_{n}}(a))=a$. Given that $C\\sigma_{a_{1}, \\dotsc , a_{n}}^{b_{1}, \\dotsc , b_{n}}$ is injective, this condition is well-defined.\n\nSince $\\mathcal{B}$ is $\\textbf{cdf}$-generated by $G(\\mathcal{B})$, we know to exist a homomorphism $g_{D}:\\mathcal{B}\\rightarrow \\mathcal{A}$ extending $g$ and the functions $D\\sigma_{b_{1}, \\dotsc , b_{n}}^{g_{D}(b_{1}), \\dotsc , g_{D}(b_{n})}$.\n\nFinally, we take $g_{D}\\circ f_{C}:\\mathcal{A}\\rightarrow\\mathcal{A}$: it extends the injection $id=g\\circ f:X\\rightarrow A$, for which $id(x)=x$; it also extends the collection of choices $E$ defined by\n\\[E\\sigma_{a_{1}, \\dotsc , a_{n}}^{a'_{1}, \\dotsc , a'_{n}}=D\\sigma_{f_{C}(a_{1}), \\dotsc , f_{C}(a_{n})}^{a'_{1}, \\dotsc , a'_{n}}\\circ C_{a_{1}, \\dotsc , a_{n}}^{f_{C}(a_{1}), \\dotsc , f_{C}(a_{n})}:\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})\\rightarrow\\sigma_{\\mathcal{A}}(a'_{1}, \\dotsc , a'_{n}),\\]\nfor $\\sigma\\in\\Sigma_{n}$ and $a_{1}, \\dotsc a_{n}, a'_{1}, \\dotsc , a'_{n}\\in A$. \nThis way, $E\\sigma_{a_{1}, \\dotsc , a_{n}}^{a_{1}, \\dotsc , a_{n}}$ is the identity on $\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})$: indeed, for any $a\\in \\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})$, \n\\[C\\sigma_{a_{1}, \\dotsc , a_{n}}^{f_{C}(a_{1}), \\dotsc , f_{C}(a_{n})}(a)=f_{C}(a)\\]\nby definition of $f_{C}$, and, given $f_{C}$ is a homomorphism, $f_{C}(a)\\in \\sigma_{\\mathcal{B}}(f_{C}(a_{1}), \\dotsc , f_{C}(a_{n}))$, meaning\\\\ $C\\sigma_{a_{1}, \\dotsc , a_{n}}^{f_{C}(a_{1}), \\dotsc , f_{C}(a_{n})}(a) \\in \\sigma_{\\mathcal{B}}(f_{C}(a_{1}), \\dotsc , f_{C}(a_{n}))$; then\n\\[E\\sigma_{a_{1}, \\dotsc , a_{n}}^{a_{1}, \\dotsc , a_{n}}(a) = D\\sigma_{f_{C}(a_{1}), \\dotsc , f_{C}(a_{n})}^{a_{1}, \\dotsc , a_{n}}(C_{a_{1}, \\dotsc , a_{n}}^{f_{C}(a_{1}), \\dotsc , f_{C}(a_{n})}(a)) = a\\]\nby definition of $D$.\n\nBut notice that the identical homomorphism $Id_{\\mathcal{A}}:\\mathcal{A}\\rightarrow \\mathcal{A}$ also extends both $id$ and $E$ and, given the unicity of such extensions on the definition of being $\\textbf{cdf}$-generated, we obtain that $Id_{\\mathcal{A}}=g_{D}\\circ f_{C}$. Of course, the fact that $f_{C}:\\mathcal{A}\\rightarrow\\mathcal{B}$ has a left inverse implies that is injective, and by definition of $\\mathcal{B}$ it is also surjective, meaning it is a bijective function; moreover, $g_{D}$ is the inverse function of $f_{C}$. Finally, for $\\sigma\\in \\Sigma_{n}$ and $a_{1}, \\dotsc , a_{n}\\in A$,\n\\[f_{C}(\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n}))\\subseteq \\sigma_{\\mathcal{B}}(f_{C}(a_{1}), \\dotsc , f_{C}(a_{n})),\\]\nsince $f_{C}$ is a homomorphism; however, given $g_{D}$ is also a homomorphism.\n\\[g_{D}(\\sigma_{\\mathcal{B}}(f_{C}(a_{1}), \\dotsc , f_{C}(a_{n})))\\subseteq \\sigma_{\\mathcal{A}}(g_{D}\\circ f_{C}(a_{1}), \\dotsc , g_{D}\\circ f_{C}(a_{n}))=\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n}),\\]\nand by applying $f_{C}$ to both sides, one obtains\n\\[\\sigma_{\\mathcal{B}}(f_{C}(a_{1}), \\dotsc , f_{C}(a_{n}))=f_{C}(g_{D}(\\sigma_{\\mathcal{B}}(f_{C}(a_{1}), \\dotsc , f_{C}(a_{n}))))\\subseteq f_{C}(\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})),\\]\nwhat proves $f_{C}$ is a full homomorphism, that is, an isomorphism.\n\\end{proof}\n\nNotice that, \\textit{mutatis mutandis}, the previous proof shows that if $\\mathcal{A}=(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma})$ is $\\textbf{cdf}$-generated by $X$, then $\\mathcal{A}$ is isomorphic to a submultialgebra of $\\textbf{mF}(\\Sigma, X, M(\\mathcal{A}))$, where\n\\[M(\\mathcal{A})=\\sup \\big\\{|\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})| \\ : \\ n\\in\\mathbb{N}, \\, \\sigma\\in\\Sigma_{n}, \\, a_{1}, \\dotsc , a_{n}\\in A \\big\\}.\\]\n\\label{M(A)}Quite obviously, $M(\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa))=\\kappa$ and $M(\\mathcal{A})\\leq \\kappa$ for any submultialgebra of $\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa)$. The value $M(\\mathcal{A})$ has been regarded for some time, in the literature of multialgebras, as one of their fundamental aspects, see for example ~\\cite{CuponaMadarasz}; unfortunately, their definition of homomorphism is grossly different from ours, meaning their results are not applicable to or studies.\n\nNotice, furthermore, that written in classical terms, the previous Theorems~\\ref{sub mF is cdf-gen ground} and~\\ref{cdf-gen is iso to sub mF} state something quite well known: an algebra is absolutely free if, and only if, it is isomorphic to some algebra of formulas over the same signature.\n\n\\begin{corollary}\\label{cdf-gen is cdf-gen by ground}\nEvery $\\textbf{cdf}$-generated multialgebra $\\mathcal{A}$ is generated by its ground $G(\\mathcal{A})$.\n\\end{corollary}\n\n\\begin{proof}\nSince every $\\textbf{cdf}$-generated multialgebra is isomorphic to a submultialgebra of some\\\\ $\\textbf{mF}(\\Sigma, X, \\kappa)$, from Theorem \\ref{cdf-gen is iso to sub mF}, and every submultialgebra of $\\textbf{mF}(\\Sigma, X, \\kappa)$ is generated by its ground, by Lemma \\ref{sub mF is gen ground}, the result follows.\n\\end{proof}\n\n\\begin{corollary}\nEvery $\\textbf{cdf}$-generated multialgebra $\\mathcal{A}$ is $\\textbf{cdf}$-generated by its ground $G(\\mathcal{A})$.\n\\end{corollary}\n\n\\begin{definition}\nA $\\Sigma$-multialgebra $\\mathcal{A}=(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma})$ is said to be disconnected\\index{Disconnected} if, for every $\\sigma\\in\\Sigma_{n}, \\theta\\in \\Sigma_{m}$, $a_{1}, \\dotsc , a_{n}, b_{1}, \\dotsc , b_{m}\\in A$,\n\\[\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})\\cap \\theta_{\\mathcal{A}}(b_{1}, \\dotsc , b_{m})\\neq\\emptyset\\]\nimplies that $n=m$, $\\sigma=\\theta$ and $a_{1}=b_{1}, \\dotsc , a_{n}=b_{m}$.\n\\end{definition}\n\n\n\\begin{example}\n$\\textbf{F}(\\Sigma, \\mathcal{V})$ is disconnected.\n\\end{example}\n\n\\begin{example}\nAll directed forests of height $\\omega$, when considered as $\\Sigma_{s}$-multialgebras, are disconnected, given that no two arrows point to the same element.\n\\end{example}\n\nIt is clear that if $\\mathcal{B}$ is a submultialgebra of $\\mathcal{A}$ and $\\mathcal{A}$ is disconnected, then $\\mathcal{B}$ is also disconnected, since if $\\sigma_{\\mathcal{B}}(a_{1}, \\dotsc , a_{n})\\cap \\theta_{\\mathcal{B}}(b_{1}, \\dotsc , b_{m})\\neq\\emptyset$, for $a_{1}, \\dotsc , a_{n}, b_{1}, \\dotsc , b_{m}\\in B$, given that $\\sigma_{\\mathcal{B}}(a_{1}, \\dotsc , a_{n})\\subseteq\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})$ and $\\theta_{\\mathcal{B}}(b_{1}, \\dotsc , b_{m})\\subseteq \\theta_{\\mathcal{A}}(b_{1}, \\dotsc , b_{m})$, we find that \n\\[\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})\\cap \\theta_{\\mathcal{A}}(b_{1}, \\dotsc , b_{m})\\neq\\emptyset\\]\nand therefore $n=m$, $\\sigma=\\theta$ and $a_{1}=b_{1}, \\dotsc , a_{n}=b_{m}$.\n\nWe noticed before that $\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa)$ is disconnected, and by Theorem \\ref{cdf-gen is iso to sub mF} we obtain that every $\\textbf{cdf}-$generated algebra is disconnected. This has a deeper meaning: being disconnected is an attempt to measure how free of identities a multialgebra is. After all, having no two multioperations to coincide, on any elements, is strongly indicative that the multialgebra does not satisfy any identities. The fact that our candidates for an absolutely free multialgebra (the submultialgebras of $\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa)$) are all disconnected suggests that they are well-deserving of the ``weakly free'' title.\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\\subsection{Being disconnected and generated by the ground}\n\nWe have offered, so far, two characterizations of the multialgebras we chose to call weakly free: first of all, they are submultialgebras of some $\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa)$, and perhaps this is to be taken as their definition; second, they are $\\textbf{cdf}$-generated. Now, we look at other possible characterizations of being weakly free that could lead to possible future definitions of relatively free multialgebras. \n\nAlgebras of formulas satisfy no identities, what would partially correspond here to the concept of being disconnected. However, there is one property that is maybe more representative of our intuition of formulas (which are, up to isomorphism, the elements of all absolutely free algebras): whenever one deals with formulas, one starts by defining them from elements that are as simple as possible (variables), and continues indefinitely by combining them trough operations (connectives). \n\nThe concept of simplest (or indecomposable) element, here, is replaced by that of being an element of the ground, so one would expect that being generated by it plays some sort of role in the objects we have defined so far: a multialgebra which is generated by a set has all of its elements either on the set, or as the result of increasingly more complex (multi)operations performed on that very set.\n\n\n\\begin{lemma}\\label{X in ground if cdf-gen by X}\nIf $\\mathcal{A}$ is $\\textbf{cdf}$-generated by $X$, then $X\\subseteq G(\\mathcal{A})$.\n\\end{lemma}\n\n\\begin{proof}\nIf $\\mathcal{A}$ is $\\textbf{cdf}$-generated by $X$, then $\\mathcal{A}$ is isomorphic to a submultialgebra of $\\textbf{mF}(\\Sigma, X, |A|)$ containing $X$, from Theorem \\ref{cdf-gen is iso to sub mF}: let us assume that $\\mathcal{A}$ is equal to this submultialgebra, without loss of generality.\n\nThen we have $X=G(\\textbf{mF}(\\Sigma, X, |A|))\\cap A\\subseteq G(\\mathcal{A})$.\n\\end{proof}\n\n\n\\begin{lemma}\\label{no proper cdf-gen sets}\nIf $\\mathcal{A}$ is $\\textbf{cdf}$-generated by both $X$ and $Y$, with $X\\subseteq Y$, then $X=Y$.\n\\end{lemma}\n\n\\begin{proof}\nSuppose $X\\neq Y$ and let $y\\in Y\\setminus X$: take a $\\Sigma$-multialgebra $\\mathcal{B}$ over the same signature as $\\mathcal{A}$ such that $|B|\\geq 2$, and a collection of choices $C$ from $\\mathcal{A}$ to $\\mathcal{B}$.\n\nTake also two functions $g, h:Y\\rightarrow B$ such that $g|_{X}=h|_{X}$ and $g(y)\\neq h(y)$, which is possible since $|B|\\geq 2$: given that $\\mathcal{A}$ is $\\textbf{cdf}$-generated by $Y$, there exist unique homomorphisms $g_{C}$ and $h_{C}$ extending both, respectively, $g$ and $C$ and $h$ and $C$.\n\nHowever, $g_{C}$ and $h_{C}$ extend both $g|_{X}:X\\rightarrow B$ and $C$, and since $\\mathcal{A}$ is $\\textbf{cdf}$-generated by $X$, we find that $g_{C}=h_{C}$. This is not possible, since $g_{C}(y)\\neq h_{C}(y)$, what must imply that $Y\\setminus X=\\emptyset$ and therefore $X=Y$.\n\\end{proof}\n\n\\begin{theorem}\\label{ground is only cdf-gen set}\nEvery $\\textbf{cdf}$-generated multialgebra $\\mathcal{A}$ is uniquely $\\textbf{cdf}$-generated by its ground.\n\\end{theorem}\n\n\\begin{proof}\nFrom Theorem \\ref{cdf-gen is cdf-gen by ground}, $\\mathcal{A}$ is $\\textbf{cdf}$-generated by $G(\\mathcal{A})$, and from Lemma \\ref{X in ground if cdf-gen by X}, if $\\mathcal{A}$ is also $\\textbf{cdf}$-generated by $X$, then $X\\subseteq G(\\mathcal{A})$. By Lemma \\ref{no proper cdf-gen sets}, this implies that $X=G(\\mathcal{A})$.\n\\end{proof}\n\n\nWe have proved so far that, if $\\mathcal{A}$ is $\\textbf{cdf}$-generated, then $\\mathcal{A}$ is generated by its ground and disconnected. We would like to prove that this is enough to characterize a $\\textbf{cdf}$-generated multialgebra: that is, if $\\mathcal{A}$ is generated by its ground and disconnected, then it is $\\textbf{cdf}$-generated, exactly by its ground.\n\nThe idea is similar to the one we used to prove, in Theorem \\ref{sub mF is cdf-gen ground}, that all submultialgebras of $\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa)$ are $\\textbf{cdf}$-generated: take a multialgebra $\\mathcal{A}$ that is both generated by its ground $G(\\mathcal{A})$, which we will denote by $X$, and disconnected, and fix a multialgebra $\\mathcal{B}$ over the same signature, a function $f:X\\rightarrow B$ and a collection of choices $C$ from $\\mathcal{A}$ to $\\mathcal{B}$.\n\nWe define a function $f_{C}:A\\rightarrow B$ using induction on the $\\langle X\\rangle_{m}$: for $m=0$, either we have an element $x\\in X$, when we define $f_{C}(x)=f(x)$, or we have an $a\\in\\sigma_{\\mathcal{A}}$, for a $\\sigma\\in\\Sigma_{0}$, when we define $f_{C}(a)=C\\sigma(a)$. Notice how, up to this point, there are no contradictions on the definition, given an element cannot belong both to $X$ and to a $\\sigma_{\\mathcal{A}}$, since $X=G(\\mathcal{A})$.\n\nSuppose we have successfully defined $f_{C}$ on $\\langle X\\rangle_{m}$ and take an $a\\in \\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})$ for $a_{1}, \\dotsc , a_{n}\\in\\langle X\\rangle_{m}$. We then define \n\\[f_{C}(a)=C\\sigma_{a_{1}, \\dotsc , a_{n}}^{f_{C}(a_{1}), \\dotsc , f_{C}(a_{n})}(a).\\]\nAgain the function remains well-defined: $a$ cannot belong to $X$, since $X=G(\\mathcal{A})$, and cannot belong to a $\\theta_{\\mathcal{A}}(b_{1}, \\dotsc , b_{p})$ unless $p=n$, $\\theta=\\sigma$ and $b_{1}=a_{1}, \\dotsc ,b_{p}=a_{n}$, since $\\mathcal{A}$ is disconnected.\n\nClearly $f_{C}$ is a homomorphism, since the image of $\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})$ under $f_{C}$ is contained in $\\sigma_{\\mathcal{B}}(f_{C}(a_{1}), \\dotsc , f_{C}(a_{n}))$, and $f_{C}$ extends both $f$ and $C$.\n\n\\begin{lemma}\\label{gen by ground and disc implies cdf-gen}\nIf a multialgebra $\\mathcal{A}$ is both generated by its ground $X$ and disconnected, $\\mathcal{A}$ is $\\textbf{cdf}$-generated by $X$.\n\\end{lemma}\n\n\\begin{proof}\nIt remains for us to show that $f_{C}$, as defined above, is the only homomorphism extending $f$ and $C$. Suppose $g$ is another such homomorphism and we shall proceed yet again by induction.\n\nOn $\\langle X\\rangle_{0}$, we have that $f_{C}(x)=f(x)=g(x)$ for all $x\\in X$; and for $a\\in\\sigma_{\\mathcal{A}}$ and $\\sigma\\in\\Sigma_{0}$ we have that \n\\[f_{C}(a)=C\\sigma(a)=g(a),\\]\nhence $f_{C}$ and $g$ coincide on $\\langle X\\rangle_{0}$. Suppose that $f_{C}$ and $g$ are equal on $\\langle X\\rangle_{m}$ and take $a\\in\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})$ for $a_{1}, \\dotsc , a_{n}\\in \\langle X\\rangle_{m}$: we have by induction hypothesis that\n\\[f_{C}(a)=C\\sigma_{a_{1}, \\dotsc , a_{n}}^{f_{C}(a_{1}), \\dotsc , f_{C}(a_{n})}(a)=C\\sigma_{a_{1}, \\dotsc , a_{n}}^{g(a_{1}), \\dotsc , g(a_{n})}(a)=g(a),\\]\nwhich concludes our proof.\n\\end{proof}\n\n\\begin{theorem}\\label{cdf-gen iff gen by ground and disc}\nA multialgebra $\\mathcal{A}$ is $\\textbf{cdf}$-generated if, and only if, $\\mathcal{A}$ is generated by its ground and disconnected.\n\\end{theorem}\n\nWe have introduced several concepts that, although dissimilar in their definitions, are intrinsically connected; so it becomes important to analyze whether they are indeed distinct: are there multialgebras that are disconnected but not generated by their grounds? Are there multialgebras that are generated by their grounds but not disconnected? Or does being generated by its ground implies being disconnected, or vice-versa? We show below that this is not the case, for we provide examples answering positively both previous questions.\n\n\\begin{example}\\label{C}\nTake the signature $\\Sigma_{s}$ with a single unary operator, first defined in Example \\ref{s}. Consider the $\\Sigma_{s}-$multialgebra $\\mathcal{C}=(\\{-1,1\\}, \\{s_{\\mathcal{C}}\\})$ such that $s_{\\mathcal{C}}(-1)=\\{1\\}$ and $s_{\\mathcal{C}}(1)=\\{-1\\}$ (that is, $s_{\\mathcal{C}}(x)=\\{-x\\}$).\n\nWe state that $\\mathcal{C}$ is disconnected, but not generated by its ground. $\\mathcal{C}$ is clearly disconnected since $s_{\\mathcal{C}}(-1)\\cap s_{\\mathcal{C}}(1)=\\emptyset$; now, $B(\\mathcal{C})=s_{\\mathcal{C}}(-1)\\cup s_{\\mathcal{C}}(1)=\\{-1,1\\}$, and so $G(\\mathcal{C})=\\emptyset$. Since $\\Sigma_{0}=\\emptyset$, $\\bigcup_{\\sigma\\in\\Sigma_{0}}\\sigma_{\\mathcal{C}}=\\emptyset$ and therefore $\\langle G(\\mathcal{C})\\rangle_{n}=\\emptyset$ for every $n\\in\\mathbb{N}$, so that $G(\\mathcal{C})$ does not generate $\\mathcal{C}$.\n\\begin{figure}[H]\n\\centering\n\\begin{tikzcd}\n -1 \\arrow[rr, bend left=50, \"s_{\\mathcal{C}}\"] && 1 \\arrow[ll, bend left=50, \"s_{\\mathcal{C}}\"]\n \\end{tikzcd}\n\\caption*{The $\\Sigma_{s}$-multialgebra $\\mathcal{C}$}\n\\end{figure}\n\\end{example}\n\n\\begin{example}\\label{B}\nTake the signature $\\Sigma_{s}$ with a single unary operator. Consider the $\\Sigma_{s}-$multialge\\-bra $\\mathcal{B}=(\\{0, 1\\}, \\{s_{\\mathcal{B}}\\})$ such that $s_{\\mathcal{B}}(0)=\\{1\\}$ and $s_{\\mathcal{B}}(1)=\\{1\\}$ (that is, $s_{\\mathcal{B}}(x)=\\{1\\}$).\n\nThen $\\mathcal{B}$ is clearly not disconnected, since $s_{\\mathcal{B}}(0)\\cap s_{\\mathcal{B}}(1)=\\{1\\}$, yet $\\mathcal{B}$ is generated by its ground: $B(\\mathcal{B})=\\{1\\}$ and so $G(\\mathcal{B})=\\{0\\}$, and we see that $\\langle G(\\mathcal{B})\\rangle_{1}$ is already $\\{0,1\\}$.\n\\begin{figure}[H]\n\\centering\n\\begin{tikzcd}\n 0 \\arrow[r, \"s_{\\mathcal{B}}\"] & 1 \\arrow[loop right, out=30, in=-30, distance=3em]{}{s_{\\mathcal{B}}}\n \\end{tikzcd}\n\\caption*{The $\\Sigma_{s}$-multialgebra $\\mathcal{B}$}\n\\end{figure}\n\n\\end{example}\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\\subsection{Being disconnected and having a strong basis}\n\nNow, we define the notion of a strong basis (a minimum generating set), and prove that, on one hand, being generated by the ground implies having a strong basis, what means that being disconnected and generated by the ground implies being disconnected and having a strong basis; reciprocally, we also prove having a strong basis and being disconnected implies being disconnected and generated by the ground (although having a strong basis does not imply being generated by the ground). This provides a third characterization of our weakly free multialgebras.\n\nOur motivation, when coining the definition of a strong basis, was to be able to weaken that very condition: after all, absolutely free algebras (that is, algebras freely generated on the variety of all algebras on a given signature) are easier to define than the relatively free ones (which are the algebras freely generated on any variety one wants to consider), so it is natural that we start this study with ``absolutely free'' multialgebras. However, we still would like to be able, in the future, to define what should be a relatively free multialgebra (whatever a variety of multialgebras may be); to this end, weakening a strong basis to be a minimal (instead of minimum) generating set makes sense, given many relatively free algebras, on domains such as that of vector spaces, indeed have basis.\n\n\n\\begin{definition}\nWe say $B\\subseteq A$ is a strong basis\\index{Strong basis} of the $\\Sigma$-multialgebra $\\mathcal{A}=(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma})$ if it is the minimum of the set $\\mathcal{G}=\\{S\\subseteq A\\ : \\ \\langle S\\rangle=A\\}$ when ordered by inclusion.\n\\end{definition}\n\n\\begin{example}\nThe set of variables $\\mathcal{V}$ is a strong basis of $\\textbf{F}(\\Sigma, \\mathcal{V})$.\n\\end{example}\n\n\\begin{example}\nThe set of elements without predecessor of a directed forest of height $\\omega$ is a strong basis of the forest, when considered as a $\\Sigma_{s}$-multialgebra.\n\\end{example}\n\n\\begin{lemma}\\label{ground cap span is in set}\nFor every subset $S$ of the universe of a $\\Sigma$-multialgebra $\\mathcal{A}$, $G(\\mathcal{A})\\cap \\langle S\\rangle\\subseteq S$.\n\\end{lemma}\n\n\\begin{proof}\nSuppose $x\\in G(\\mathcal{A})\\cap \\langle S\\rangle$: if $x\\notin S$, we will show that $x$ cannot be in $\\langle S\\rangle$, which contradicts our assumption. Indeed, if $x\\notin S$ then \n\\[x\\notin\\langle S\\rangle_{0}=S\\cup \\bigcup_{\\sigma\\in\\Sigma_{0}}\\sigma_{\\mathcal{A}},\\]\nsince $x\\notin S$, and $x\\in G(\\mathcal{A})$ implies that \n\\[x\\in A\\setminus B(\\mathcal{A}) \\subseteq A\\setminus \\bigcup_{\\sigma\\in\\Sigma_{0}}\\sigma_{\\mathcal{A}}.\\]\n\nNow, for induction hypothesis, suppose that $x\\notin\\langle S\\rangle_{m}$; then\n\\[x\\notin \\langle S\\rangle_{m+1}=\\langle S\\rangle_{m}\\cup\\bigcup \\big\\{\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n}) \\ : \\ n\\in\\mathbb{N}, \\, \\sigma\\in\\Sigma_{n}, \\, a_{1}, \\dotsc , a_{n}\\in \\langle S\\rangle_{m} \\big\\}\\]\nsince $x\\notin \\langle S\\rangle_{m}$, and $x\\in G(\\mathcal{A})$ implies that \n\\[x\\in A\\setminus B(\\mathcal{A}) \\subseteq A\\setminus \\bigcup \\big\\{\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n}) \\ : \\ n\\in\\mathbb{N}, \\, \\sigma\\in\\Sigma_{n}, \\, a_{1}, \\dotsc , a_{n}\\in \\langle S\\rangle_{m} \\big\\}.\\]\n\\end{proof}\n\n\\begin{theorem}\\label{ground is in basis}\nIf the $\\Sigma$-multialgebra $\\mathcal{A}$ has a strong basis $B$, $G(\\mathcal{A})\\subseteq B$.\n\\end{theorem}\n\n\\begin{proof}\nBy lemma \\ref{ground cap span is in set}, $G(\\mathcal{A})=G(\\mathcal{A})\\cap A=G(\\mathcal{A})\\cap\\langle B\\rangle\\subseteq B$.\n\\end{proof}\n\nNotice lemma \\ref{ground cap span is in set} leads us too to the fact that, if $\\mathcal{A}$ is generated by its ground, then it has the ground as a strong basis: this is because, if $\\langle S\\rangle=A$, $G(\\mathcal{A})=G(\\mathcal{A})\\cap\\langle S\\rangle\\subseteq S$, and therefore $G(\\mathcal{A})$ becomes a minimum generating set.\n\n\\begin{definition}\nIf $B$ is a strong basis of a disconnected $\\Sigma$-multialgebra $\\mathcal{A}$, we define the $B$-order\\index{Order, $B$-}\\label{Border} of an element $a\\in A$ as the natural number \n\\[o_{B}(a)=\\min \\big\\{k\\in\\mathbb{N}\\ : \\ a\\in\\langle B\\rangle_{k}\\big\\}.\\]\n\\end{definition}\n\nThis is a clear generalization of the order, or complexity, of a formula: in fact, the order of a formula in ${T}(\\Sigma, \\mathcal{V})$ is exactly its $\\mathcal{V}$-order.\n\n\\begin{proposition}\nIf $a\\in \\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})$ and $o_{B}(a)\\geq 1$, then $o_{B}(a_{1}), \\dotsc , o_{B}(a_{n})0$, $o_{B}(a)>o_{B}(a_{1})\\geq 0$, which contradicts the fact that $a\\in B$ and therefore $o_{B}(a)=0$.\n\nIf $n=0$, it is clear that $B^{*}=B\\setminus \\{a\\}$ is a generating set smaller than $B$: generating set because, if $a\\in \\sigma_{\\mathcal{A}}$, $a\\in \\bigcup_{\\sigma\\in\\Sigma_{0}}\\sigma_{\\mathcal{A}}$ and therefore $B\\subseteq \\langle B^{*}\\rangle_{0}$, so that $\\langle B\\rangle_{m}\\subseteq\\langle B^{*}\\rangle_{m+1}$ and $\\bigcup_{m\\in\\mathbb{N}}\\langle B\\rangle_{m}=\\bigcup_{m\\in\\mathbb{N}}\\langle B^{*}\\rangle_{m}$. This is also a contradiction, since $B$ is a strong basis.\n\\end{proof}\n\n\\begin{theorem}\\label{theorem 3}\n$\\mathcal{A}$ is generated by its ground and disconnected if, and only if, it has a strong basis and it is disconnected.\n\\end{theorem}\n\n\\begin{proof}\nWe already proved, in Lemma \\ref{ground is basis if disc}, that if $\\mathcal{A}$ is disconnected and has a strong basis $B$, then it is generated by its ground and disconnected. Reciprocally, if $\\mathcal{A}$ is disconnected and generated by its ground, first of all it is clearly disconnected.\n\nNow, if $\\langle G(\\mathcal{A})\\rangle=A$ one has that $G(\\mathcal{A})\\subseteq S$ for every $S\\in \\{S\\subseteq A\\ : \\ \\langle S\\rangle=A\\}$, by Lemma \\ref{ground cap span is in set}. Therefore, the ground is a strong basis.\n\\end{proof}\n\nOnce again, we ask ourselves whether the concepts we have defined in this section are truly independent: does being disconnected imply having a strong basis? Does having a strong basis imply being disconnected? We show that neither is the case by providing examples of a multialgebra that is disconnected but does not have a strong basis and one of a multialgebra that has a strong basis but is not disconnected.\n\n\\begin{example}\nTake the $\\Sigma_{s}-$multialgebra $\\mathcal{C}$ from Example \\ref{C}.\n\nWe know that $\\mathcal{C}$ is disconnected, but we also state that it does not have a strong basis: in fact, we see that the set $\\{S\\subseteq \\{-1,1\\}\\ :\\ \\langle S\\rangle=\\{-1,1\\}\\}$ is exactly $\\{\\{-1\\}, \\{1\\}, \\{-1,1\\}\\}$, and this set has no minimum.\n\\end{example}\n\n\\begin{example}\nTake the $\\Sigma_{s}-$multialgebra $\\mathcal{B}$ from Example \\ref{B}.\n\nAs we saw before, $\\mathcal{B}$ is not disconnected, but we state that it has a strong basis: $B=\\{0\\}$ generates $\\mathcal{B}$ and, since $\\{1\\}$ does not generate the multialgebra, we find that $B$ is a minimum generating set.\n\\end{example}\n\nIn these two examples we presented a multialgebra ($\\mathcal{B}$) which has a strong basis and is generated by its ground, and one multialgebra ($\\mathcal{C}$) which does not have a strong basis and is not generated by its ground. Clearly being generated by its ground implies having a strong basis, so it is natural to hypothesize that having a strong basis and being generated by its ground could be equivalent notions; however, as we show in the example below, having a basis does not imply being generated by its ground.\n\n\\begin{example}\nTake the signature $\\Sigma_{s}$ with a single unary operator from Example \\ref{s}. Consider the $\\Sigma_{s}-$multialgebra $\\mathcal{M}=(\\{-1, 0, 1\\}, \\{s_{\\mathcal{M}}\\})$ such that $s_{\\mathcal{M}}(0)=\\{0\\}$, $s_{\\mathcal{M}}(1)=\\{1\\}$ and $s_{\\mathcal{M}}(-1)=\\{1\\}$ (that is, $s_{\\mathcal{M}}(x)=\\{|x|\\}$, where $|X|$ denotes the absolute value of $x$).\n\nWe have that $G(\\mathcal{M})=\\{-1\\}$ and that $\\langle\\{-1\\}\\rangle=\\{-1, 1\\}$, so that $\\mathcal{M}$ is not generated by its ground. But we state that $\\{-1, 0\\}$ is a strong basis: first of all, it clearly generates $\\mathcal{M}$; furthermore, the generating sets of $\\mathcal{M}$ are only $\\{-1, 0\\}$ and $\\{-1, 0, 1\\}$, so that $\\{-1, 0\\}$ is in fact the smallest generating set.\n\n\\begin{figure}[H]\n\\centering\n\\begin{tikzcd}\n -1 \\arrow[rr, bend right=50]{}{s_{\\mathcal{M}}} & 0 \\arrow[u, loop]{}{s_{\\mathcal{M}}} & 1 \\arrow[loop right, out=30, in=-30, distance=3em]{}{s_{\\mathcal{M}}}\n \\end{tikzcd}\n\\caption*{The $\\Sigma_{s}$-multialgebra $\\mathcal{M}$}\n\\end{figure}\n\\end{example}\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\\subsection{Being disconnected and chainless}\n\nThe last equivalence to being a submultialgebra of $\\textbf{mF}(\\Sigma, \\mathcal{V}, \\kappa)$ we give depends on the notion of being chainless, which is very graph-theoretical in nature. Think of a tree that ramifies ever downward: one can pick any vertex and proceed, against the arrows, upwards until an element without predecessor is reached. More than that, it is not possible to find an infinite path, starting in any one vertex, by always going against the arrows: such a path, if it existed, would be what we shall call a chain. A multialgebra without chains is, very naturally, chainless.\n\nAs it was in the case of strong bases, there isn't a parallel concept to being chainless among the theory of universal algebra: it seems that this concept is far more natural when dealing with multioperations, although it can be easily applied to algebras if one wishes to do so. A close, although not equivalent, entity are the branches in the formation trees of formulas: if allowed to grow infinitely, these would became chains.\n\nThe main result of this section is, perhaps, the fact that being chainless implies being generated by its ground, which, as we know, implies in turn having a strong basis. Of course, when adding disconnectedness to the equation, all three concepts become equivalent to one another and to the fact that the multialgebra at hand is weakly free. We have, then, the following schematic diagram, that offers an overview of the aforementioned results.\n\n\\begin{figure}[H]\n\\centering\n\\begin{minipage}[t]{3cm}\n\\centering\n\\begin{tikzcd}[arrows=Rightarrow, every arrow\/.append style={shift right=3ex}]\n\\text{$\\mathcal{A}$ is chainless} \\arrow[Rightarrow]{d}\\\\\n\\text{$\\mathcal{A}$ is generated by $G(\\mathcal{A})$} \\arrow[Rightarrow]{d}\\arrow[negated]{u}\\\\\n\\text{$\\mathcal{A}$ has a strong basis} \\arrow[negated]{u}\n \\end{tikzcd}\n\\end{minipage}\n\\hspace{2cm}\n\\centering\n\\begin{minipage}[t]{6cm}\n\\centering\n\\begin{tikzcd}[arrows=Rightarrow, every arrow\/.append style={shift right=3ex}]\n\\begin{tabular}{c}$\\mathcal{A}$ is chainless\\\\ and disconnected\\end{tabular} \\arrow[Rightarrow]{d}\\\\\n\\begin{tabular}{c}$\\mathcal{A}$ is generated by $G(\\mathcal{A})$\\\\ and disconnected\\end{tabular} \\arrow[Rightarrow]{d}\\arrow{u}\\\\\n\\begin{tabular}{c}$\\mathcal{A}$ has a strong basis\\\\ and is disconnected\\end{tabular} \\arrow{u}\n \\end{tikzcd}\n\\end{minipage}\n\\end{figure}\n\n\n\nGiven a function $\\tau:\\{1, \\dotsc , n\\}\\rightarrow\\{1, \\dotsc , n\\}$ in $S_{n}$, the group of permutations on $n$ elements (meaning $\\tau$ is bijective), the action of $\\tau$ in an $n$-tuple $(x_{1}, \\dotsc , x_{n})\\in X^{n}$ is given by\n\\[\\tau(x_{1}, \\dotsc , x_{n})=(x_{\\tau(1)}, \\dotsc , x_{\\tau(n)}).\\]\nGiven $1\\leq i, j\\leq n$, we define $[i,j]$ to be the permutation such that $[i,j](i)=j$, $[i,j](j)=i$ and, for $k\\in\\{1, \\dotsc , n\\}$ different from $i$ and $j$, $[i,j](k)=k$.\n\n\\begin{definition}\nGiven a $\\Sigma$-multialgebra $\\mathcal{A}$, a sequence $\\{a_{n}\\}_{n\\in\\mathbb{N}}\\subseteq A$ is said to be a chain\\index{Chain} if, for every $n\\in\\mathbb{N}$, there exist a natural number $m_{n}\\in\\mathbb{N}$, a functional symbol $\\sigma^{n}\\in\\Sigma_{m_{n}}$, a permutation $\\tau_{n}\\in S_{m_{n}}$ and elements $a_{1}^{n}, \\dotsc , a_{m_{n}-1}^{n}\\in A$ such that \n\\[a_{n}\\in \\sigma^{n}_{\\mathcal{A}}(\\tau_{n}(a_{n+1}, a_{1}^{n}, \\dotsc , a_{m_{n}-1}^{n})).\\]\n\nA $\\Sigma$-multialgebra is said to be chainless\\index{Chainless} when it has no chains.\n\\end{definition}\n\n\\begin{example}\nTake a directed forest of height $\\omega$ and add to it a loop, that is, choose a vertex $v$ and add an arrow from $v$ to $v$: then $\\{a_{n}\\}_{n\\in\\mathbb{N}}$, such that $a_{n}=v$ for every $n\\in\\mathbb{N}$, is a chain.\n\\end{example}\n\n\\begin{example}\n$\\textbf{F}(\\Sigma, \\mathcal{V})$ is chainless.\n\\end{example}\n\n\\begin{lemma}\\label{chainless implies gen by ground}\nIf $\\mathcal{A}$ is chainless, then it is generated by its ground.\n\\end{lemma}\n\n\\begin{proof}\nSuppose that this not hold, so $A\\setminus\\langle G(\\mathcal{A})\\rangle$ is not empty, and must therefore contain some element $a_{0}$. We create a chain $\\{a_{n}\\}_{n\\in\\mathbb{N}}$ by induction, being the case $n=0$ already done.\n\nSo, suppose we have created a finite sequence of elements $a_{0}, \\dotsc , a_{k}\\in A\\setminus \\langle G(\\mathcal{A})\\rangle$ such that, for each $0\\leq n< k$, there exist a positive integer $m_{n}\\in\\mathbb{N}\\setminus\\{0\\}$, a functional symbol $\\sigma^{n}\\in\\Sigma_{m_{n}}$, a permutation $\\tau_{n}\\in S_{m_{n}}$ and elements $a_{1}^{n}, \\dotsc , a_{m_{n}-1}^{n}\\in A$ such that \n\\[a_{n}\\in \\sigma^{n}_{\\mathcal{A}}(\\tau_{n}(a_{n+1}, a_{1}^{n}, \\dotsc , a_{m_{n}-1}^{n})).\\]\n\nSince $a_{k}\\in A\\setminus \\langle G(\\mathcal{A})\\rangle$, we have that $a_{k}$ is not an element of the ground; so, there must exist $m_{k}\\in\\mathbb{N}$, a functional symbol $\\sigma^{k}\\in\\Sigma_{m_{k}}$ and elements $b_{1}^{k}, \\dotsc , b_{m_{k}}^{k}\\in A$ such that\n\\[a_{k}\\in\\sigma^{k}_{\\mathcal{A}}(b_{1}^{k}, \\dotsc , b_{m_{k}}^{k}).\\]\nNow, if all $b_{1}^{k}, \\dotsc , b_{m_{k}}^{k}$ belonged to $\\langle G(\\mathcal{A})\\rangle$, so would $a_{k}$: there must be an element $a_{k+1}\\in\\{b_{1}^{k}, \\dotsc , b_{m_{k}}^{k}\\}$, say $b_{l}^{k}$, such that $a_{k+1}\\in A\\setminus \\langle G(\\mathcal{A})\\rangle$. We then define $a_{i}^{k}$ as $b_{j}^{k}$, for $i\\in\\{1, \\dotsc , m_{k}-1\\}$ and $j=\\min\\{i\\leq p\\leq m_{k}\\ : \\ p\\neq l\\}$, and \n\\[\\tau_{k}=[l-1, l]\\circ\\cdots\\circ[1,2],\\]\nand then it is clear that $\\{a_{n}\\}_{n\\in\\mathbb{N}}$ becomes a chain, with the extra condition that $\\{a_{n}\\}_{n\\in\\mathbb{N}}\\subseteq A\\setminus \\langle G(\\mathcal{A})\\rangle$. Since $\\mathcal{A}$ is chainless, we reach a contradiction, so we must have instead that $A\\setminus \\langle G(\\mathcal{A})\\rangle=\\emptyset$, and therefore $\\mathcal{A}$ is generated by its ground.\n\\end{proof}\n\nSo, a disconnected, chainless multialgebra is, by Lemma \\ref{chainless implies gen by ground}, disconnected and generated by its ground. We state, that, in fact, the reciprocal also holds, when we arrive at yet another characterization of being an weakly free multialgebra.\n\nSo, suppose $\\mathcal{A}$ is disconnected and generated by its ground, and let $\\{a_{n}\\}_{n\\in\\mathbb{N}}$ be a chain in $\\mathcal{A}$: clearly no $a_{n}$ can belong to the ground, since \n\\[a_{n}\\in \\sigma^{n}_{\\mathcal{A}}(\\tau_{n}(a_{n+1}, a_{1}^{n}, \\dotsc , a_{m_{n}-1}^{n})),\\]\nand therefore $o_{G(\\mathcal{A})}(a_{n+1})a$.\n\n\\begin{definition}\nGiven a poset $(A, \\leq)$, an element $a\\in A$ is:\n\\begin{enumerate}\n\\item a maximum if, for all $b\\in A$, $b\\leq a$;\n\\item a minimum if, for all $b\\in A$, $a\\leq b$;\n\\item maximal if, for all $b\\in A$, $a\\leq b$ implies $a=b$;\n\\item minimal if, for all $b\\in A$, $b\\leq a$ implies $a=b$.\n\\end{enumerate}\n\nFurthermore, given a subset $S\\subseteq A$, we say $a\\in A$ is:\n\\begin{enumerate}\n\\item an upper bound for $S$ if, for any $s\\in S$, $s\\leq a$;\n\\item a lower bound for $S$ if, for any $s\\in S$, $a\\leq s$;\n\\item the supremum of $S$, when we write $a=\\sup S$, if it is the minimum of all upper bounds for $S$, meaning that:\n\\begin{enumerate}\n\\item for any $s\\in S$, $s\\leq a$;\n\\item if $b\\in A$ is such that, for any $s\\in S$, $s\\leq b$, then $a\\leq b$;\\footnote{Notice that there is indeed only one supremum of a set, as well as only one infimum: if $a$ and $b$ are both minimum upper bounds for $S$, the fact that $a$ is a minimum upper bound and that $b$ is an upper bound gives us $a\\leq b$; reciprocally, the fact $b$ is a minimum upper bound and $a$ is an upper bound implies $b\\leq a$, and so $a=b$. The same reasoning applies to infima.}\n\\end{enumerate}\n\\item the infimum of $S$, when we write $a=\\inf S$, if it is the maximum of all lower bounds for $S$, meaning that:\n\\begin{enumerate}\n\\item for any $s\\in S$, $a\\leq s$;\n\\item if $b\\in A$ is such that, for any $s\\in S$, $a\\leq s$, then $b\\leq a$.\n\\end{enumerate}\n\\end{enumerate}\n\\end{definition}\n\nOf course, we may define maxima, minima, maximal and minimal elements for a subset $S\\subseteq A$ of $A$, by merely restricting the order of $(A, \\leq)$ to $S$.\n\nA Boolean algebra, to which we have already given on Section \\ref{Lattices, and Boolean... } a purely algebraic formulation, is a partially-ordered sets $(A, \\leq)$ such that: there are a maximum (denoted by $1$) and a minimum ($0$) elements, which we shall assume distinct; for every pair of elements $a, b\\in A$, the set $\\{a, b\\}$ has a supremum, denoted by $a\\vee b$, and an infimum, denoted by $a\\wedge b$; and every element $a$ has a complement $b$, which satisfies\n\\[b=\\inf\\{c\\in A: \\sup\\{a, c\\}=1\\}\\]\nand\n\\[b=\\sup\\{c\\in A: \\inf\\{a, c\\}=0\\}.\\]\n\nA poset $(A, \\leq)$ is said to be complete\\index{Poset, complete} if every $S\\subseteq A$ has a supremum and an infimum. \n\n\\begin{lemma}\n\\begin{enumerate} \n\\item Every Boolean algebra $(A, \\leq)$ is distributive, meaning \n\\[a\\vee (b\\wedge c)=(a\\vee b)\\wedge(a\\vee c)\\quad\\text{and}\\quad a\\wedge(b\\vee c)=(a\\wedge b)\\vee(a\\wedge c),\\]\nfor any $a, b, c\\in A$;\n\n\\item every complete Boolean algebra $(A, \\leq)$ is infinite distributive, meaning that for any $S\\cup\\{a\\}\\subseteq A$, \n\\[\\sup\\{\\inf\\{a, s\\}: s\\in S\\}=\\inf\\{a, \\sup S\\}\\quad\\text{and}\\quad\\inf\\{\\sup\\{a, s\\}: s\\in S\\}=\\sup\\{a, \\inf S\\}.\\]\n\\end{enumerate}\n\\end{lemma}\n\n\\begin{proof}\nWe only prove the first equation related to distributivity, given that the proof for the other is very similar. Denote $p=a\\vee (b\\wedge c)$, meaning \n\\[p=\\sup\\{a, \\inf\\{b, c\\}\\},\\]\nand $q=(a\\vee b)\\wedge (a\\vee c)$, that is \n\\[q=\\inf\\{\\sup\\{a, b\\}, \\sup\\{a, c\\}\\};\\]\nsince $\\inf\\{b, c\\}\\leq b$ and $\\inf\\{b, c\\}\\leq c$, $p=\\sup\\{a, \\inf\\{b, c\\}\\}\\leq \\sup\\{a, b\\}$ and, analogously, $p\\leq \\sup\\{a, c\\}$, implying $p$ is a lower bound for $\\{\\sup\\{a, b\\}, \\sup\\{a, c\\}\\}$. Since $q$ is the largest of these lower bounds, we get $p\\leq q$.\n\nReciprocally, $p=\\sup\\{a, \\inf\\{b, c\\}\\}$ implies $a\\leq p$ and $\\inf\\{b, c\\}\\leq p$: if $\\inf\\{b, c\\}=p$, this means $a\\leq \\inf\\{b, c\\}$ and therefore $a\\leq b$ and $a\\leq c$, meaning that $\\sup\\{a, b\\}=b$ and $\\sup\\{a, c\\}=c$ and therefore $q=\\inf\\{b, c\\}=p$; if, otherwise, $\\inf\\{b, c\\}2$, with axiom schemata\n\\begin{enumerate}\n\\item[\\textbf{Ax\\: 1}] $\\alpha\\rightarrow(\\beta\\rightarrow\\alpha)$;\n\\item[\\textbf{Ax\\: 2}] $\\big(\\alpha\\rightarrow (\\beta\\rightarrow \\gamma)\\big)\\rightarrow\\big((\\alpha\\rightarrow\\beta)\\rightarrow(\\alpha\\rightarrow\\gamma)\\big)$;\n\\item[\\textbf{Ax\\: 3}] $\\alpha\\rightarrow\\big(\\beta\\rightarrow(\\alpha\\wedge\\beta)\\big)$;\n\\item[\\textbf{Ax\\: 4}] $(\\alpha\\wedge\\beta)\\rightarrow \\alpha$;\n\\item[\\textbf{Ax\\: 5}] $(\\alpha\\wedge\\beta)\\rightarrow \\beta$;\n\\item[\\textbf{Ax\\: 6}] $\\alpha\\rightarrow(\\alpha\\vee\\beta)$;\n\\item[\\textbf{Ax\\: 7}] $\\beta\\rightarrow(\\alpha\\vee\\beta)$;\n\\item[\\textbf{Ax\\: 8}] $(\\alpha\\rightarrow\\gamma)\\rightarrow\\Big((\\beta\\rightarrow\\gamma)\\rightarrow \\big((\\alpha\\vee\\beta)\\rightarrow\\gamma\\big)\\Big)$;\n\\item[\\textbf{Ax\\: 9}] $(\\alpha\\rightarrow\\beta)\\rightarrow\\big((\\alpha\\rightarrow{\\sim} \\beta)\\rightarrow{\\sim} \\alpha\\big)$;\n\\item[\\textbf{Ax\\: 10}] $\\alpha\\rightarrow({\\sim} \\alpha\\rightarrow\\beta)$;\n\\item[\\textbf{Ax\\: 11}] $\\alpha\\vee{\\sim} \\alpha$;\n\\item[\\textbf{Ax\\: 12}] $\\bot\\rightarrow\\alpha$;\n\\item[\\textbf{Ax\\: 13}] $\\alpha\\rightarrow\\top$;\n\\end{enumerate}\nand following the inference rule of Modus Ponens\n\\[\\frac{\\alpha\\quad\\alpha\\rightarrow\\beta}{\\beta}.\\]\n\n\nAs opposed to $\\textbf{CPL}$, a tarskian logic $\\mathscr{L}$ will be said to be paraconsistent\\index{Logic, Paraconsistent} when it possesses an unary symbol \"$\\neg $\", that we shall refer to as a negation, such that there exist formulas $\\alpha$ and $\\beta$ of $\\mathscr{L}$ satisfying $\\alpha, \\neg\\alpha\\not\\vdash_{\\mathscr{L}}\\beta$. A good, intuitive way to look at paraconsistent logics is through the paradigm introduced by da Costa that the formulas of such a logic can, at times, be divided into well-behaved and badly-behaved (nowadays more commonly referred to as inconsistent) ones. Think of a scientist performing an experiment: from her school years, said scientist knows that opposite poles of a magnet attract each other; so, if during the experiment two poles she thought to be opposite are instead repealing each other, she can be certain that she was wrong, and the two poles are instead the same, both north or both south. Her reasoning works because the sentence ``opposite poles attract'' is well-behaved (that is, it can not coexist with its negation) and true, while what she had apparently observed was that ``opposite poles repeal'', a well-behaved although false sentence.\n\nNow, suppose that our scientist has discovered that all monopoles (conjectured magnetic particles that, instead of having both north and south poles, have only one pole, hence the name) have exactly the same pole and is now testing two hypotheses: ``all monopoles are north poles'' and ``all monopoles are south poles''. Again, we have contradictory sentences, as was the case with ``opposite poles attract'' and ``opposite poles repeal'', but this time our scientist can not reach the conclusion that, at some prior step, she made a mistake: why is that?\\footnote{Notice that we are presenting this logical problem in a somewhat convoluted way in order to avoid explicit explosivity of the theory at hand: scientists, and most mathematicians, appear to prefer thinking about these topics in terms of non-contradiction (what is eventually equivalent).} Simply put: none of the hypotheses to be tested, ``all monopoles are north poles'' and ``all monopoles are south poles'', is well-behaved, meaning that we can not discard their negations as necessarily false.\n\nWhat we are doing is dividing those sentences found in science between true-or-false sentences (such as those involving attracting or repealing poles) and hypothetical sentences (such as those involving all monopoles). But this is not exclusive to scientists: mathematicians also have true-or-false sentences and conjectural sentences: after all, some mathematicians believe that, e.g., the Riemann hypothesis is true, while others believe it to be false, and that doesn't make mathematics as a whole trivial. In daily life, conjectural and hypothetical sentences may be replaced with rumors, that can also be contradictory without making logical reasoning unfeasible. da Costa's hierarchy, which deals with these distinctions plus higher degrees of consistency, will be studied algebraically in Chapters \\ref{Chapter5} and \\ref{Chapter6}. \n\nIf our logic $\\mathscr{L}$ also possesses a binary symbol \"$\\rightarrow$\" satisfying the deduction meta-theorem, i.e., $\\Gamma, \\psi\\vdash_{\\mathscr{L}}\\varphi$ if and only if $\\Gamma\\vdash_{\\mathscr{L}}\\psi\\rightarrow\\varphi$, then the condition that there exist formulas $\\alpha$ and $\\beta$ such that $\\alpha, \\neg\\alpha\\not\\vdash_{\\mathscr{L}}\\beta$ is equivalent to the existence of formulas $\\alpha$ and $\\beta$ such that $\\not\\vdash_{\\mathscr{L}}\\alpha\\rightarrow(\\neg\\alpha\\rightarrow\\beta)$, which is know as the failure of the explosion law\\index{Explosion law}; if this result were to be true for any $\\alpha$ and $\\beta$, that would mean the explosion law would be valid, which indeed happens in $\\textbf{CPL}$ as one can see by its $\\textbf{Ax\\: 10}$.\n\nIf a formula $\\psi$ has all its propositional variables among the set $\\{p_{1}, \\dotsc , p_{n}\\}$, we may write $\\psi(p_{1}, \\dotsc , p_{n})$\\label{varformula}: to define logics of formal inconsistency, we shall need a set of formulas $\\bigcirc(p)$, each of them dependent exactly on the propositional variable $p$, that is, each one of them contains $p$ and solely $p$ as a variable.\n\nAnd if we apply to a formula $\\psi(p_{1}, \\dotsc , p_{n})$ a homomorphism $\\sigma$ sending $p_{i}$ to the formula $\\alpha_{i}$, for $i\\in\\{1, \\dotsc , n\\}$, we shall write $\\psi(\\alpha_{1}, \\dotsc , \\alpha_{n})$ for $\\sigma(\\psi)$: this way, $\\bigcirc(\\alpha)$ will be the set of formulas obtained from $\\bigcirc(p)$ after we apply to each of its elements the homomorphism taking $p$ to $\\alpha$, that is,\n\\[\\bigcirc(\\alpha)=\\{\\psi(\\alpha)\\ :\\ \\psi(p)\\in\\bigcirc(p)\\}.\\]\n\n\\begin{definition}\\label{Definition of LFI}\nA tarskian, finitary and structural logic $\\mathscr{L}$ containing a negation and a set of formulas $\\bigcirc(p)\\neq\\emptyset$ depending exactly on the propositional variable $p$ is a logic of formal inconsistency ($\\textbf{LFI}$)\\index{Logic of formal inconsistency}\\label{LFI} when:\n\\begin{enumerate}\n\\item there exist formulas $\\phi$ and $\\theta$ in $\\mathscr{L}$ such that $\\phi, \\neg\\phi\\not\\vdash_{\\mathscr{L}}\\theta$;\n\\item there exist formulas $\\alpha$ and $\\beta$ in $\\mathscr{L}$ such that\n\\begin{enumerate}\n\\item $\\bigcirc(\\alpha), \\alpha\\not\\vdash_{\\mathscr{L}}\\beta$ and\n\\item $\\bigcirc(\\alpha), \\neg\\alpha\\not\\vdash_{\\mathscr{L}}\\beta$;\n\\end{enumerate}\n\\item for all formulas $\\varphi$ and $\\psi$ in $\\mathscr{L}$,\n\\[\\bigcirc(\\varphi), \\varphi, \\neg\\varphi\\vdash_{\\mathscr{L}}\\psi.\\]\n\\end{enumerate}\n\\end{definition}\n\nThe logic $\\mathscr{L}$ will be called a weak $\\textbf{LFI}$\\index{Weak $\\textbf{LFI}$} if the second condition of Definition \\ref{Definition of LFI} is replaced by there existing formulas $\\alpha_{1}$ and $\\beta_{1}$ such that\n\\[\\bigcirc(\\alpha_{1}), \\alpha_{1}\\not\\vdash_{\\mathscr{L}}\\beta_{1}\\]\nand $\\alpha_{2}$ and $\\beta_{2}$, possibly different from respectively $\\alpha_{1}$ and $\\beta_{1}$, such that\n\\[\\bigcirc(\\alpha_{2}), \\neg\\alpha_{2}\\not\\vdash_{\\mathscr{L}}\\beta_{2}.\\]\n\nThe logic $\\mathscr{L}$ will be called a strong $\\textbf{LFI}$\\index{Strong $\\textbf{LFI}$} if the first and second conditions of Definition \\ref{Definition of LFI} are replaced by there existing a single pair of formulas $\\alpha$ and $\\beta$ satisfying simultaneously\n\\begin{enumerate}\n\\item $\\alpha, \\neg\\alpha\\not\\vdash_{\\mathscr{L}}\\beta$,\n\\item $\\bigcirc(\\alpha), \\alpha\\not\\vdash_{\\mathscr{L}}\\beta$ and\n\\item $\\bigcirc(\\alpha), \\neg\\alpha\\not\\vdash_{\\mathscr{L}}\\beta$.\n\\end{enumerate}\nAll these definitions are the Definitions $2.1.7$, $2.1.8$ and $2.1.9$ of \\cite{ParLog}.\n\nWhen the set $\\bigcirc(p)$ contains a single formula, this formula will be denoted by $\\circ p$\\label{circ}, being the consistency of $p$. More often than not, we will want \"$\\circ$\" to be a primitive connective, as well as consistency a primitive notion.\n\nThe $\\textbf{LFI}$ we will most frequently work with is one that, beyond having a primitive consistency, emulates only the positive fragment of classical propositional logic and excluded middle: this way, $\\textbf{mbC}$\\label{mbC} is often regarded as the simplest logic of formal inconsistency. It has the signature we will denote by $\\Sigma_{\\textbf{LFI}}$\\label{SigmaLFI}, with $(\\Sigma_{\\textbf{LFI}})_{0}=\\emptyset$, $(\\Sigma_{\\textbf{LFI}})_{1}=\\{\\neg , \\circ\\}$, $(\\Sigma_{\\textbf{LFI}})_{2}=\\{\\vee, \\wedge, \\rightarrow\\}$ and $(\\Sigma_{\\textbf{LFI}})_{n}=\\emptyset$ for $n>2$; it has as axiom schemata\n\\begin{enumerate}\n\\item[\\textbf{Ax\\: 1}] $\\alpha\\rightarrow(\\beta\\rightarrow\\alpha)$;\n\\item[\\textbf{Ax\\: 2}] $\\big(\\alpha\\rightarrow (\\beta\\rightarrow \\gamma)\\big)\\rightarrow\\big((\\alpha\\rightarrow\\beta)\\rightarrow(\\alpha\\rightarrow\\gamma)\\big)$;\n\\item[\\textbf{Ax\\: 3}] $\\alpha\\rightarrow\\big(\\beta\\rightarrow(\\alpha\\wedge\\beta)\\big)$;\n\\item[\\textbf{Ax\\: 4}] $(\\alpha\\wedge\\beta)\\rightarrow \\alpha$;\n\\item[\\textbf{Ax\\: 5}] $(\\alpha\\wedge\\beta)\\rightarrow \\beta$;\n\\item[\\textbf{Ax\\: 6}] $\\alpha\\rightarrow(\\alpha\\vee\\beta)$;\n\\item[\\textbf{Ax\\: 7}] $\\beta\\rightarrow(\\alpha\\vee\\beta)$;\n\\item[\\textbf{Ax\\: 8}] $(\\alpha\\rightarrow\\gamma)\\rightarrow\\Big((\\beta\\rightarrow\\gamma)\\rightarrow \\big((\\alpha\\vee\\beta)\\rightarrow\\gamma\\big)\\Big)$;\n\\item[$\\textbf{Ax\\: 9}^{*}$] $(\\alpha\\rightarrow \\beta)\\vee\\alpha$;\n\\item[$\\textbf{Ax\\: 11}^{*}$] $\\alpha\\vee\\neg \\alpha$,\n\\end{enumerate}\nplus \n\\[\\tag{\\textbf{bc1}}\\circ\\alpha\\rightarrow(\\alpha\\rightarrow(\\neg \\alpha\\rightarrow \\beta)),\\]\nand as inference rules that of Modus Ponens,\n\\[\\frac{\\alpha\\quad\\alpha\\rightarrow\\beta}{\\beta}.\\]\n\nOther logics of formal incompatibility we will make use of are:\n\\begin{enumerate}\n\\item $\\textbf{mbCciw}$\\label{mbCciw}, obtained from the Hilbert system for $\\textbf{mbC}$ by adding the axiom schema\\label{ciw}\n\\[\\tag{\\textbf{ciw}} \\circ\\alpha\\vee(\\alpha\\wedge\\neg \\alpha);\\]\n\\item $\\textbf{mbCci}$\\label{mbCci}, obtained from $\\textbf{mbC}$ by adding\\label{ci}\n\\[\\tag{\\textbf{ci}}\\neg \\circ\\alpha\\rightarrow(\\alpha\\wedge\\neg \\alpha);\\]\n\\item $\\textbf{mbCcl}$\\label{mbCcl}, obtained from $\\textbf{mbC}$ by adding\\label{cl}\n\\[\\tag{\\textbf{cl}}\\neg(\\alpha\\wedge\\neg \\alpha)\\rightarrow\\circ\\alpha.\\]\n\\end{enumerate}\n\n\n\n\n\\section{Matrices and generalizations}\n\nA logical matrix\\index{Matrix, Logical} over a signature $\\Sigma$ is a pair $\\mathcal{M}=(\\mathcal{A}, D)$ such that $\\mathcal{A}=(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma})$ is a $\\Sigma$-algebra and $D\\subseteq A$ is said to be the set of designated elements\\index{Designated elements} of the matrix. Given formulas $\\Gamma\\cup\\{\\varphi\\}$ over the signature $\\Sigma$, we say $\\Gamma$ semantically proves $\\varphi$ according to $\\mathcal{M}$ (and write $\\Gamma\\vDash_{\\mathcal{M}}\\varphi$\\label{semant.proves}) if, for every homomorphism $\\nu:\\textbf{F}(\\Sigma, \\mathcal{V})\\rightarrow\\mathcal{A}$ such that $\\nu(\\gamma)\\in D$, for every $\\gamma\\in\\Gamma$, one has that $\\nu(\\varphi)\\in D$.\n\nA logic $\\mathcal{L}$, over the signature $\\Sigma$, is said to be characterized\\index{Characterized} by $\\mathcal{M}$ if, for every set of formulas $\\Gamma\\cup\\{\\varphi\\}$ over $\\Sigma$, $\\Gamma\\vdash_{\\mathcal{L}}\\varphi$ if and only if $\\Gamma\\vDash_{\\mathcal{M}}\\varphi$. \n\nQuite analogously, given a class $\\mathbb{M}$ of matrices $\\mathcal{M}$ over the same signature, we write $\\Gamma\\vDash_{\\mathbb{M}}\\varphi$\\label{vDashM} if $\\Gamma\\vDash_{\\mathcal{M}}\\varphi$ for every $\\mathcal{M}\\in\\mathbb{M}$, and a logic $\\mathcal{L}$ is said to be characterized by $\\mathbb{M}$ when $\\Gamma\\vdash_{\\mathcal{L}}\\varphi$ if and only if $\\Gamma\\vDash_{\\mathbb{M}}\\varphi$.\n\nIt is a well know result by W\\'ojcicki\\index{W\\'ojciki} that every tarskian logic can be characterized by a suitable class of logical matrices, see \\cite{Woj}. And, although one could see such a result as settling the matter of logical matrices, many times the suitable class of logical matrices obtained for a logic is not efficient, as in, to name one example, its algebras can be too large or complex, or the class may be infinite. So alternatives have been offered to classes of logical matrices, as in W\\'ojcicki \\cite{Woj2}.\n\nAnother approach is the one we will call that of restricted matrices\\index{Matrix, Restricted}, or Rmatrices\\index{Rmatrix}, (although that was not its original nomenclature, Piochi\\index{Piochi} named then $\\mathcal{E}$-matrices instead), see \\cite{Piochi} or, for an approach focusing more on structurality of the related closure operators, \\cite{Piochi2} and \\cite{Piochi3}. A restricted matrix over a signature $\\Sigma$ is a triple $\\mathcal{M}=(\\mathcal{A}, D, \\mathcal{F})$ such that:\n\\begin{enumerate}\n\\item $\\mathcal{A}=(A,\\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma})$ is a $\\Sigma$-algebra;\n\\item $D$ is a subset of $A$;\n\\item $\\mathcal{F}$ is a set of homomorphisms from $\\textbf{F}(\\Sigma, \\mathcal{V})$ to $\\mathcal{A}$.\n\\end{enumerate}\n\nThe set $\\mathcal{F}$ will be called the set of restrictions\\index{Restrictions}. Given formulas $\\Gamma\\cup\\{\\varphi\\}$ over the signature $\\Sigma$, we say $\\Gamma$ proves $\\varphi$ according to a restricted matrix $\\mathcal{M}$ (also over $\\Sigma$) and write $\\Gamma\\vDash_{\\mathcal{M}}\\varphi$ if, for every homomorphism $\\nu\\in\\mathcal{F}$, $\\nu(\\gamma)\\in D$, for every $\\gamma\\in\\Gamma$, implies $\\nu(\\varphi)\\in D$; we say a logic $\\mathcal{L}$ is characterized by a restricted matrix $\\mathcal{M}$ when $\\Gamma\\vdash_{\\mathcal{L}}\\varphi$ if and only if $\\Gamma\\vDash_{\\mathcal{M}}\\varphi$, and it is possible to prove that every tarskian logic can be characterized by a, potentially infinite, restricted matrix (see Piochi's \\cite{Piochi} and \\cite{Piochi2}).\n\nAnd yet, given sometimes this is the most efficient approach, we may define $\\vDash_{\\mathbb{M}}$ for $\\mathbb{M}$ a class of restricted matrices in much the same way we did for a class of matrices. \n\nMany of these generalizations have one simple objective, which is to offer a reasonable decision method for a logic; of course, methods that depend on infinite matrices or infinite classes of matrices are not really decision methods, and so finite matrices, or finite sets of finite matrices, are the preferable outcomes of the algebraization process of a logic. Of course, such outcomes are not always possible, as many results on uncharacterizability of logics show: probably the earliest one is G{\\\"o}del's\\index{G{\\\"o}del} proof that propositional, intuitionistic logic is not characterizable by a single, finite logical matrix, found in \\cite{Godel}.\n\nInspired by G{\\\"o}del's proof, Dugundji (\\cite{Dugundji})\\index{Dugundji} proved that no system lying between the modal logics $\\textbf{S1}$ and $\\textbf{S5}$ admits a single, finite logic matrix which characterizes it as well, and in many places we may refer to results regarding uncharacterizability of logics as ``Dugundji-like'' theorems. As we enter the domain of paraconsistent logics, these results abound, given the intrinsic complexity of many of those systems: one example would be Avron's\\index{Avron} proof that many logics, including da Costa's $C_{1}$, do not possess a characterizing finite Nmatrix, or even a characterizing finite set of finite Nmatrices (\\cite{Avron, Avron3}).\n\nMost of the work shown in Sections \\ref{Restricted Nmatrices}, \\ref{A brief history of RNmatrices}, \\ref{Structurality} and \\ref{Examples: characterizing some logics with RNmatrices} was submitted to an online repository in \\cite{CostaRNmatrix}, and then finally published in \\cite{TwoDecisionProcedures}.\n\n\n\n\n\\subsection{Restricted Nmatrices}\\label{Restricted Nmatrices}\n\nAs we explained before, although the problem of finding semantics for a given logic is somewhat solved in the case the logic at hand is tarskian, the corresponding class of matrices or even restricted matrix associated to the logic may not be sufficiently efficient, and in the realm of paraconsistency and, furthermore, in the presence of incompatibility, which we will further study ahead, non-deterministic matrices have been proven fruitful. The first approach to non-deterministic matrices is found in the work of Rescher\\index{Rescher} (see \\cite{Rescher}) and Ivlev\\index{Ivlev} (see \\cite{Lev}, \\cite{Lev2}, \\cite{Lev3} and \\cite{Lev4}), although our reasoning will be closer to that of \\cite{Avron}.\n\n\\begin{definition}\nGiven a signature $\\Sigma$, a non-deterministic matrix over $\\Sigma$, or Nmatrix\\index{Nmatrix}\\index{Matrix, Non-deterministic}, is a pair $\\mathcal{M}=(\\mathcal{A}, D)$ such that $\\mathcal{A}=(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\Sigma})$ is a $\\Sigma-$multialgebra and $D$ is a subset of $A$, said to be its set of designated elements. \n\\end{definition}\n\nOne may check definition $6.3.1$ of \\cite{ParLog} and \\cite{Avron} for equivalent definitions, with slightly different emphases. \n\nGiven formulas $\\Gamma\\cup\\{\\varphi\\}$ over the signature $\\Sigma$ and an Nmatrix $\\mathcal{M}$, we say that $\\Gamma$ proves $\\varphi$ according to $\\mathcal{M}$ if, for every homomorphism of multialgebras $\\nu:\\textbf{F}(\\Sigma, \\mathcal{V})\\rightarrow \\mathcal{A}$ such that $\\nu(\\gamma)\\in D$ for every $\\gamma\\in\\Gamma$, one has $\\nu(\\varphi)\\in D$, when we then write $\\Gamma\\vDash_{\\mathcal{M}}\\varphi$: notice how close this definition is of that for $\\vDash_{\\mathcal{M}}$ when $\\mathcal{M}$ is simply a matrix. \n\nOne, perhaps very important, observation is that Nmatrices, as defined, are slightly redundant, being enough to add a symbol to our signature in order to work only their underlying multialgebra: consider an Nmatrix $\\mathcal{M}=(\\mathcal{A}, D)$ over the signature $\\Sigma$, add a symbol $\\top$ to $\\Sigma_{0}$ therefore producing the signature $\\Sigma^{\\top}$\\label{Sigmatop} and define the $\\Sigma^{\\top}$-multialgebra $\\mathcal{A}^{\\top}$ such that\n\\[\\sigma_{\\mathcal{A}^{\\top}}=\\sigma_{\\mathcal{A}}, \\quad\\text{for every $\\sigma$ in $\\Sigma$},\\]\nand $\\top_{\\mathcal{A}^{\\top}}=D$. Then, given formulas $\\Gamma\\cup\\{\\varphi\\}$ in the signature $\\Sigma$, we say that $\\Gamma$ proves $\\varphi$ according to $\\mathcal{A}^{\\top}$, and write $\\Gamma\\vDash_{\\mathcal{A}^{\\top}}\\varphi$, if for every homomorphism of $\\Sigma$-multialgebras $\\nu:\\textbf{F}(\\Sigma, \\mathcal{V})\\rightarrow\\mathcal{A}^{\\top}$,\n\\[\\nu(\\Gamma)\\subseteq \\top_{\\mathcal{A}^{\\top}}\\quad\\text{implies}\\quad \\nu(\\varphi)\\in \\top_{\\mathcal{A}^{\\top}},\\]\nwhat is possible given that $\\Sigma\\subseteq \\Sigma^{\\top}$ implies $\\Gamma\\cup\\{\\varphi\\}\\subseteq F(\\Sigma, \\mathcal{V})\\subseteq F(\\Sigma^{\\top}, \\mathcal{V})$.\nOf course, such a simplification is not possible when dealing with matrices, unless the set of designated values is a singleton.\n\nGiven a class $\\mathbb{M}$ of Nmatrices over the same signature $\\Sigma$, we say $\\Gamma\\vDash_{\\mathbb{M}}\\varphi$ if, for every $\\mathcal{M}\\in \\mathbb{M}$, $\\Gamma\\vDash_{\\mathcal{M}}\\varphi$. Given every matrix is an Nmatrix, we have that every tarskian logic may be characterized by a class of Nmatrices.\n\n\\begin{definition}\nA restricted non-deterministic matrix, or restricted Nmatrix or RNmatrix\\index{RNmatrix}\\index{Matrix, Restricted non-deterministic}, over a signature $\\Sigma$ is a triple \n\\[\\mathcal{M}=(\\mathcal{A}, D, \\mathcal{F})\\]\nsuch that:\n\\begin{enumerate}\n\\item $(\\mathcal{A}, D)$ is a non-deterministic matrix over $\\Sigma$;\n\\item $\\mathcal{F}$ is a subset of the set of all homomorphisms from $\\textbf{F}(\\Sigma, \\mathcal{V})$ to $\\mathcal{A}$.\n\\end{enumerate}\n\\end{definition}\n\n\nWe define $\\vDash_{\\mathcal{M}}$ as in the case that $\\mathcal{M}$ is a restricted matrix. Notice that every restricted matrix is a restricted Nmatrix, and then every tarskian logic may be characterized by a restricted Nmatrix, but we can achieve a more powerful result, similar to the one commonly known as Suszko`s Thesis, found in \\cite{Suszko}; but, unlike Suszko, we do not focus on bivaluations, and we do not wish to advocate that all logics are two-valued.\n\n\\begin{theorem}\\label{2-valued RNmatrix}\nEvery tarskian logic is characterizable by a two-valued RNmatrix.\\footnote{By an $n$-valued RNmatrix we understand an RNmatrix $\\mathcal{M}=(\\mathcal{A}, D, \\mathcal{F})$ where the universe of $\\mathcal{A}$ has $n$ elements.}\n\\end{theorem}\n\n\\begin{proof}\nLet $\\mathfrak{L}=(\\mathcal{L}, \\vdash)$ be a tarskian logic over the signature $\\Sigma$. Consider then the $\\Sigma$-multialgebra $\\textbf{2}(\\Sigma)$\\label{2Sigma} with universe $\\{0,1\\}$ and, for an $n$-ary $\\sigma\\in\\Sigma$, operations defined by\n\\[\\sigma_{\\textbf{2}(\\Sigma)}(x_{1}, \\dotsc , x_{n})=\\{0,1\\}, \\forall x_{1}, \\dotsc , x_{n}\\in \\{0,1\\}.\\]\nWe then define the set $\\mathcal{F}_{\\mathfrak{L}}$ of valuations $\\nu:\\textbf{F}(\\Sigma, \\mathcal{V})\\rightarrow\\textbf{2}(\\Sigma)$ such that there exists a closed set of formulas $\\Gamma$ over $\\Sigma$ for which $\\nu(\\gamma)=1$ if, and only if, $\\gamma\\in\\Gamma$; notice that all functions from $\\textbf{F}(\\Sigma, \\mathcal{V})$ to $\\textbf{2}(\\Sigma)$ are homomorphisms, and therefore no further restrictions are necessary.\n\nWe then define the RNmatrix \n\\[\\textbf{2}(\\mathfrak{L})=(\\textbf{2}(\\Sigma), \\{1\\}, \\mathcal{F}_{\\mathfrak{L}}),\\]\n\\label{2L}and we state that $\\Gamma\\vdash\\varphi$ if, and only if, $\\Gamma\\vDash_{\\textbf{2}(\\mathfrak{L})}\\varphi$. \n\nRegarding the first direction, if $\\Gamma\\vdash\\varphi$, for any valuation $\\nu\\in\\mathcal{F}_{\\mathfrak{L}}$ such that $\\nu(\\Gamma)\\subseteq\\{1\\}$ there must exist a closed set $\\Delta$ of formulas over $\\Sigma$ such that $\\nu(\\delta)=1$ if, and only if, $\\delta\\in\\Delta$, and since $\\nu(\\Gamma)\\subseteq\\{1\\}$ we have $\\Gamma\\subseteq \\Delta$; now, given $\\Gamma\\vdash\\varphi$ and $\\Gamma\\subseteq \\Delta$, it follows that $\\Delta\\vdash\\varphi$ and, since $\\Delta$ is closed, $\\varphi\\in \\Delta$, meaning $\\nu(\\varphi)=1$ and, therefore, $\\Gamma\\vDash_{\\textbf{2}(\\mathfrak{L})}\\varphi$.\n\nFor the second direction, assume $\\Gamma\\vDash_{\\textbf{2}(\\mathfrak{L})}\\varphi$: if $\\Gamma\\not\\vdash\\varphi$, by Lindenbaum-\\L oz there exists a non-trivial extension $\\Delta$ of $\\Gamma$ such that $\\varphi\\notin\\Delta$; if $\\nu$ is the valuation of $\\mathcal{F}_{\\mathfrak{L}}$ such that $\\nu(\\delta)=1$ if, and only if, $\\delta\\in\\Delta$, this means that $\\nu(\\Delta)\\subseteq\\{1\\}$, and therefore $\\nu(\\Gamma)\\subseteq\\{1\\}$ since $\\Gamma\\subseteq\\Delta$, and $\\nu(\\varphi)=0$, contradicting the fact that $\\Gamma\\vDash_{\\textbf{2}(\\mathfrak{L})}\\varphi$. This proves $\\Gamma\\vDash_{\\textbf{2}(\\mathfrak{L})}\\varphi$ implies $\\Gamma\\vdash\\varphi$, and the theorem is proved.\n\\end{proof}\n\nFor a restricted Nmatrix $\\mathcal{M}$ we will want to consider, for a subset $\\Gamma$ of $F(\\Sigma, \\mathcal{V})$, the closure $K_{\\mathcal{M}}(\\Gamma)$\\label{KMGamma}, that is, the set of formulas $\\varphi\\in F(\\Sigma, \\mathcal{V})$ such that $\\Gamma\\vDash_{\\mathcal{M}}\\varphi$.\n\n\\begin{definition}\nGiven a signature $\\Sigma$, an operator $K:\\mathcal{P}(F(\\Sigma, \\mathcal{V}))\\rightarrow \\mathcal{P}(F(\\Sigma, \\mathcal{V}))$ is said to be a tarskian\\index{Operator, Tarskian} (or closure) one if, for all $\\Gamma, \\Theta\\subseteq F(\\Sigma, \\mathcal{V})$, it satisfies:\n\\begin{enumerate}\n\\item $\\Gamma\\subseteq K(\\Gamma)$;\n\\item if $\\Theta\\subseteq \\Gamma$, $K(\\Theta)\\subseteq K(\\Gamma)$;\n\\item if $\\Theta=K(\\Gamma)$, $K(\\Theta)=\\Theta$.\n\\end{enumerate}\n\\end{definition}\n\nNotice how this generalizes the notion of a logic being tarskian: the consequence relation in a tarskian logic is a tarskian operator; for such a reason, one may often call the pair $(\\textbf{F}(\\Sigma, \\mathcal{V}), K)$, for $K$ a tarskian operator on $F(\\Sigma, \\mathcal{V})$, a sentential logic itself. Even more, we may define a logic $\\mathcal{L}$ over the signature $\\Sigma$ by $\\Gamma\\vdash_{\\mathcal{L}}\\varphi$ if, and only if, $\\varphi\\in K(\\Gamma)$, and it is clear how $\\mathcal{L}$ is tarskian: to every tarskian operator there corresponds a tarskian logic and vice-versa.\n\n\\begin{proposition}\nGiven a restricted Nmatrix $\\mathcal{M}=(\\mathcal{A}, D, \\mathcal{F})$, the operator $\\Gamma\\mapsto K_{\\mathcal{M}}(\\Gamma)$ is a tarskian one.\n\\end{proposition}\n\n\\begin{proof}\n\\begin{enumerate}\n\\item Take $\\varphi\\in \\Gamma$, and suppose $\\nu\\in\\mathcal{F}$ is such that $\\nu(\\gamma)\\in D$ for all $\\gamma\\in\\Gamma$; then $\\nu(\\varphi)\\in D$ and therefore $\\varphi\\in K_{\\mathcal{M}}(\\Gamma)$, so that $\\Gamma\\subseteq K_{\\mathcal{M}}(\\Gamma)$.\n\n\\item Suppose $\\Theta\\subseteq\\Gamma$ and that $\\varphi\\in K_{\\mathcal{M}}(\\Theta)$: then, if $\\nu\\in\\mathcal{F}$ satisfies that $\\nu(\\theta)\\in D$, for every $\\theta\\in\\Theta$, one has $\\nu(\\varphi)\\in D$.\n\nNow, take a restricted homomorphism $\\nu\\in\\mathcal{F}$ such that $\\nu(\\gamma)\\in D$, for every $\\gamma\\in\\Gamma$: since $\\Theta\\subseteq \\Gamma$, in this case $\\nu(\\theta)\\in D$, $\\forall \\theta\\in \\Theta$, and therefore $\\nu(\\varphi)\\in D$, implying that $\\varphi\\in K_{\\mathcal{M}}(\\Gamma)$. It follows that $K_{\\mathcal{M}}(\\Theta)\\subseteq K_{\\mathcal{M}}(\\Gamma)$.\n\n\\item By the first item above, $\\Theta\\subseteq K_{\\mathcal{M}}(\\Theta)$, so it only remains to be shown that $K_{\\mathcal{M}}(\\Theta)\\subseteq \\Theta$. Given one $\\varphi\\in K_{\\mathcal{M}}(\\Theta)$, for any $\\nu\\in\\mathcal{F}$ such that $\\nu(\\theta)\\in D$, for all $\\theta\\in\\Theta$, $\\nu(\\varphi)\\in D$.\n\nNow suppose $\\nu\\in\\mathcal{F}$ satisfies that $\\nu(\\gamma)\\in D$, for every $\\gamma\\in\\Gamma$: then, since $\\Theta=K_{\\mathcal{M}}(\\Gamma)$, $\\nu(\\theta)\\in D$ for all $\\theta\\in\\Theta$, and therefore $\\nu(\\varphi)\\in D$. It follows that $\\varphi\\in K_{\\mathcal{M}}(\\Gamma)=\\Theta$ and $K_{\\mathcal{M}}(\\Theta)=\\Theta$.\n\\end{enumerate}\n\\end{proof}\n\nGiven a class $\\mathbb{M}$ of RNmatrices, we can then consider the closure of a set $\\Gamma\\subseteq F(\\Sigma, X)$ under $\\mathbb{M}$, that is, $\\varphi\\in K_{\\mathbb{M}}(\\Gamma)$\\label{KMMGamma} if and only if $\\Gamma\\vDash_{\\mathbb{M}}\\varphi$. Clearly \n\\[K_{\\mathbb{M}}(\\Gamma)=\\bigcap_{\\mathcal{M}\\in\\mathbb{M}}K_{\\mathcal{M}}(\\Gamma).\\]\n\n\\begin{lemma}\nIf $K_{\\lambda}$ is a tarskian operator for every $\\lambda\\in\\Lambda$, $K$ defined as \n\\[K(\\Gamma)=\\bigcap_{\\lambda\\in\\Lambda}K_{\\lambda}(\\Gamma)\\]\nis also tarskian.\n\\end{lemma}\n\n\\begin{proof}\n\\begin{enumerate}\n\\item Since, for every $\\lambda\\in\\Lambda$ we have that $\\Gamma\\subseteq K_{\\lambda}(\\Gamma)$, we have that $\\Gamma\\subseteq \\bigcap_{\\lambda\\in\\Lambda}K_{\\lambda}(\\Gamma)=K(\\Gamma)$.\n\n\\item If $\\Theta\\subseteq \\Gamma$, we have $K_{\\lambda}(\\Theta)\\subseteq K_{\\lambda}(\\Gamma)$ for every $\\lambda\\in\\Lambda$, so \n\\[K(\\Theta)=\\bigcap_{\\lambda\\in\\Lambda}K_{\\lambda}(\\Theta)\\subseteq\\bigcap_{\\lambda\\in\\Lambda}K_{\\lambda}(\\Gamma)=K(\\Gamma).\\]\n\n\\item Finally, if $\\Theta=K(\\Gamma)$, then $\\bigcap_{\\lambda\\in\\Lambda}K_{\\lambda}(\\Gamma)=\\Theta$, so that $\\Theta\\subseteq K_{\\lambda}(\\Gamma)$ for every $\\lambda\\in\\Lambda$: clearly $\\Theta\\subseteq K(\\Theta)$, so it remains for us to show that $K(\\Theta)\\subseteq \\Theta$.\n\nLet us denote $K_{\\lambda}(\\Gamma)$ by $\\Theta_{\\lambda}$, and since $\\Theta\\subseteq K_{\\lambda}(\\Gamma)$, $K_{\\lambda}(\\Theta)\\subseteq K_{\\lambda}(\\Theta_{\\lambda})=\\Theta_{\\lambda}$, for every $\\lambda\\in\\Lambda$; notice that $\\Theta=\\bigcap_{\\lambda\\in\\Lambda}\\Theta_{\\lambda}$.\n\nThen, \n\\[K(\\Theta)=\\bigcap_{\\lambda\\in\\Lambda}K_{\\lambda}(\\Theta)\\subseteq \\bigcap_{\\lambda\\in\\Lambda}\\Theta_{\\lambda}=\\Theta,\\]\nwhat finishes the proof.\n\\end{enumerate}\n\\end{proof}\n\n\\begin{theorem}\nGiven a class $\\mathbb{M}$ of RNmatrices, $K_{\\mathbb{M}}$ is a tarskian operator.\n\\end{theorem}\n\nIn other words, classes of restricted Nmatrices can, at most, describe tarskian logics, and from Theorem \\ref{2-valued RNmatrix} they indeed characterize all of these. So no non-tarskian logic can be characterized by either RNmatrices or their classes,\\footnote{Although slight generalizations of RNmatrices could possibly change this.} but the fact is we do not see that as a real problem: we are not looking exclusively for expressive power in our semantics, being efficiency a more desirable property. That is, the point of restricted Nmatrices will not be what they can express, but how easily will be to define and use them. Most importantly, however, is that RNmatrices will be able to provide decision methods through what are, essentially, truth-tables, where none were available before: one example, we will stress repeatedly, is that of $C_{1}$; although decidable, this system is not only not characterizable by finite matrices, but neither by finite sets of finite matrices, finite Nmatrices or finite sets of finite Nmatrices (\\cite{Avron}). And, despite all of this, this logic, and in fact all of da Costa's hierarchy, may be characterized by finite RNmatrices, what we will achieve in Chapter \\ref{Chapter5}.\n\n\n\n\n\n\n\n\n\n\n\n\\subsection{A brief history of RNmatrices}\\label{A brief history of RNmatrices}\n\nWe first developed RNmatrices by studying semantics for the logic $\\textbf{mbC}$: when analyzing Fidel structures for that very system, we hoped to simplify the ($3$-valued) Nmatrix for $\\textbf{mbC}$ to a matrix, by slightly altering its signature. By replacing the unary connective, standing for consistency, for a binary connective ($\\uparrow$) for incompatibility, and therefore creating the logics of incompatibility we study in Chapters \\ref{Chapter7}, \\ref{Chapter8} and \\ref{Chapter9}, we created a system (within $\\textbf{nbI}^{-}$, that is, $\\textbf{nbI}$ without the commutativity of $\\uparrow$) equivalent to $\\textbf{mbC}$, with simpler semantics, but which was also, unfortunately, not characterizable by finite matrices.\n\nWell, a very natural system related to $\\textbf{nbI}^{-}$ is $\\textbf{bI}$, which adds the commutativity of the binary incompatibility connective and subtracts the negation; rather unfortunate was then our discovery that this logic is not characterizable by, not only finite matrices, but rather finite Nmatrices as well. But we had a semantic for it almost ready, simply by adapting the decision method (a two-valued Nmatrix) for $\\textbf{bI}^{-}$: it was enough to demand that every valuation to be taken into consideration should satisfy $\\nu(\\alpha\\uparrow\\beta)=\\nu(\\beta\\uparrow\\alpha)$. \n\nOf course, within a semantics of matrices, or even Nmatrices, this is not permissible; but, inspired by Piochi's work on $\\mathcal{E}$-matrices (which do restrict valuations) and our previous knowledge of Nmatrices, we have coined what we begun to call restricted non-deterministic matrices, or RNmatrices. We quickly grew to realize that RNmatrices are actually recurrent in the history of non-classical logics, although not defined as such but rather as a specific solution to a difficult system: examples abound, including \\index{Bivaluation}bivaluations, Fidel structures, Kearn's $4$-valued semantics for modal logics, static semantics, PNmatrices, among others.\n\nTo our great surprise, we found out that RNmatrices semantics have quite recently started to resurface: Pawlowski and Urbaniak, in \\cite{Pawlowski, PawlowskUrbaniak}, had noticed, shortly before us, previous uses of the semantics of RNmatrces (although their nomenclature is, naturally, different), specially in the areas of informal provability and modal logic. We add to the topic a more comprehensive analysis of the literature and a more in depth study of the expressiveness of these semantics, in addition to the first explicit use of RNmatrices to paraconsistent logics.\n\nAnd it is important, here, to explain what we mean by the expressiveness of RNmatrix semantics: usually, the expressiveness of semantics refers to which logics can these semantics characterize, a concept that is distinctively easy to understand when our semantics are matricial in nature. However, according to Theorem \\ref{2-valued RNmatrix}, RNmatrices are, in a way, as expressive as possible, meaning that any tarskian logic can be characterized by a finite RNmatrix. Hence, we are actually more concerned about, first of all, the expressiveness of decidable RNmatrices, \\textit{i. e.} RNmatrices where the problem of identifying which valuations are restricted valuations is decidable, see Section \\ref{Decision methods} of Chapter \\ref{Chapter5} for more details; second, the expressiveness of other, related semantics, such as PNmatrices of Example \\ref{example of PNMatrices} below, which we know to be no more expressive than RNmatrices (that is, all logics characterized by PNmatrices can also be characterized by RNmatrices), but are not sure whether they are as expressive as RNmatrices (meaning, if PNmatrices characterize all tarskian logics, as RNmatrices do). \n\n\n\n\n\n\n\n\n\\begin{example}\\label{Example Kearns}\nAs we have already discussed, Dugundji proved that the logics between the modal systems $\\textbf{S1}$ and $\\textbf{S5}$ can not be characterized by single, finite matrices; to circumvent that, J. Kearns\\index{Kearns} (\\cite{Kearns}) proposed a (four-valued) Nmatrix semantics for the modal $\\textbf{T}$, $\\textbf{S4}$ and $\\textbf{S5}$ for which just some specific valuations can be considered, what clearly characterizes these as RNmatrices semantics. Kearns idea proved itself to be very popular, and generalizations can be found in, \\textit{exempli gratia}, \\cite{CCP} (and \\cite{CCPErr}) and \\cite{OmoriSkurt}.\n\nUsing a more modern approach to Kearns technique, as in \\cite{CCP}, we proceed as follows:\\label{Semantics, Kearns} consider the universe $\\{T, t, f, F\\}$; operations are defined as necessary for each logic $\\textbf{L}\\in\\{\\textbf{T}, \\textbf{S4}, \\textbf{S5}\\}$. Consider the set $Val^{\\textbf{L}}$ of all valuations for $\\textbf{L}$; we define $Val_{k}^{\\textbf{L}}\\subseteq Val^{\\textbf{L}}$ by induction as $Val_{0}^{\\textbf{L}}=Val^{\\textbf{L}}$ and, for $k\\in\\mathbb{N}$,\\[Val_{k+1}^{\\textbf{L}}=\\{\\nu\\in Val_{k}^{\\textbf{L}} : \\text{for every formula $\\alpha$, $Val_{k}^{\\textbf{L}}(\\alpha)\\subseteq D$ implies $\\nu(\\alpha)=T$}\\},\\]\nwhere $Val_{k}^{\\textbf{L}}(\\alpha)=\\{\\mu(\\alpha) : \\mu\\in Val_{k}^{\\textbf{L}}\\}$. We then define the restricted valuations as $\\mathcal{F}_{\\textbf{L}}=\\bigcap_{k\\in\\mathbb{N}}Val_{k}^{\\textbf{L}}$, and Kearns proved that $\\vdash_{\\textbf{L}}\\alpha$ if, and only if, $\\nu(\\alpha)=T$ for every $\\nu\\in\\mathcal{F}_{\\textbf{L}}$, making his semantics for $\\textbf{L}$ correspond to the RNmatrix $\\mathcal{K}_{\\textbf{L}}=(\\mathcal{A}_{\\textbf{L}}, \\{T\\}, \\mathcal{F}_{\\textbf{L}})$, for $\\mathcal{A}_{\\textbf{L}}$ the multialgebra for $\\textbf{L}$, with universe $\\{T, t, f, F\\}$, whose precise definition we omitted.\n\nNow, we can also prove $\\mathcal{K}_{\\textbf{L}}$\\label{KL} is structural: through an induction on $k$, one shows that, for every valuation $\\nu\\in Val_{k}^{\\textbf{L}}$ and every substitution $\\rho$, $\\nu\\circ\\rho\\in Val_{k}^{\\textbf{L}}$, being the case $k=0$ trivially true; so, suppose the result holds for a certain $k\\in\\mathbb{N}$. Given $\\nu\\in Val_{k+1}^{\\textbf{L}}$ and a substitution $\\rho$, by definition of $Val_{k+1}^{\\textbf{L}}$ we have $\\nu\\in Val_{k}^{\\textbf{L}}$ and, by induction hypothesis, $\\nu\\circ\\rho\\in Val_{k}^{\\textbf{L}}$. Let $\\alpha$ be a formula satisfying $Val_{k}^{\\textbf{L}}(\\alpha)\\subseteq D$: for every $\\mu\\in Val_{k}^{\\textbf{L}}$, $\\mu\\circ\\rho\\in Val_{k}^{\\textbf{L}}$ (again by induction hypothesis), meaning that $\\mu\\circ\\rho(\\alpha)=\\mu(\\rho(\\alpha))\\in D$, and therefore $Val_{k}^{\\textbf{L}}(\\rho(\\alpha))\\subseteq D$.\n\nThis implies, since $\\nu\\in Val_{k+1}^{\\textbf{L}}$, that $\\nu\\circ\\rho(\\alpha)=\\nu(\\rho(\\alpha))=T$, and henceforth $\\nu\\circ\\rho\\in Val_{k+1}^{\\textbf{L}}$. Obviously this proves that $\\nu\\in Val_{k}^{\\textbf{L}}$ implies $\\nu\\circ\\rho\\in Val_{k}^{\\textbf{L}}$ for all $k\\in\\mathbb{N}$: so, if $\\nu\\in \\mathcal{F}_{\\textbf{L}}$ and $\\rho$ is a substitution, for every $k\\in \\mathbb{N}$, $\\nu\\in Val_{k}^{\\textbf{L}}$ implies $\\nu\\circ\\rho\\in Val_{k}^{\\textbf{L}}$ (again, for every $k\\in\\mathbb{N}$), and so $\\nu\\circ\\rho\\in \\bigcap_{k\\in\\mathbb{N}}Val_{k}^{\\textbf{L}}=\\mathcal{F}_{\\textbf{L}}$.\n\\end{example}\n\n\\begin{example}\nThe first proof of the decidability of da Costa's calculi $C_{n}$ (see Chapter \\ref{Chapter5} for a definition of these logics) is due to Fidel, which in 1977 created a new class of structures, both algebraic and relational, to provide precisely such a demonstration in \\cite{Fidel3}: these objects are now known as \\textit{Fidel structures}\\index{Fidel structure}. Essentially, a \\index{Fidel structure for $C_{n}$}Fidel structure for $C_{n}$, which in this presentation are equipped with a connective $(n)$, is a triple $\\mathcal{N}=(\\mathcal{A}, \\{N_{a}\\}_{a\\in A}, \\{N_{a}^{(n)}\\}_{a\\in A})$ such that:\n\\begin{enumerate}\n\\item $\\mathcal{A}$ is a Boolean algebra with universe $A$;\n\\item for every $a\\in A$, $N_{a}$\\label{Na} and $N_{a}^{(n)}$ are subsets of $A$ (and therefore unary relations indexed by $A$) satisfying certain desired properties.\n\\end{enumerate}\nA valuation over a Fidel structure $\\mathcal{N}$ for $C_{n}$ is a function $\\nu$ from the formulas of $C_{n}$ to $A$ satisfying, among other conditions, that $\\nu(\\alpha\\#\\beta)=\\nu(\\alpha)\\#\\nu(\\beta)$, for $\\#\\in\\{\\vee, \\wedge, \\rightarrow\\}$, \n\\[\\nu(\\neg\\alpha)\\in N_{\\nu(\\alpha)}\\quad\\text{and}\\quad\\nu(\\alpha^{(n)})\\in N_{\\nu(\\alpha)}^{(n)}.\\]\nIf, for every element $a$ of $\\mathcal{A}$, we define $\\neg a=N_{a}$ and $n(a)=N_{a}^{(n)}$, we have enriched $\\mathcal{A}$ to become a multialgebra, which we denote by $\\mathcal{A}_{\\mathcal{N}}^{n}$; it is not difficult to see, then, that the consequence operator induced by the Fidel structure $\\mathcal{N}$ equals the one produced by the RNmatrix $\\mathcal{M}_{\\mathcal{N}}^{n}=(\\mathcal{A}_{\\mathcal{N}}^{n}, D, \\mathcal{F}_{\\mathcal{N}}^{n})$, for $D=\\{1\\}$ and $\\mathcal{F}_{\\mathcal{N}}^{n}$ the set of valuations for $C_{n}$ over $\\mathcal{N}$.\n\nThus, we understand Fidel structures as one of the earliest, and best understood, applications of RNmatrices; more important, however, is its success in surviving the test of time, being a methodology that is still relevantly applied today. See, for an example, the characterizations of $\\textbf{LFI}$'s such as $\\textbf{mbC}$, $\\textbf{mbCcl}$ and $\\textbf{CILA}$ in \\cite{ParLog} through Fidel structures, all of which may be recast as RNmatrices, one may add.\n\nNow, to be more precise, \\cite{Fidel3} uses a Fidel structure called $\\textbf{C}$ over the two-element Boolean algebra $\\textbf{2}$ as a decision procedure for all of $C_{n}$: that way, for any $n\\geq 1$, the multialgebras $\\mathcal{A}_{\\textbf{C}}^{n}$ are the same and satisfy $\\neg 0=n(0)=\\{1\\}$ and $\\neg1=n(1)=\\{0,1\\}$; this forces most of the semantical power of these structures to lie in their valuations, making of $\\mathcal{F}_{\\textbf{C}}^{n}$ rather complicated sets (look, for an example of a similar nature, to $\\textbf{2}(\\mathfrak{L})$ in Theorem \\ref{2-valued RNmatrix}). In a certain sense, when we present our own RNmatrices for $C_{n}$ in Chapter \\ref{Chapter5} we make a compromise: we allow for the underlying multialgebras to have larger universes (with $n+2$ elements for $C_{n}$), but in return we have far simpler valuations.\n\\end{example}\n\n\n\\begin{example}\nIn the work of Avron and Konikowska \\cite{AK:05}, valuations over Nmatrices induce semantics which they called dynamic semantics\\index{Semantics, Dynamic} over Nmatrices; as opposed to these, they have also considered semantics which restrict the usual valuations, what they have baptized as static semantics\\index{Semantics, Static}. Essentially, given an Nmatrix $\\mathcal{M}=(\\mathcal{A}, D)$, its static semantics is given by $\\mathcal{M}^{s}=(\\mathcal{A}, D, \\mathcal{F}^{s}_{\\mathcal{M}})$\\label{Ms}, for $\\mathcal{F}^{s}_{\\mathcal{M}}$ the set of valuations $\\nu$ over $\\mathcal{M}$ satisfying that: for an $n$-ary connective $\\sigma$, and formulas $\\sigma(\\alpha_{1}, \\dotsc , \\alpha_{n})$ and $\\sigma(\\beta_{1}, \\dotsc , \\beta_{n})$, $\\nu(\\alpha_{i})=\\nu(\\beta_{i})$, for every $i\\in\\{1, \\dotsc , n\\}$, implies\n\\[\\nu(\\sigma(\\alpha_{1}, \\dotsc , \\alpha_{n}))=\\nu(\\sigma(\\beta_{1}, \\dotsc , \\beta_{n})).\\]\nIt is easy to prove that the RNmatrices $\\mathcal{M}^{s}$ are all structural.\n\\end{example}\n\n\\begin{example}\\label{example of PNMatrices}\nPNmatrices\\index{PNmatrix}\\index{Matrix, Partial non-deterministic}, whose first appearance in the literature was in \\cite{Baaz}, generalize Nmatrices by replacing multialgebras for partial multialgebras (that is, algebras of relations). Our brief presentation here, however, is closer to that of \\cite{CalMar}: a partial $\\Sigma$-multialgebra is a pair $\\mathcal{A}=(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma})$ such that, if $\\sigma\\in\\Sigma_{n}$, $\\sigma_{\\mathcal{A}}$ is a function of the form $\\sigma_{\\mathcal{A}}:A^{n}\\rightarrow\\mathcal{P}(A)$ (no longer $\\mathcal{P}(A)\\setminus\\{\\emptyset\\}$ as in the case of multialgebras). A valuation for a partial $\\Sigma$-multialgebra $\\mathcal{A}$, as described, is then a map $\\nu: F(\\Sigma, \\mathcal{V})\\rightarrow A$ such that\n\\[\\nu(\\sigma(\\alpha_{1}, \\dotsc , \\alpha_{n}))\\in\\sigma_{\\mathcal{A}}(\\nu(\\alpha_{1}), \\dotsc , \\nu(\\alpha_{n})),\\]\nfor every $\\sigma\\in\\Sigma_{n}$ and formulas $\\alpha_{1}, \\dotsc , \\alpha_{n}$; notice that, if $\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})=\\emptyset$, then there cannot exist a valuation $\\nu$ and formulas $\\alpha_{1}$ trough $\\alpha_{n}$ such that $\\nu(\\alpha_{i})=a_{i}$, for $1\\leq i\\leq n$, given $\\nu$ is supposed to be a total function. Finally, given a partial $\\Sigma$-multialgebra $\\mathcal{A}$ and a subset $D$ of its universe, $\\mathcal{M}=(\\mathcal{A}, D)$ is a PNmatrix: given formulas $\\Gamma\\cup\\{\\varphi\\}$ on $\\Sigma$, we say $\\Gamma$ proves $\\varphi$ according to $\\mathcal{M}$, and write $\\Gamma\\vDash_{\\mathcal{M}}\\varphi$, whenever, for every valuation $\\nu$ for $\\mathcal{A}$, $\\nu(\\Gamma)\\subseteq D$ implies $\\nu(\\varphi)\\in D$.\n\nWhat we can prove, however, is that the semantics induced by PNmatrices can be also induced by RNmatrices, making of the latter a more expressive technique. Given a PNmatrix $\\mathcal{M}=(\\mathcal{A}, D)$, over $\\Sigma$, we consider the $\\Sigma$-multialgebra $\\mathcal{A}^{\\emptyset}=(A\\cup\\{o\\}, \\{\\sigma_{\\mathcal{A}^{\\emptyset}}\\}_{\\sigma\\in\\Sigma})$, for an element $o\\notin A$, such that\n\\[\\sigma_{\\mathcal{A}^{\\emptyset}}(a_{1}, \\dotsc , a_{n})=\n\\begin{cases*}\n\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n}) & if $a_{1}, \\dotsc , a_{n}\\in A$ and $\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})\\neq\\emptyset$,\\\\\n\\{o\\} & otherwise,\n\\end{cases*}\\]\nfor all $a_{1}, \\dotsc , a_{n}\\in A\\cup\\{o\\}$. By defining\n\\[\\mathcal{F}=\\{\\nu:\\textbf{F}(\\Sigma, \\mathcal{V})\\rightarrow\\mathcal{A}^{\\emptyset}\\ :\\ o\\notin\\nu(F(\\Sigma, \\mathcal{V}))\\},\\]\nwhere $\\nu(F(\\Sigma, \\mathcal{V}))=\\{\\nu(\\alpha) : \\alpha\\in F(\\Sigma, \\mathcal{V})\\}$, one sees that the valuations for $\\mathcal{A}$ correspond, more or less, to $\\mathcal{F}$, that is, the set of valuations for $\\mathcal{A}^{\\emptyset}$ without $o$ in their range; then $\\mathcal{M}^{\\emptyset}=(\\mathcal{A}^{\\emptyset}, D, \\mathcal{F})$\\label{Mo} has the same deduction operator as $\\mathcal{A}$.\n\\end{example}\n\n\\begin{example}\\label{PTS}\nPossible-translations semantics were first defined in \\cite{PTS-first}, but here we use an approach similar to that found in \\cite{PTS-defined}, the difference being that we only focus on zeroth order logics: take signatures $\\Sigma$ and $\\Sigma^{*}$ and its respective languages, $\\mathcal{L}_{\\Sigma}$ and $\\mathcal{L}_{\\Sigma^{*}}$; a translation between logics $\\mathscr{L}=(\\mathcal{L}_{\\Sigma}, \\vdash_{\\mathscr{L}})$ and $\\mathscr{L}^{*}=(\\mathcal{L}_{\\Sigma^{*}}, \\vdash_{\\mathscr{L}^{*}})$ is any function $t:\\mathcal{L}_{\\Sigma}\\rightarrow \\mathcal{L}_{\\Sigma^{*}}$ such that, for any set of formulas $\\Gamma\\cup\\{\\varphi\\}$ of $\\mathcal{L}_{\\Sigma}$,\n\\[\\text{$\\Gamma\\vdash_{\\mathscr{L}}\\varphi$\\quad implies\\quad $t(\\Gamma)\\vdash_{\\mathscr{L}^{*}}t(\\varphi)$,}\\]\nwhere $t(\\Gamma)=\\{t(\\gamma)\\ :\\ \\gamma\\in\\Gamma\\}$. Fixed a system $\\mathscr{L}$ over $\\Sigma$, a possible-translations semantics\\index{Possible-translations semantics} for $\\mathscr{L}$ is then a pair $\\mathcal{PT}=(\\textbf{Log}, \\textbf{Tr})$ where $\\textbf{Log}=\\{\\mathscr{L}_{i}\\}_{i\\in I}$ are logics over possibly distinct signatures $\\Sigma^{i}$, and $\\textbf{Tr}=\\{t_{i}\\}_{i\\in I}$ are translations from $\\mathscr{L}$ to $\\mathscr{L}_{i}$. The possible-translations semantics leads to a consequence operator $\\Vdash_{\\mathcal{PT}}$ on the formulas of the signature $\\Sigma$ whenever we define, for a set of formulas $\\Gamma\\cup\\{\\varphi\\}$ over $\\Sigma$,\n\\[\\text{$\\Gamma\\Vdash_{\\mathcal{PT}}\\varphi$\\quad if and only if\\quad $t_{i}(\\Gamma_{i})\\vdash_{\\mathscr{L}_{i}}t_{i}(\\varphi)$,\\quad for all $i\\in I$.}\\]\nOf course, $\\mathscr{L}$ is characterized by $\\mathcal{PT}$ when $\\Gamma\\vdash_{\\mathscr{L}}\\varphi$ iff $\\Gamma\\Vdash_{\\mathcal{PT}}\\varphi$; intuitively, a logic is characterized by a given possible-translations semantics whenever $\\mathscr{L}$ can be defined by specific properties of the $\\mathscr{L}_{i}$, which the translations $t_{i}$ combine together to obtain $\\Vdash_{\\mathcal{PT}}$. \n\nOf course, RNmatrices cannot possibly be as expressive as even this weaker definition of possible-translations semantics, given that they can only deal with one signature at a time,\\footnote{It seems, however, possible to combine RNmatrices over different signatures as long as one has at their disposal adequate maps between signatures, sometimes also known as translations.} but a connection between the two semantics is self-evident: both of them shift the expressive power of semantics, from structures to maps themselves; specifically, translations in the case of possible-translations semantics, and restricted valuations in the case of RNmatrices.\n\nHowever, consider a specific case of possible-translations semantics: suppose all logics of $\\textbf{Log}$ have the same signature $\\Sigma$ as $\\mathscr{L}$, and that all of them are characterized by RNmatrices $\\mathcal{M}_{i}$ or, what is equivalent, that all of them are tarskian (and, therefore, so is $\\mathscr{L}$); then $\\mathscr{L}$ is characterized by a class of of RNmatrices in a straightforward way. In fact, if $\\mathcal{M}_{i}=(\\mathcal{A}_{i}, D_{i}, \\mathcal{F}_{i})$, then \n\\[\\text{$\\Gamma\\vdash_{\\mathscr{L}_{i}}\\varphi$\\quad iff \\quad $\\nu(\\Gamma)\\subseteq D_{i}$\\quad implies\\quad $\\nu(\\varphi)\\in D_{i}$\\quad for every $\\nu\\in\\mathcal{F}_{i}$;}\\]\nby defining $\\mathcal{F}_{i}^{t}=\\{\\nu\\circ t_{i}\\ :\\ \\nu\\in\\mathcal{F}_{i}\\}$ and $\\mathcal{M}_{i}^{t}=(\\mathcal{A}_{i}, D_{i}, \\mathcal{F}_{i}^{t})$, it is easy to see that $t_{i}(\\Gamma)\\vdash_{\\mathscr{L}_{i}}t_{i}(\\varphi)$ if, and only if, $\\Gamma\\vDash_{\\mathcal{M}_{i}^{t}}\\varphi$. By making $\\mathbb{M}=\\{\\mathcal{M}_{i}^{t}\\}_{i\\in I}$, we see that $\\Gamma\\Vdash_{\\mathcal{PT}}\\varphi$ is equivalent to $\\Gamma\\vDash_{\\mathbb{M}}\\varphi$, and so $\\mathscr{L}$ is characterized by $\\mathbb{M}$.\n\n\n\n\n\n\n\\end{example}\n\n\n\n\n\\subsection{Structurality}\\label{Structurality}\n\n\\begin{definition}\nFor any subsemigroup $\\mathcal{E}$ of the semigroup $\\textit{End}(F(\\Sigma, \\mathcal{V}))$ of endomorphisms on $F(\\Sigma, \\mathcal{V})$, we say an operator $K:\\mathcal{P}(F(\\Sigma, \\mathcal{V}))\\rightarrow\\mathcal{P}(F(\\Sigma, \\mathcal{V}))$ is $\\mathcal{E}$-structural\\index{Structural, $\\mathcal{E}$-} when \n\\[\\{\\sigma(\\varphi)\\ :\\ \\varphi\\in K(\\Gamma)\\}\\subseteq K(\\{\\sigma(\\varphi)\\ :\\ \\varphi\\in\\Gamma\\})\\]\nfor every $\\sigma\\in\\mathcal{E}$, or as we shall write it, $\\sigma K(\\Gamma)\\subseteq K(\\sigma\\Gamma)$; the operator is said to be structural when it is $\\textit{End}(F(\\Sigma, \\mathcal{V}))$-structural. \n\\end{definition}\n\nStructurality is rather important when dealing with logics arising from a Hilbert calculus given that instances of axiom schemata and rules of inference are obtained by application of endomorphisms; so, when studying a deduction operator, it is important to understand how structural it is. While most of the systems we study here are structural, not only $\\mathcal{E}$-structural for a proper subsemigroup $\\mathcal{E}$, developing a general theory of structurality for RNmatrices, as Piochi did for Rmatrices in \\cite{Piochi}, \\cite{Piochi2} and \\cite{Piochi3}, can have useful applications on non-structural logics.\n\n\\begin{proposition}\nIf $\\mathcal{M}=(\\mathcal{A}, D, \\mathcal{F})$ is an RNmatrix such that, for every $\\nu\\in\\mathcal{F}$ and $\\sigma\\in\\mathcal{E}$, $\\nu\\circ\\sigma\\in\\mathcal{F}$, then $K_{\\mathcal{M}}$ is $\\mathcal{E}$-structural.\n\\end{proposition}\n\n\\begin{proof}\nLet $\\Gamma$ be a subset of $F(\\Sigma, \\mathcal{V})$ and $\\varphi\\in K_{\\mathcal{M}}(\\Gamma)$: we must prove that, for a given $\\sigma\\in \\mathcal{E}$, $\\sigma(\\varphi)$ is in $K_{\\mathcal{M}}(\\sigma\\Gamma)$, or what is equivalent, that $\\sigma\\Gamma\\vDash_{\\mathcal{M}}\\sigma(\\varphi)$. So, let $\\nu\\in\\mathcal{F}$ be a valuation satisfying $\\nu(\\sigma(\\gamma))=\\nu\\circ\\sigma(\\gamma)\\in D$, $\\forall\\gamma\\in\\Gamma$, and we must prove that $\\nu(\\sigma(\\varphi))=\\nu\\circ\\sigma(\\varphi)\\in D$.\n\nBy our hypothesis, $\\nu\\circ\\sigma\\in\\mathcal{F}$, and since $\\nu\\circ\\sigma(\\gamma)\\in D$ for every $\\gamma\\in\\Gamma$ and $\\varphi\\in K_{\\mathcal{M}}(\\Gamma)$ (meaning $\\Gamma\\vDash_{\\mathcal{M}}\\varphi$), we find that $\\nu\\circ\\sigma(\\varphi)\\in D$, what ends the proof.\n\\end{proof}\n\n\\begin{lemma}\nIf all operators $K_{\\lambda}$, for $\\lambda\\in\\Lambda$, are $\\mathcal{E}$-structural, then so it is $K$ defined by \n\\[K(\\Gamma)=\\bigcap_{\\lambda\\in\\Lambda}K_{\\lambda}(\\Gamma),\\]\nfor every $\\Gamma\\subseteq F(\\Sigma, \\mathcal{V})$.\n\\end{lemma}\n\n\\begin{proof}\nTake $\\sigma\\in\\mathcal{E}$ and $\\varphi\\in K(\\Gamma)$: we have that $\\varphi\\in K_{\\lambda}(\\Gamma)$ for every $\\lambda\\in\\Lambda$, and then $\\sigma(\\varphi)\\in K_{\\lambda}(\\sigma\\Gamma)$, since each $K_{\\lambda}$ is structural.\n\nIt follows that $\\sigma(\\varphi)\\in\\bigcap_{\\lambda\\in\\Lambda}K_{\\lambda}(\\sigma\\Gamma)=K(\\sigma\\Gamma)$, and so $\\sigma K(\\Gamma)\\subseteq K(\\sigma\\Gamma)$.\n\\end{proof}\n\n\\begin{theorem}\nGiven a class $\\mathbb{M}$ of RNmatrices $\\mathcal{M}=(\\mathcal{A}, D, \\mathcal{F})$ such that, for every $\\nu\\in\\mathcal{F}$ and $\\sigma\\in\\mathcal{E}$, $\\nu\\circ\\sigma\\in\\mathcal{F}$, $K_{\\mathbb{M}}$ is a $\\mathcal{E}$-structural operator.\n\\end{theorem}\n\n\nNow, for any tarskian logic $\\mathfrak{L}$, we remember for an instant the two-valued RNmatrix $\\textbf{2}(\\mathfrak{L})$ which characterizes $\\mathfrak{L}$ from Theorem \\ref{2-valued RNmatrix}: notice that, if $\\mathfrak{L}$ is $\\mathcal{E}$-structural, so is the deduction operator induced by $\\textbf{2}(\\mathfrak{L})$. To see that, take an endomorphism $\\sigma$ in $\\mathcal{E}$; if $\\Gamma\\vDash_{\\textbf{2}(\\mathfrak{L})}\\varphi$, $\\Gamma\\vdash\\varphi$, and from the fact that $\\mathfrak{L}$ is $\\mathcal{E}$-structural, $\\{\\sigma(\\gamma)\\ :\\ \\gamma\\in\\Gamma\\}\\vdash\\sigma(\\varphi)$; this means, of course, that \n\\[\\{\\sigma(\\gamma)\\ :\\ \\gamma\\in\\Gamma\\}\\vDash_{\\textbf{2}(\\mathfrak{L})}\\sigma(\\varphi),\\]\nwhat proves the result. Of course, if $\\mathfrak{L}$ is structural, so is the operator of $\\textbf{2}(\\mathfrak{L})$.\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\\section{Examples: characterizing some logics with RNmatrices}\\label{Examples: characterizing some logics with RNmatrices}\n\nWe would like to know that our approach using restricted Nmatrices has something to offer that other approaches don't. It is a classical result by Avron, found in \\cite{Avron}, that no finite Nmatrix can characterize the logics between $\\mbCcl$ and $\\CILA$, both included: the problem arises specifically from axiom $\\cl$. This is an important Dugundji-like uncharacterizability theorem, that limits greatly the expressive power of Nmatrices. We will show, by offering examples of such finite RNmatrices, that no such restriction exists for finite RNmatrices over those logics.\n\n\\subsection{$\\mbCcl$}\\label{RNmatrix for mbCcl}\n\nThis logic, together with the closely related $\\mbCci$, were first defined, under these names, in \\cite{ParLog}; however, the system $\\textbf{B}[\\{\\textbf{i1}, \\textbf{i2}\\}]$\\label{Bi1i2}, equivalent to $\\mbCci$, appeared already in \\cite{Avron2}, while $\\textbf{Bi}$\\label{Bi} and $\\textbf{Bl}$\\label{Bl}, equivalent to respectively $\\mbCci$ and $\\mbCcl$, had also been defined in \\cite{Avron}.\n\nConsider the $\\Sigma_{\\textbf{LFI}}$-multialgebra $\\mathcal{A}_{\\mbCciw}$ with universe $\\{F, t, T\\}$ and operations given by the tables below.\n\\begin{figure}[H]\n\\centering\n\\begin{minipage}[t]{4cm}\n\\centering\n\\begin{tabular}{|l|c|c|r|}\n\\hline\n$\\vee$ & $F$ & $t$ & $T$ \\\\ \\hline\n$F$ & $\\{F\\}$ & $\\{t, T\\}$ & $\\{t, T\\}$ \\\\ \\hline\n$t$ & $\\{t, T\\}$ & $\\{t, T\\}$ & $\\{t, T\\}$ \\\\ \\hline\n$T$ & $\\{t, T\\}$ & $\\{t, T\\}$ & $\\{t, T\\}$ \\\\ \\hline\n\\end{tabular}\n\\caption*{Table for Disjunction}\n\\end{minipage}\n\\hspace{3cm}\n\\centering\n\\begin{minipage}[t]{4cm}\n\\centering\n\\begin{tabular}{|l|c|c|r|}\n\\hline\n$\\wedge$ & $F$ & $t$ & $T$ \\\\ \\hline\n$F$ & $\\{F\\}$ & $\\{F\\}$ & $\\{F\\}$ \\\\ \\hline\n$t$ & $\\{F\\}$ & $\\{t, T\\}$ & $\\{t, T\\}$ \\\\ \\hline\n$T$ & $\\{F\\}$ & $\\{t, T\\}$ & $\\{t, T\\}$ \\\\ \\hline\n\\end{tabular}\n\\caption*{Table for Conjunction}\n\\end{minipage}\n\\end{figure}\n\n\\begin{figure}[H]\n\\centering\n\\begin{minipage}[t]{3cm}\n\\centering\n\\begin{tabular}{|l|r|}\n\\hline\n & $\\neg$ \\\\ \\hline\n$F$ & $\\{t, T\\}$\\\\ \\hline\n$t$ & $\\{t, T\\}$\\\\ \\hline\n$T$ & $\\{F\\}$ \\\\ \\hline\n\\end{tabular}\n\\caption*{Table for $\\neg$}\n\\end{minipage}\n\\begin{minipage}[t]{5cm}\n\\centering\n\\begin{tabular}{|l|c|c|r|}\n\\hline\n$\\rightarrow$ & $F$ & $t$ & $T$ \\\\ \\hline\n$F$ & $\\{t, T\\}$ & $\\{t, T\\}$ & $\\{t, T\\}$ \\\\ \\hline\n$t$ & $\\{F\\}$ & $\\{t, T\\}$ & $\\{t, T\\}$ \\\\ \\hline\n$T$ & $\\{F\\}$ & $\\{t, T\\}$ & $\\{t, T\\}$ \\\\ \\hline\n\\end{tabular}\n\\caption*{Table for Implication}\n\\end{minipage}\n\\begin{minipage}[t]{3cm}\n\\centering\n\\begin{tabular}{|l|r|}\n\\hline\n & $\\circ$ \\\\ \\hline\n$F$ & $\\{t, T\\}$\\\\ \\hline\n$t$ & $\\{F\\}$\\\\ \\hline\n$T$ & $\\{t,T\\}$ \\\\ \\hline\n\\end{tabular}\n\\caption*{Table for $\\circ$}\n\\end{minipage}\n\\end{figure}\n\nWhen making $D=\\{t, T\\}$, is it is shown in Corollary $6.5.5$ of \\cite{ParLog} that the Nmatrix $\\mathcal{M}_{\\mbCciw}=(\\mathcal{A}_{\\mbCciw}, D)$\\label{MmbCciw} is adequate for $\\mbCciw$, that is, for any set of formulas $\\Gamma\\cup\\{\\varphi\\}$ over the signature $\\Sigma_{\\textbf{LFI}}$ we have $\\Gamma\\vdash_{\\mbCciw}\\varphi$ if, and only if, $\\Gamma\\vDash_{\\mathcal{M}_{\\mbCciw}}\\varphi$.\n\nNow consider the restricted Nmatrix \\label{MmbCcl}\n\\[\\mathcal{M}_{\\mbCcl}=(\\mathcal{A}_{\\mbCciw}, D, \\mathcal{F}_{\\mbCcl})\\]\nsuch that $\\mathcal{F}_{\\mbCcl}$ is the set of homomorphisms $\\nu:\\textbf{F}(\\Sigma_{\\textbf{LFI}},\\mathcal{V})\\rightarrow\\mathcal{A}_{\\mbCcl}$ satisfying that, if $\\nu(\\alpha)=t$, then $\\nu(\\alpha\\wedge\\neg\\alpha)=T$. Clearly such an RNmatrix is structural, since for any endomorphism $\\sigma$ of $\\textbf{F}(\\Sigma_{\\textbf{LFI}},\\mathcal{V})$ we have that if $\\nu\\circ\\sigma(\\alpha)=t$, then $\\nu(\\sigma(\\alpha))=t$, meaning $\\nu(\\sigma(\\alpha)\\wedge\\neg\\sigma(\\alpha))=T$ and, therefore, $\\nu\\circ\\sigma(\\alpha\\wedge\\neg\\alpha)=T$.\n\nIt is easy to see $\\mathcal{M}_{\\mbCcl}$ models the axiom schemata and rules of inference of\\\\ $\\mbCciw$, but it is also true that, given an instance $\\psi=\\neg(\\alpha\\wedge\\neg\\alpha)\\rightarrow\\circ\\alpha$ of $\\cl$, we have $\\vDash_{\\mathcal{M}_{\\mbCcl}}\\psi$: assume that, for $\\nu\\in\\mathcal{F}_{\\mbCcl}$, $\\nu(\\psi)\\notin D$, meaning that $\\nu(\\neg(\\alpha\\wedge\\neg\\alpha))\\in D$ and $\\nu(\\circ\\alpha)=F$, which in turn implies $\\nu(\\alpha)=t$.\n\nSince $\\nu\\in\\mathcal{F}_{\\mbCcl}$, $\\nu(\\alpha)=t$ implies $\\nu(\\alpha\\wedge\\neg\\alpha)=T$, and therefore $\\nu(\\neg(\\alpha\\wedge\\neg\\alpha))=F$, reaching a contradiction. We have, then, that $\\nu(\\psi)\\in D$, for any $\\nu\\in\\mathcal{F}_{\\mbCcl}$.\n\n\\begin{theorem}\nGiven formulas $\\Gamma\\cup\\{\\varphi\\}$ of $\\mbCcl$, if $\\Gamma\\vdash_{\\mbCcl}\\varphi$ then $\\Gamma\\vDash_{\\mathcal{M}_{\\mbCcl}}\\varphi$.\n\\end{theorem}\n\n\\begin{proof}\nIf $\\Gamma\\vdash_{\\mbCcl}\\varphi$, there exists a demonstration $\\alpha_{1}, \\dotsc , \\alpha_{n}$ of $\\varphi$ from $\\Gamma$, with $\\alpha_{n}=\\varphi$.\n\nLet $\\nu\\in\\mathcal{F}_{\\mbCcl}$ be a valuation satisfying that that $\\nu(\\Gamma)\\subseteq D$: we want to prove that, in this case, $\\nu(\\varphi)\\in D$; so we prove, by induction, that $\\alpha_{1}$ through $\\alpha_{n}$ have image in $D$ under $\\nu$, and therefore $\\nu(\\varphi)=\\nu(\\alpha_{n})=1$.\n\nThe formula $\\alpha_{1}$ is either an instance of an axiom, when $\\nu(\\alpha_{1})\\in D$ since all instances of axioms have image in $D$ through any elements of $\\mathcal{F}_{\\mbCcl}$, or $\\alpha_{1}$ is a premise, that is, an element of $\\Gamma$, and since $\\nu(\\Gamma)\\subseteq D$ we have that $\\nu(\\alpha_{1})\\in D$.\n\nSuppose then that $\\nu(\\alpha_{1}), \\dotsc , \\nu(\\alpha_{i-1})\\in D$, and we have three cases to consider:\n\\begin{enumerate}\n\\item if $\\alpha_{i}$ is an instance of an axiom, as mentioned above $\\nu(\\alpha_{i})\\in D$;\n\\item if $\\alpha_{i}$ is a premise, $\\nu(\\alpha_{i})\\in D$ since $\\alpha_{i}\\in\\Gamma$ and $\\nu(\\Gamma)\\subseteq D$;\n\\item if there are $\\alpha_{j}$ and $\\alpha_{k}$ with $j,k2$.\n\nFor simplicity, we define $\\alpha^{0}=\\alpha$\\label{alpha^} and\n\\[\\alpha^{n+1}=\\neg(\\alpha^{n}\\wedge\\neg(\\alpha^{n}))\\]\nfor $n\\in\\mathbb{N}$; we also define $\\alpha^{(0)}=\\alpha$, $\\alpha^{(1)}=\\alpha^{1}$\\label{alpha^()} and \n\\[\\alpha^{(n+1)}=\\alpha^{(n)}\\wedge\\alpha^{n+1}\\]\nfor $n\\in\\mathbb{N}\\setminus\\{0\\}$. Again for simplicity, $\\alpha^{1}$ may be denoted by $\\alpha^{\\circ}$\\label{alphacirc}, since this formula will play a role equivalent to $\\circ\\alpha$'s role in, for example, $\\mbC$.\n\nThe Hilbert calculus for the positive fragment of intuitionistic logic\\index{Logic, Intuitionistic} is composed of the following axiom schemata\n\\begin{enumerate}\n\\item[\\textbf{Ax\\: 1}] $\\alpha\\rightarrow(\\beta\\rightarrow\\alpha)$;\n\\item[\\textbf{Ax\\: 2}] $\\big(\\alpha\\rightarrow (\\beta\\rightarrow \\gamma)\\big)\\rightarrow\\big((\\alpha\\rightarrow\\beta)\\rightarrow(\\alpha\\rightarrow\\gamma)\\big)$;\n\\item[\\textbf{Ax\\: 3}] $\\alpha\\rightarrow\\big(\\beta\\rightarrow(\\alpha\\wedge\\beta)\\big)$;\n\\item[\\textbf{Ax\\: 4}] $(\\alpha\\wedge\\beta)\\rightarrow \\alpha$;\n\\item[\\textbf{Ax\\: 5}] $(\\alpha\\wedge\\beta)\\rightarrow \\beta$;\n\\item[\\textbf{Ax\\: 6}] $\\alpha\\rightarrow(\\alpha\\vee\\beta)$;\n\\item[\\textbf{Ax\\: 7}] $\\beta\\rightarrow(\\alpha\\vee\\beta)$;\n\\item[\\textbf{Ax\\: 8}] $(\\alpha\\rightarrow\\gamma)\\rightarrow\\Big((\\beta\\rightarrow\\gamma)\\rightarrow \\big((\\alpha\\vee\\beta)\\rightarrow\\gamma\\big)\\Big)$;\n\\end{enumerate}\nplus Modus Ponens as inference rule\n\\[\\frac{\\alpha\\quad\\alpha\\rightarrow\\beta}{\\beta}.\\]\n\nWe define the da costa's system $C_{\\omega}$\\label{Comega}, over $\\Sigma_{\\textbf{C}}$, by adding to the Hilbert calculus for the positive fragment of intuitionistic logic the axiom schemata\n\\begin{enumerate}\n\\item[\\textbf{Ax\\: 10}] $\\alpha\\vee\\neg\\alpha$;\n\\item[\\textbf{cf}] $\\neg\\neg\\alpha\\rightarrow\\alpha$.\n\\end{enumerate}\n\n\\begin{definition}\nFor $n\\in\\mathbb{N}\\setminus\\{0\\}$, we define the da Costa's\\index{Hierarchy, da Costa's} system $C_{n}$\\label{Cn}, over the signature $\\Sigma_{\\textbf{C}}$, as the logic obtained from $C_{\\omega}$ by addition of the axiom schemata\\label{bcn}\\label{phashn}\n\\[\\tag{$\\textbf{bc}_{n}$}\\alpha^{(n)}\\rightarrow\\big(\\alpha\\rightarrow(\\neg\\alpha\\rightarrow\\beta)\\big);\\]\n\\[\\tag{$\\textbf{p}\\vee_{n}$}(\\alpha^{(n)}\\wedge\\beta^{(n)})\\rightarrow(\\alpha\\vee\\beta)^{(n)};\\]\n\\[\\tag{$\\textbf{p}\\wedge_{n}$}(\\alpha^{(n)}\\wedge\\beta^{(n)})\\rightarrow(\\alpha\\wedge\\beta)^{(n)};\\]\n\\[\\tag{$\\textbf{p}\\rightarrow_{n}$}(\\alpha^{(n)}\\wedge\\beta^{(n)})\\rightarrow(\\alpha\\rightarrow\\beta)^{(n)}.\\]\n\\end{definition}\n\nThese systems were originally developed by Newton da Costa, in his seminal work \\cite{Costa3}; originally, the axiom $\\textbf{bc}_{n}$ was presented as\n\\[\\alpha^{(n)}\\rightarrow\\Big((\\beta\\rightarrow\\alpha)\\rightarrow\\big((\\beta\\rightarrow\\neg\\alpha)\\rightarrow\\neg\\beta\\big)\\Big),\\]\nwhich deals with non-contradiction instead of explosivity, but it is clear how both approaches are equivalent. For further comments on this finer distinction, refer back to \\cite{ParLog}.\n\n\\begin{definition}\\label{bival-def}\nA bivaluation for $C_{n}$, also known as a $C_{n}$-bivaluation\\index{Bivaluation for $C_{n}$}, is a function\\\\ $\\mathsf{b}:\\textbf{F}(\\Sigma_{\\textbf{C}}, \\mathcal{V})\\rightarrow\\{0,1\\}$ satisfying:\n\\begin{enumerate}\n\\item[$(B1)$] $\\mathsf{b}(\\alpha\\wedge\\beta)=1$ if and only if $\\mathsf{b}(\\alpha)=1$ and $\\mathsf{b}(\\beta)=1$;\n\\item[$(B2)$] $\\mathsf{b}(\\alpha\\vee\\beta)=1$ if and only if $\\mathsf{b}(\\alpha)=1$ or $\\mathsf{b}(\\beta)=1$;\n\\item[$(B3)$] $\\mathsf{b}(\\alpha\\rightarrow\\beta)=1$ if and only if $\\mathsf{b}(\\alpha)=0$ or $\\mathsf{b}(\\beta)=1$;\n\\item[$(B4)$] $\\mathsf{b}(\\alpha)=0$ implies $\\mathsf{b}(\\neg\\alpha)=1$;\n\\item[$(B5)$] $\\mathsf{b}(\\neg\\neg\\alpha)=1$ implies $\\mathsf{b}(\\alpha)=1$;\n\\item[$(B6)_{n}$] $\\mathsf{b}(\\alpha^{n-1})=\\mathsf{b}(\\neg(\\alpha^{n-1}))$ if and only if $\\mathsf{b}(\\alpha^{n})=0$;\n\\item[$(B7)$] $\\mathsf{b}(\\alpha)=\\mathsf{b}(\\neg\\alpha)$ if and only if $\\mathsf{b}(\\neg(\\alpha^{1}))=1$;\n\\item[$(B8)$] for any $\\#\\in\\{\\vee, \\wedge, \\rightarrow\\}$, $\\mathsf{b}(\\alpha)\\neq\\mathsf{b}(\\neg\\alpha)$ and $\\mathsf{b}(\\beta)\\neq\\mathsf{b}(\\neg\\beta)$ imply, together, that $\\mathsf{b}(\\alpha\\#\\beta)\\neq\\mathsf{b}(\\neg(\\alpha\\#\\beta))$.\n\\end{enumerate}\n\\end{definition}\n\nThe previous notion of bivaluations for $C_{n}$ has its origin in \\cite{Loparic}. If we denote the fact that, for every $C_{n}$-bivaluation $\\mathsf{b}$, $\\mathsf{b}(\\Gamma)\\subseteq\\{1\\}$ implies $\\mathsf{b}(\\varphi)=1$ by $\\Gamma\\vDash_{C_{n}}\\varphi$, it is proved in the same paper that \n\\[\\Gamma\\vdash_{C_{n}}\\varphi\\quad\\text{if and only if}\\quad\\Gamma\\vDash_{C_{n}}\\varphi,\\]\nfor any set of formulas $\\Gamma\\cup\\{\\varphi\\}$ over the signature $\\Sigma_{\\textbf{C}}$.\n\nWe shall provide, in the next sections, a semantical approach to da Costa's hierarchy trough finite restricted Nmatrices.\n\n\\section{$C_{2}$}\n\nWe will start from the simpler case that is $C_{2}$. For $\\mathsf{b}$ a $C_{2}$-bivaluation, $(B6)_{2}$ implies that $\\mathsf{b}(\\alpha^{1})=\\mathsf{b}(\\neg(\\alpha^{1}))$ if and only if $\\mathsf{b}(\\alpha^{2})=0$; so $\\mathsf{b}(\\alpha)=\\mathsf{b}(\\neg\\alpha)=\\mathsf{b}(\\alpha^{1})=1$ implies, from $(B7)$, that $\\mathsf{b}(\\alpha^{2})=0$. Furthermore, $\\mathsf{b}(\\alpha)=0$ leads us to $\\mathsf{b}(\\neg\\neg\\alpha)=0$ from $(B5)$, and so the following four scenarios are possible.\n\n\\begin{figure}[H]\n\\centering\n\\begin{tabular}{|l|c|c|c|c|c|c|r|}\n\\hline\n$\\alpha$ & $\\neg\\alpha$ & $\\alpha\\wedge\\neg\\alpha$ & $\\alpha^{1}$ & $\\neg(\\alpha^{1})$ & $\\alpha^{1}\\wedge\\neg(\\alpha^{1})$ & $\\alpha^{2}$ & $\\alpha^{(2)}$\\\\ \\hline\n\\multirow{3}{*}{$1$} & \\multirow{2}{*}{$1$} & \\multirow{2}{*}{$1$} & $1$ & $1$ & $1$ & $0$ & $0$\\\\\\cline{4-8}\n& & & $0$ & $1$ & $0$ & $1$ & $0$\\\\\\cline{2-8}\n& $0$ & $0$ & $1$ & $0$ & $0$ & $1$ & $1$\\\\\\hline\n$0$ & $1$ & $0$ & $1$ & $0$ & $0$ & $1$ & $1$\\\\\\hline\n\\end{tabular}\n\\caption*{Table for the scenarios on $C_{2}$}\n\\end{figure}\n\nNow, for a formula $\\alpha$ and a $C_{2}$-bivaluation, we would like to consider the triple $(\\mathsf{b}(\\alpha), \\mathsf{b}(\\neg\\alpha), \\mathsf{b}(\\alpha^{1}))$ in $\\{0,1\\}^{3}$; from the previous table, we see that there are only $4$ possibilities, that we list below:\n\\[T_{2}=(1,0,1),\\quad t^{2}_{0}=(1,1,0),\\quad t^{2}_{1}=(1,1,1,)\\quad\\text{and}\\quad F_{2}=(0,1,1).\\]\nWe call the set of those elements $B_{2}$, and it is clear that \n\\[B_{2}=\\{z\\in \\{0,1\\}^{3}\\ :\\ z_{1}\\vee z_{2}=1\\quad\\text{and}\\quad (z_{1}\\wedge z_{2})\\vee z_{3}=1\\},\\]\nwhere $z_{i}$ will denote the $i$th coordinate of an element $z\\in\\{0,1\\}^{3}$. We then define the $\\Sigma_{\\textbf{C}}$-multialgebra $\\mathcal{A}_{C_{2}}$ with universe $B_{2}$ and operations given by\n\\[\\tilde{\\neg}z=\\{w\\in B_{2}\\ :\\ w_{1}=z_{2}\\quad\\text{and}\\quad w_{2}\\leq z_{1}\\}\\]\nand\n\\[\\text{for}\\quad \\#\\in\\{\\vee, \\wedge, \\rightarrow\\},\\quad z\\tilde{\\#}w=\\begin{cases*}\n\\{u\\in B_{2}\\ :\\ u_{1}=z_{1}\\# w_{1}\\}\\cap Boo_{2} & if $z, w\\in Boo_{2}$\\\\\n\\{u\\in B_{2}\\ :\\ u_{1}=z_{1}\\# w_{1}\\} & otherwise\n\\end{cases*},\\]\nwhere: we will denote an operation $\\sigma_{\\mathcal{A}_{C_{2}}}$ on $\\mathcal{A}_{C_{2}}$ simply by $\\tilde{\\sigma}$; the operations $\\#\\in\\{\\vee, \\wedge, \\rightarrow\\}$ are defined as usual in $\\{0,1\\}$; and $Boo_{2}=\\{F_{2}, T_{2}\\}$ is the set of classically-behaving elements.\n\nOne could argue that a more natural definition would be that $z\\tilde{\\#}w$ always equals $\\{u\\in B_{2}\\ :\\ u_{1}=z_{1}\\# w_{1}\\}$, but that would be problematic. Condition $(B8)$ for being a bivaluation for $C_{n}$ states that $\\mathsf{b}(\\alpha)\\neq\\mathsf{b}(\\neg\\alpha)$ and $\\mathsf{b}(\\beta\\neq\\mathsf{b}(\\neg\\beta)$ imply, together, that $\\mathsf{b}(\\alpha\\#\\beta)\\neq\\mathsf{b}(\\neg(\\alpha\\#\\beta))$; and one easily sees that the elements $(\\mathsf{b}(\\alpha), \\mathsf{b}(\\neg\\alpha), \\mathsf{b}(\\alpha^{1}))$ of $B_{2}$ satisfying $\\mathsf{b}(\\alpha)\\neq\\mathsf{b}(\\neg\\alpha)$ are precisely $F_{2}$ and $T_{2}$. This all means that $(B8)$ translates to demanding that, if $x,y\\in Boo_{2}$, then one must necessarily have $x\\tilde{\\#}y\\in Boo_{2}$. Notice that this also justifies the nomenclature of ``classical'' for the elements $F_{2}$ and $T_{2}$: they correspond to formulas such that their negation, and the formula itself, are not simultaneously true.\n\nIf we denote the set $\\{t_{1}^{2}, t_{0}^{2}, T_{2}\\}$ by $D_{2}$, the tables for $\\mathcal{A}_{C_{2}}$ are as below. We will also use the notation $\\{t_{0}^{2}, t_{1}^{2}\\}=I_{2}$ when necessary.\n\\begin{figure}[H]\n\\centering\n\\begin{minipage}[t]{4cm}\n\\centering\n\\begin{tabular}{|l|c|c|c|r|}\n\\hline\n$\\vee$ & $F_{2}$ & $t_{1}^{2}$ & $t_{0}^{2}$ & $T_{2}$ \\\\ \\hline\n$F_{2}$ & $\\{F_{2}\\}$ & $D_{2}$ & $D_{2}$ & $\\{T_{2}\\}$ \\\\ \\hline\n$t_{1}^{2}$ & $D_{2}$ & $D_{2}$ & $D_{2}$ & $D_{2}$ \\\\ \\hline\n$t_{0}^{2}$ & $D_{2}$ & $D_{2}$ & $D_{2}$ & $D_{2}$ \\\\ \\hline\n$T_{2}$ & $\\{T_{2}\\}$ & $D_{2}$ & $D_{2}$ & $\\{T_{2}\\}$ \\\\ \\hline\n\\end{tabular}\n\\caption*{Table for Disjunction}\n\\end{minipage}\n\\hspace{3cm}\n\\centering\n\\begin{minipage}[t]{4cm}\n\\centering\n\\begin{tabular}{|l|c|c|c|r|}\n\\hline\n$\\wedge$ & $F_{2}$ & $t_{1}^{2}$ & $t_{0}^{2}$ & $T_{2}$ \\\\ \\hline\n$F_{2}$ & $\\{F_{2}\\}$ & $\\{F_{2}\\}$ & $\\{F_{2}\\}$ & $\\{F_{2}\\}$ \\\\ \\hline\n$t_{1}^{2}$ & $\\{F_{2}\\}$ & $D_{2}$ & $D_{2}$ & $D_{2}$ \\\\ \\hline\n$t_{0}^{2}$ & $\\{F_{2}\\}$ & $D_{2}$ & $D_{2}$ & $D_{2}$ \\\\ \\hline\n$T_{2}$ & $\\{F_{2}\\}$ & $D_{2}$ & $D_{2}$ & $\\{T_{2}\\}$ \\\\ \\hline\n\\end{tabular}\n\\caption*{Table for Conjunction}\n\\end{minipage}\n\\end{figure}\n\n\\begin{figure}[H]\n\\centering\n\\begin{minipage}[t]{4cm}\n\\centering\n\\begin{tabular}{|l|r|}\n\\hline\n & $\\neg$ \\\\ \\hline\n$F_{2}$ & $\\{T_{2}\\}$\\\\ \\hline\n$t_{1}^{2}$ & $D_{2}$\\\\ \\hline\n$t_{0}^{2}$ & $D_{2}$\\\\ \\hline\n$T_{2}$ & $\\{F_{2}\\}$ \\\\ \\hline\n\\end{tabular}\n\\caption*{Table for negation}\n\\end{minipage}\n\\hspace{3cm}\n\\centering\n\\begin{minipage}[t]{5cm}\n\\centering\n\\begin{tabular}{|l|c|c|c|r|}\n\\hline\n$\\rightarrow$ & $F_{2}$ & $t_{1}^{2}$ & $t_{0}^{2}$ & $T_{2}$ \\\\ \\hline\n$F_{2}$ & $\\{T_{2}\\}$ & $D_{2}$ & $D_{2}$ & $\\{T_{2}\\}$ \\\\ \\hline\n$t_{1}^{2}$ & $\\{F_{2}\\}$ & $D_{2}$ & $D_{2}$ & $D_{2}$ \\\\ \\hline\n$t_{0}^{2}$ & $\\{F_{2}\\}$ & $D_{2}$ & $D_{2}$ & $D_{2}$ \\\\ \\hline\n$T_{2}$ & $\\{F_{2}\\}$ & $D_{2}$ & $D_{2}$ & $\\{T_{2}\\}$ \\\\ \\hline\n\\end{tabular}\n\\caption*{Table for Implication}\n\\end{minipage}\n\\end{figure}\n\nIf we consider the Nmatrix $\\mathcal{M}_{C_{2}}=(\\mathcal{A}_{C_{2}}, D_{2})$, we cannot hope to characterize $C_{2}$ with it given that $C_{2}$ is not characterizable by a single finite Nmatrix (\\cite{Avron}). What we will do instead, and which will be a successful endeavor, is to restrict the set of valuations for this Nmatrix, creating an RNmatrix, in order to characterize $C_{2}$.\n\n\\begin{definition}\\label{FC2}\nLet $\\mathcal{F}_{C_{2}}$ be the set of homomorphisms $\\nu:\\textbf{F}(\\Sigma_{\\textbf{C}}, \\mathcal{V})\\rightarrow \\mathcal{A}_{C_{2}}$ (which are called valuations over $\\mathcal{A}_{C_{2}}$) such that:\n\\begin{enumerate}\n\\item if $\\nu(\\alpha)=t_{0}^{2}$, then $\\nu(\\alpha\\wedge\\neg\\alpha)=T_{2}$;\n\\item if $\\nu(\\alpha)=t_{1}^{2}$, then $\\nu(\\alpha\\wedge\\neg\\alpha)\\in I_{2}$ and $\\nu(\\alpha^{\\circ})=t_{0}^{2}$.\n\\end{enumerate}\nWe will denote the restricted Nmatrix $(\\mathcal{A}_{C_{2}}, D_{2}, \\mathcal{F}_{C_{2}})$ by $\\mathcal{RM}_{C_{2}}$.\n\\end{definition}\n\nSuppose $\\nu\\in\\mathcal{F}_{C_{2}}$: notice that, if $\\nu(\\alpha)=t_{0}^{2}$, then $\\nu(\\alpha\\wedge\\neg\\alpha)=T_{2}$ and so $\\nu(\\alpha^{\\circ})=F_{2}$; if $\\nu(\\alpha)=t_{1}^{2}$, $\\nu(\\alpha^{\\circ})=t_{0}^{2}$ and therefore $\\nu(\\alpha^{\\circ\\circ})=F_{2}$. Let $\\bot_{\\alpha}$ denote $(\\alpha\\wedge\\neg\\alpha)\\wedge\\alpha^{(2)}$ and ${\\sim}\\alpha$ denote $\\alpha\\rightarrow\\bot_{\\alpha}$ (the strong negation definable in $C_{2}$), and we arrive at the following table, where an asterisk signifies that a certain value would be different if the table were constructed using the Nmatrix $(\\mathcal{A}_{C_{2}}, D_{2})$.\n\n\\begin{figure}[H]\n\\centering\n\\begin{tabular}{|l|c|c|c|c|c|c|c|c|r|}\n\\hline\n$\\alpha$ & $\\neg\\alpha$ & $\\alpha\\wedge\\neg\\alpha$ & $\\alpha^{\\circ}$ & $\\neg(\\alpha^{\\circ})$ & $\\alpha^{\\circ}\\wedge\\neg(\\alpha^{\\circ})$ & $\\alpha^{\\circ\\circ}$ & $\\alpha^{(2)}$ & $\\bot_{\\alpha}$ & ${\\sim}\\alpha$ \\\\ \\hline\n$T_{2}$ & $F_{2}$ & $F_{2}$ & $T_{2}$ & $F_{2}$ & $F_{2}$ & $T_{2}$ & $T_{2}$ & $F_{2}$ & $F_{2}$\\\\ \\hline\n$t_{0}^{2}$ & $D_{2}$ & $T_{2}^{*}$ & $F_{2}$ & $T_{2}$ & $F_{2}$ & $T_{2}$ & $F_{2}$ & $F_{2}$ & $F_{2}$\\\\ \\hline\n$t_{1}^{2}$ & $D_{2}$ & $I_{2}^{*}$ & $t_{0}^{2 *}$ & $D_{2}$ & $T_{2}^{*}$ & $F_{2}$ & $F_{2}$ & $F_{2}$ & $F_{2}$ \\\\ \\hline\n$F_{2}$ & $T_{2}$ & $F_{2}$ & $T_{2}$ & $F_{2}$ & $F_{2}$ & $T_{2}$ & $T_{2}$ & $F_{2}$ & $T_{2}$ \\\\ \\hline\n\\end{tabular}\n\\caption*{Table for the scenarios in $\\mathcal{RM}_{C_{2}}$}\n\\end{figure}\n\n\\begin{proposition}\nFor a homomorphism $\\nu$ in $\\mathcal{F}_{C_{2}}$ and an endomorphism $\\sigma:\\textbf{F}(\\Sigma_{\\textbf{C}}, \\mathcal{V})\\rightarrow\\textbf{F}(\\Sigma_{\\textbf{C}}, \\mathcal{V})$, $\\nu\\circ\\sigma\\in \\mathcal{F}_{C_{2}}$.\n\\end{proposition}\n\n\\begin{proof}\nOf course $\\nu\\circ\\sigma:\\textbf{F}(\\Sigma_{\\textbf{C}}, \\mathcal{V})\\rightarrow\\mathcal{A}_{C_{2}}$ remains a homomorphism.\n\\begin{enumerate}\n\\item If $\\nu\\circ\\sigma(\\alpha)=t_{0}^{2}$, given $\\nu$ is in $\\mathcal{F}_{C_{2}}$ we derive that $\\nu(\\sigma(\\alpha)\\wedge\\neg\\sigma(\\alpha))=T_{2}$; since $\\sigma$ is an endomorphism of $\\textbf{F}(\\Sigma_{\\textbf{C}}, \\mathcal{V})$, $\\sigma(\\alpha)\\wedge\\neg\\sigma(\\alpha)=\\sigma(\\alpha\\wedge\\neg\\alpha)$, and so $\\nu\\circ\\sigma(\\alpha\\wedge\\neg\\alpha)=T_{2}$.\n\n\\item If $\\nu\\circ\\sigma(\\alpha)=t_{1}^{2}$, since $\\nu\\in\\mathcal{F}_{C_{2}}$ we get $\\nu(\\sigma(\\alpha)\\wedge\\neg\\sigma(\\alpha))\\in I_{2}$ and \n\\[\\nu(\\sigma(\\alpha)^{\\circ})=\\nu\\Big(\\neg(\\sigma(\\alpha)\\wedge\\neg\\sigma(\\alpha))\\Big)=t_{0}^{2};\\]\ngiven $\\sigma(\\alpha)\\wedge\\neg\\sigma(\\alpha)=\\sigma(\\alpha\\wedge\\neg\\alpha)$ and $\\neg(\\sigma(\\alpha)\\wedge\\neg\\sigma(\\alpha))=\\sigma(\\neg(\\alpha\\wedge\\neg\\alpha))$, we obtain that $\\nu\\circ\\sigma(\\alpha\\wedge\\neg\\alpha)\\in I_{2}$ and $\\nu\\circ\\sigma(\\neg(\\alpha\\wedge\\neg\\alpha))=\\nu\\circ\\sigma(\\alpha^{\\circ})=t_{0}^{2}$, what finishes the proof.\n\\end{enumerate}\n\\end{proof}\n\nThe previous proposition implies $\\mathcal{RM}_{C_{2}}$ is structural. \n\nThe following technical lemmas are necessary in order to prove the desired Theorem \\ref{Char C2}.\n\n\\begin{lemma}\\label{Sound C2}\nFor $\\nu$ a valuation in $\\mathcal{F}_{C_{2}}$, the mapping $\\mathsf{b}: F(\\Sigma_{\\textbf{C}}, \\mathcal{V})\\rightarrow \\{0,1\\}$, such that $\\mathsf{b}(\\alpha)=1$ if and only if $\\nu(\\alpha)\\in D_{2}$, is a $C_{2}$-bivaluation.\n\\end{lemma}\n\n\\begin{proof}\nLet us see that $\\mathsf{b}$ satisfies the clauses of Definition \\ref{bival-def} for $n=2$. Clauses $(B1)$ through $(B3)$ are clearly satisfied: since \n\\[\\mathsf{b}(\\alpha \\# \\beta) = \\nu(\\alpha \\# \\beta)_1 = \\nu(\\alpha)_1 \\# \\nu(\\beta)_1\\]\nfor $\\# \\in \\{\\vee, \\wedge, \\rightarrow\\}$. \n\nOn the other hand, $\\mathsf{b}(\\neg\\alpha) = \\nu(\\neg\\alpha)_1 = \\nu(\\alpha)_2$, hence $\\mathsf{b}(\\alpha) \\vee \\mathsf{b}(\\neg\\alpha)=1$, by definition of $B_{2}$. From this, clause $(B2)$ is satisfied. Since $\\mathsf{b}(\\neg\\neg\\alpha)=\\nu(\\neg\\neg\\alpha)_1=\\nu(\\neg\\alpha)_2\\leq \\nu(\\alpha)_1 = \\mathsf{b}(\\alpha)$ it follows that $\\mathsf{b}$ satisfies $(B5)$. \n\nConcerning $(B6)_2$, suppose that $\\mathsf{b}(\\alpha^\\circ)=\\mathsf{b}(\\neg(\\alpha^\\circ))=1$: this means that $\\nu(\\alpha^\\circ)_1=\\nu(\\neg(\\alpha^\\circ))_1=\\nu(\\alpha^\\circ)_2=1$. That is, $\\nu(\\alpha^\\circ) \\in \\{t^2_0,t^2_1\\}$. From the possible scenarios in $\\mathcal{RM}_{C_{2}}$ it follows that $\\nu(\\alpha^\\circ) = t^2_1$ and so $\\nu(\\alpha^2)=F$, what implies that $\\mathsf{b}(\\alpha^2)=\\nu(\\alpha^2)_1=0$. Conversely, $\\mathsf{b}(\\alpha^2)=\\nu(\\alpha^2)_1=0$ means that $\\nu(\\alpha^2)=F$, which implies that $\\nu(\\alpha^\\circ) = t^2_1$, and so $\\mathsf{b}(\\alpha^\\circ)=\\mathsf{b}(\\neg(\\alpha^\\circ))=1$. From this, clause $(B6)_2$ is satisfied. \n\nNow, $\\mathsf{b}(\\alpha)=\\mathsf{b}(\\neg\\alpha)$ if and only if $\\nu(\\alpha)_1 = \\nu(\\alpha)_2=1$, what in turn happens if and only if $\\nu(\\alpha) \\in \\{t^2_0,t^2_1\\}$. This last condition is equivalent to the fact that $\\nu(\\neg(\\alpha^\\circ)) \\in D_2$, by the possible scenarios in $\\mathcal{RM}_{C_{2}}$, which is equivalent to $\\mathsf{b}(\\neg(\\alpha^\\circ))=1$. Then, $(B7)$ is satisfied. \n\nFinally, suppose that $\\mathsf{b}(\\alpha) \\neq \\mathsf{b}(\\neg\\alpha)$ and $v(\\beta) \\neq \\mathsf{b}(\\neg\\beta)$. This means that $\\nu(\\alpha),\\nu(\\beta) \\in \\{T_2,F_2\\}$ and so $\\nu(\\alpha\\# \\beta) \\in \\{T_2,F_2\\}$, by definition of $\\mathcal{A}_{C_2}$. Hence clause $(B8)$ is fulfilled, and the proof is complete.\n\\end{proof}\n\n\\begin{lemma}\\label{Comp C2}\nFor $\\mathsf{b}$ a $C_{2}$-bivaluation, the mapping $\\nu:\\textbf{F}(\\Sigma_{\\textbf{C}}, \\mathcal{V})\\rightarrow \\mathcal{A}_{C_{2}}$, such that $\\nu(\\alpha)=(\\mathsf{b}(\\alpha), \\mathsf{b}(\\neg\\alpha), \\mathsf{b}(\\alpha^{\\circ}))$, is a valuation which lies in $\\mathcal{F}_{C_{2}}$ and satisfies that $\\mathsf{b}(\\alpha)=1$ if, and only if, $\\nu(\\alpha)\\in D_{2}$.\n\\end{lemma}\n\n\\begin{proof}\nBy definition, $\\nu(\\neg\\alpha)=(\\mathsf{b}(\\neg\\alpha),\\mathsf{b}(\\neg\\neg\\alpha),\\mathsf{b}((\\neg\\alpha)^\\circ))$, and clearly this belongs to $\\tilde{\\neg} \\nu(\\alpha)$ according to the definition of $\\tilde{\\neg}$, by the property $(B5)$ of $\\mathsf{b}$. Analogously, by definition of $\\tilde{\\#}$ it follows that $\\nu(\\alpha\\# \\beta) \\in \\nu(\\alpha) \\tilde{\\#} \\nu(\\beta)$, with use of the properties $(B1)$ through $(B3)$ of $\\mathsf{b}$. This shows that $\\nu$ is a valuation for $\\mathcal{A}_{C_2}$. It remains to be shown that $\\nu$ satisfies conditions $1$ and $2$ of Definition \\ref{FC2}.\n\nRegarding the first of these conditions, assume that $\\nu(\\alpha) = t^2_0$. This means that $\\mathsf{b}(\\alpha)=\\mathsf{b}(\\neg\\alpha)=1$ and $\\mathsf{b}(\\alpha^\\circ)=0$. Let $\\beta=\\alpha\\wedge\\neg\\alpha$ (thus $\\neg\\beta=\\alpha^\\circ$). By $(B1)$, $\\mathsf{b}(\\beta)=1$. Since, by hypothesis, $\\mathsf{b}(\\beta) \\neq \\mathsf{b}(\\neg\\beta)$, it follows that $\\mathsf{b}(\\neg(\\beta^\\circ))=0$, by~$( B7)$, and so $\\mathsf{b}(\\beta^\\circ)=1$, by~$(B4)$. This shows that \n\\[\\nu(\\alpha\\wedge\\neg\\alpha) = (1,0,1)=T_2,\\]\nas we wished to prove. \n\nRegarding the second condition, suppose that $\\nu(\\alpha) = t^2_1$. Then, $\\mathsf{b}(\\alpha)=\\mathsf{b}(\\neg\\alpha)=\\mathsf{b}(\\alpha^\\circ)=1$. Consider again $\\beta=\\alpha \\land \\neg\\alpha$. From this, $\\mathsf{b}(\\beta)=\\mathsf{b}(\\neg\\beta)=1$ and so \n\\[\\nu(\\alpha\\wedge\\neg\\alpha) = (\\mathsf{b}(\\beta), \\mathsf{b}(\\neg\\beta), \\mathsf{b}(\\beta^\\circ)) \\in \\{t^2_1,t^2_1\\}.\\]\nSince $\\mathsf{b}(\\alpha)=\\mathsf{b}(\\neg\\alpha)$ it follows by $(B7)$ that $\\mathsf{b}(\\neg(\\alpha^\\circ))=1$. Hence, since $\\mathsf{b}(\\alpha^1)=\\mathsf{b}(\\neg(\\alpha^1))$, we infer that $\\mathsf{b}(\\alpha^2)=0$, by $(B6)_2$. That is, $\\nu(\\alpha^\\circ) = (1,1,0)=t^2_0$. This, of course, finishes proving that $\\nu$ is in $\\mathcal{F}_{C_{2}}$.\n\\end{proof}\n\n\\begin{theorem}\\label{Char C2}\nGiven formulas $\\Gamma\\cup\\{\\varphi\\}$ of $C_{2}$, $\\Gamma\\vdash_{C_{2}}\\varphi$ if, and only if, $\\Gamma\\vDash_{\\mathcal{RM}_{C_{2}}}\\varphi$.\n\\end{theorem}\n\n\\begin{proof}\nSuppose first that $\\Gamma \\vdash_{C_{2}} \\varphi$, and let $\\nu$ be a valuation in $\\mathcal{F}_{C_2}$ such that $\\nu(\\gamma) \\in D_{2}$ for every $\\gamma \\in \\Gamma$. By Lemma \\ref{Sound C2}, $\\mathsf{b}$ defined as $\\mathsf{b}(\\alpha)=\\nu(\\alpha)_1$, for every $\\alpha$, is a $C_2$-bivaluation such that, by definition, $\\mathsf{b}(\\gamma) =1$ for every $\\gamma \\in \\Gamma$. By hypothesis and by soundness of $C_2$ with respect to bivaluations, it follows that $\\mathsf{b}(\\varphi)=1$. That is, $\\nu(\\varphi) \\in D_2$. This shows that $\\Gamma\\vDash_{\\mathcal{RM}_{C_{2}}}\\varphi$.\n\nNow assume that $\\Gamma\\vDash_{\\mathcal{RM}_{C_{2}}}\\varphi$, and let $\\mathsf{b}$ be a $C_2$-bivaluation such that $\\mathsf{b}(\\gamma) =1$ for every $\\gamma \\in \\Gamma$. By Lemma \\ref{Comp C2}, the function $\\nu$ defined as $\\nu(\\alpha)=(\\mathsf{b}(\\alpha),\\mathsf{b}(\\neg\\alpha),\\mathsf{b}(\\alpha^\\circ))$, for every $\\alpha$, is a valuation in $\\mathcal{F}_{C_2}$ such that, by definition, $\\nu(\\gamma) \\in D_2$ for every $\\gamma \\in \\Gamma$. By hypothesis, $\\nu(\\varphi) \\in D_2$, which means that $\\mathsf{b}(\\varphi)=1$. By completeness of $C_2$ with respect to bivaluations, it follows that $\\Gamma \\vdash_{C_2} \\varphi$.\n\\end{proof}\n\n\n\n\n\\section{The general case}\\label{The general case}\n\n\\begin{lemma}\\label{Values of i-consistency}\nGiven a $C_{n}$-bivaluation $\\mathsf{b}$, if $\\mathsf{b}(\\alpha^{i})=1$ and $\\mathsf{b}(\\neg(\\alpha^{i}))=0$ for some $i\\in\\mathbb{N}$, then $\\mathsf{b}(\\alpha^{j})=1$ for all $j\\geq i$.\n\\end{lemma}\n\n\\begin{proof}\nWe prove that, for all $j\\geq i$, $\\mathsf{b}(\\alpha^{j})=1$ and $\\mathsf{b}(\\neg(\\alpha^{j}))=0$ by induction, being the base case done. So, suppose $\\mathsf{b}(\\alpha^{j})=1$ and $\\mathsf{b}(\\neg(\\alpha^{j}))=0$: we have $\\mathsf{b}(\\alpha^{j}\\wedge\\neg(\\alpha^{j}))=0$ from $(B1)$, and $\\mathsf{b}(\\alpha^{j+1})=\\mathsf{b}(\\neg(\\alpha^{j}\\wedge\\neg(\\alpha^{j})))=1$ from $(B4)$; and from $(B7)$, $\\mathsf{b}(\\neg(\\alpha^{j+1}))=\\mathsf{b}(\\neg((\\alpha^{j})^{1}))=0$, since $\\mathsf{b}(\\alpha^{j})\\neq\\mathsf{b}(\\neg(\\alpha^{j}))$, what finishes the proof.\n\\end{proof}\n\n\\begin{proposition}\nIf $\\mathsf{b}$ is a $C_{n}$-bivaluation, at most one of the elements $\\mathsf{b}(\\alpha)$, $\\mathsf{b}(\\neg\\alpha)$, $\\mathsf{b}(\\alpha^{1}), \\dotsc ,\n$\\\\$\\mathsf{b}(\\alpha^{n-1})$ equals $0$.\n\\end{proposition}\n\n\\begin{proof}\nSuppose $\\mathsf{b}(\\alpha)=0$: from condition $(B4)$, this means $\\mathsf{b}(\\neg\\alpha)=1$. Then $\\mathsf{b}(\\alpha\\wedge\\neg\\alpha)=0$ (from $(B1)$), meaning $\\mathsf{b}(\\alpha^{1})=\\mathsf{b}(\\neg(\\alpha\\wedge\\neg\\alpha))=1$, again from $(B4)$; but notice, furthermore, that, from the converse of $(B5)$, $\\mathsf{b}(\\alpha\\wedge\\neg\\alpha)=0$ implies $\\mathsf{b}(\\neg(\\alpha^{1}))=\\mathsf{b}(\\neg\\neg(\\alpha\\wedge\\neg\\alpha))=0$. From Lemma \\ref{Values of i-consistency}, we find $\\mathsf{b}(\\alpha^{2})=\\cdots=\\mathsf{b}(\\alpha^{n-1})=1$, what finishes proving that, if $\\mathsf{b}(\\alpha)=0$, then $\\mathsf{b}(\\neg\\alpha)=\\mathsf{b}(\\alpha^{1})=\\cdots=\\mathsf{b}(\\alpha^{n-1})=1$.\n\nNow, suppose $\\mathsf{b}(\\neg\\alpha)=0$: we already have $\\mathsf{b}(\\alpha)=1$, since otherwise one could derive $\\mathsf{b}(\\neg\\alpha)=1$ from the previous remarks. So $\\mathsf{b}(\\alpha\\wedge\\neg\\alpha)=0$ and therefore $\\mathsf{b}(\\alpha^{1})=1$. Again by the converse of $(B5)$, one finds $\\mathsf{b}(\\neg(\\alpha^{1}))=0$, and again from Lemma \\ref{Values of i-consistency}, we obtain $\\mathsf{b}(\\alpha^{1})=\\cdots=\\mathsf{b}(\\alpha^{n-1})=1$.\n\nFinally, suppose that for some $1\\leq i n$, for any formula $\\alpha$, since they must always take the values, respectively, $1$ and $0$.} We start by looking at $\\mathsf{b}(\\neg\\neg\\alpha)$, which we make equal to\n\\begin{enumerate}\n\\item $\\mathsf{b}(\\gamma)\\wedge\\mathsf{b}(\\neg\\gamma)$, if $\\neg\\alpha=\\gamma^{1}$ (thus validating $(V5)$);\n\\item any value between $\\mathsf{b}(\\alpha)$ and ${\\sim}\\mathsf{b}(\\neg\\alpha)$, both included, if the previous case does not apply (validating $(V2)$ and $(V3)$).\n\\end{enumerate}\nRegarding $\\mathsf{b}((\\neg\\alpha)^{i})$ and $\\mathsf{b}(\\neg(\\neg\\alpha)^{i})$, for $1\\leq i\\leq n-1$, there isn't a lot we must do: supposing we have defined $\\mathsf{b}((\\neg\\alpha)^{i-1})$ and $\\mathsf{b}(\\neg(\\neg\\alpha)^{i-1})$, it is sufficient to demand that $\\mathsf{b}((\\neg\\alpha)^{i})$ is greater or equal to \n\\[{\\sim}[\\mathsf{b}((\\neg\\alpha)^{i-1}\\wedge\\neg(\\neg\\alpha)^{i-1})],\\]\nin order to satisfy clause $(V2)$, and that $\\mathsf{b}(\\neg(\\neg\\alpha)^{i})$ equals $\\mathsf{b}((\\neg\\alpha)^{i-1})\\wedge\\mathsf{b}(\\neg(\\neg\\alpha)^{i-1})$, in order to validate $(V5)$ (unless $\\neg\\alpha=\\gamma^{k}$ for some $1\\leq k\\leq n-1$ such that $k+i=n$, in which case $\\mathsf{b}((\\neg\\alpha)^{i})=\\mathsf{b}(\\gamma^{n})$ is defined according to $(V4)_{n}$). Of course, $\\mathsf{b}((\\neg\\alpha)^{n})$ and $\\mathsf{b}(\\neg(\\neg\\alpha)^{n})$ are given values according to clauses, respectively, $(V4)_{n}$ and $(V5)$, being expressed in terms of the already defined $\\mathsf{b}((\\neg\\alpha)^{n-1})$ and $\\mathsf{b}(\\neg(\\neg\\alpha)^{n-1})$.\n\n\nDifficulties finally arise when defining $\\mathsf{b}(\\neg(\\alpha\\#\\beta))$, $\\mathsf{b}((\\alpha\\#\\beta)^{i})$ and $\\mathsf{b}(\\neg(\\alpha\\#\\beta)^{i})$, because we must be mindful of clause $(V6)_{n}$: denote by $a_{0}$ the value $\\mathsf{b}(\\alpha_{0})=\\mathsf{b}(\\alpha)\\#\\mathsf{b}(\\beta)$, and by $a_{\\#}$ the value $\\mathsf{b}(\\alpha^{(n)})\\wedge\\mathsf{b}(\\beta^{(n)})=\\bigwedge_{j=1}^{n}\\mathsf{b}(\\alpha^{j})\\wedge\\mathsf{b}(\\beta^{j})$ (all of $\\mathsf{b}(\\alpha^{j})$ and $\\mathsf{b}(\\beta^{j})$, by hypothesis, already defined).\n\n\\begin{enumerate}[align=left]\n\\item[\\underline{$\\mathsf{b}(\\neg(\\alpha\\#\\beta))$} - ] The value for $\\mathsf{b}(\\neg(\\alpha\\#\\beta))$ will be $a_{\\neg}$: on the off chance that $\\#=\\wedge$, $\\alpha=\\gamma^{n-1}$ and $\\beta=\\neg\\gamma^{n-1}$, we make $a_{\\neg}$ equal to ${\\sim}a_{0}={\\sim}[\\mathsf{b}(\\gamma^{n-1})\\wedge\\mathsf{b}(\\neg(\\gamma^{n-1}))]$ (as required by $(V4)_{n}$); otherwise, $a_{\\neg}$ can be any value such that \n\\[a_{\\neg}\\geq {\\sim}a_{0}\\quad\\text{and}\\quad a_{\\neg}\\wedge a_{0}\\leq{\\sim}a_{\\#}\\]\n(this is possible, one example of such an element being $a_{\\neg}={\\sim}a_{0}$); this last case of course implies $\\mathsf{b}(\\neg(\\alpha\\#\\beta))=a_{\\neg}\\geq{\\sim}a_{0}={\\sim}\\mathsf{b}(\\alpha\\#\\beta)$, and thus also validates $(V2)$. \n\n\\item[\\underline{$\\mathsf{b}((\\alpha\\#\\beta)^{1})$} - ] Having defined $a_{\\neg}$, we define $\\mathsf{b}((\\alpha\\#\\beta)^{1})$ to be any value $a_{1}$ such that $a_{1}\\geq {\\sim}(a_{0}\\wedge a_{\\neg})$ (and therefore we have\n\\[\\mathsf{b}\\big(\\neg((\\alpha\\#\\beta)\\wedge\\neg(\\alpha\\#\\beta))\\big)=\\mathsf{b}((\\alpha\\#\\beta)^{1})=a_{1}\\geq {\\sim}(a_{0}\\wedge a_{\\neg})={\\sim}\\mathsf{b}\\big((\\alpha\\#\\beta)\\wedge\\neg(\\alpha\\#\\beta)\\big),\\]\nand thus $(V2)$ remains valid) and ${\\sim}(a_{0}\\wedge a_{\\neg})\\wedge a_{1}\\geq a_{\\#}$ (what would still be possible, being enough to choose, for one, $a_{1}={\\sim}(a_{0}\\wedge a_{\\neg})$).\n\nAnd we make $\\mathsf{b}(\\neg(\\alpha\\#\\beta)^{1})=\\mathsf{b}(\\alpha\\#\\beta)\\wedge\\mathsf{b}(\\neg(\\alpha\\#\\beta))=a_{0}\\wedge a_{\\neg}$, thus respecting clause $(V5)$ (and incidentally also $(V2)$ and $(V3)$).\n\n\\item[\\underline{Inductive step} - ] Supposing we have defined $a_{1}=\\mathsf{b}((\\alpha\\#\\beta)^{1})$ trough $a_{k}=\\mathsf{b}((\\alpha\\#\\beta)^{k})$, for $1\\leq k\\leq n-2$, we make $\\mathsf{b}((\\alpha\\#\\beta)^{k+1})$ equal to any value $a_{k+1}$ satisfying \n\\[a_{k+1}\\geq{\\sim}(a_{0}\\wedge a_{\\neg}\\wedge \\bigwedge_{j=1}^{k}a_{j})\\quad\\text{and}\\quad{\\sim}(a_{0}\\wedge a_{\\neg})\\wedge\\bigwedge_{j=2}^{k}a_{j}\\geq a_{\\#};\\]\n\nInductively, we get $\\mathsf{b}(\\neg(\\alpha\\#\\beta)^{k})=a_{0}\\wedge a_{\\neg}\\wedge \\bigwedge_{j=1}^{k-1}a_{j}$, and so: first of all, this means that $\\mathsf{b}((\\alpha\\#\\beta)^{k})\\wedge\\mathsf{b}(\\neg(\\alpha\\#\\beta)^{k})=a_{0}\\wedge a_{\\neg}\\wedge \\bigwedge_{j=1}^{k}a_{j}$ and therefore \n\\[\\mathsf{b}((\\alpha\\#\\beta)^{k+1})=\\mathsf{b}(\\neg((\\alpha\\#\\beta)^{k}\\wedge\\neg(\\alpha\\#\\beta)^{k})\\geq {\\sim}\\big(\\mathsf{b}((\\alpha\\#\\beta)^{k})\\wedge\\mathsf{b}(\\neg(\\alpha\\#\\beta)^{k})\\big),\\]\nmeaning $(V2)$ is respected; second, by making $\\mathsf{b}(\\neg(\\alpha\\#\\beta)^{k+1})=\\mathsf{b}((\\alpha\\#\\beta)^{k})\\wedge \\mathsf{b}(\\neg(\\alpha\\#\\beta)^{k})$, so as to respect $(V5)$, we obtain that $\\mathsf{b}(\\neg(\\alpha\\#\\beta)^{k+1})=a_{0}\\wedge a_{\\neg}\\wedge \\bigwedge_{j=1}^{k}a_{j}$ and the pattern remains.\n\n\\item[\\underline{$\\mathsf{b}((\\alpha\\#\\beta)^{n})$} - ] Completing the process outlined in the previous items, we make $\\mathsf{b}((\\alpha\\#\\beta)^{n})$ equal to \n\\[a_{n}={\\sim}(\\mathsf{b}((\\alpha\\#\\beta)^{n-1})\\wedge \\mathsf{b}(\\neg(\\alpha\\#\\beta)^{n-1}))={\\sim}(a_{0}\\wedge a_{\\neg}\\wedge \\bigwedge_{j=1}^{n-1}a_{j}),\\]\ngiven clause $(V4)_{n}$; and define $\\mathsf{b}(\\neg(\\alpha\\#\\beta)^{n})$ to be simply $\\mathsf{b}((\\alpha\\#\\beta)^{n-1})\\wedge\\mathsf{b}(\\neg(\\alpha\\#\\beta)^{n-1})$, that is, $a_{0}\\wedge a_{\\neg}\\wedge\\bigwedge_{j=1}^{n-1}a_{j}$.\n\\end{enumerate}\n\nFinally, now that $a_{1}$ trough $a_{n}$ are defined, given that $(\\alpha\\#\\beta)^{(n)}=\\bigwedge_{j=1}^{n}(\\alpha\\#\\beta)^{j}$, one obtains \n\\[\\mathsf{b}((\\alpha\\#\\beta)^{(n)})=\\bigwedge_{j=1}^{n}\\mathsf{b}((\\alpha\\#\\beta)^{j})=a_{n}\\wedge\\bigwedge_{j=1}^{n-1}a_{j};\\]\nso\n\\[\\mathsf{b}((\\alpha\\#\\beta)^{(n)})=[{\\sim}(a_{0}\\wedge a_{\\neg})\\vee{\\sim}\\bigwedge_{j=1}^{n-1}a_{j})]\\wedge\\bigwedge_{j=1}^{n-1}a_{j}={\\sim}(a_{0}\\wedge a_{\\neg})\\wedge\\bigwedge_{j=1}^{n-1}a_{j}\\geq a_{\\#}=\\mathsf{b}(\\alpha^{(n)})\\wedge\\mathsf{b}(\\beta^{(n)}),\\]\nsatisfying $(V6)_{n}$ and proving, with some difficulty, $\\mathsf{b}$ is a $\\mathcal{B}$-valuation, as we wanted to show.\n\n\n\n\n\n\n\n\n\\section{Counting snapshots}\\label{Counting snapshots}\n\nWe would like to prove, for completeness sake, that if $\\mathcal{B}$ is a finite Boolean algebra with $2^{m}$ elements, then $B_{n}^{\\mathcal{B}}$ has $(n+2)^{m}$ elements. To this end, we begin with a few observations intended to make our job easier. We notice that the very useful equality\n\\[B_{n+1}^{\\mathcal{B}}=\\{(a_{1}, \\dotsc , a_{n+2})\\in |\\mathcal{B}|^{n+2}\\ :\\ (a_{1}, \\dotsc , a_{n+1})\\in B_{n}^{\\mathcal{B}}\\quad\\text{and}\\quad a_{n+2}\\vee\\bigwedge_{i=1}^{n+1}a_{i}=1\\}\\]\nholds for any $n\\in\\mathbb{N}\\setminus\\{0\\}$.\n\nNow, given all finite Boolean algebras are isomorphic to the field of subsets of some set, we will simply assume all finite Boolean algebras here involved are, in fact, fields of subsets: in that case, the Boolean algebra with $2^{m}$ elements is $\\mathcal{P}(X_{m})$, for $X_{m}=\\{x_{1}, \\dotsc , x_{m}\\}$\\label{Xm} a canonical set with $m$ elements. We will say that an element $a$ of a finite Boolean algebra $\\mathcal{P}(X_{m})$ is of order $k$ if it is a subset of $X_{m}$ with $k$ elements; it is clear that there are $\\binom{m}{0}=1$ elements of order $0$ (namely, $\\emptyset$, the bottom), $\\binom{m}{1}$ elements of order $1$, \\dots , and $\\binom{m}{m}=1$ elements of order $m$ (namely, $X_{m}$, the top). For what will follow, it is important to remember that the binomial coefficient is given by\n\\[\\binom{m}{n}=\\frac{m!}{n!(m-n)!},\\]\nfor $m, n\\in\\mathbb{N}$ and $n\\leq m$, and that the binomial theorem states \n\\[(x+y)^{m}=\\sum_{k=0}^{m}\\binom{m}{k}x^{k}y^{m-k},\\]\nfor $m\\in\\mathbb{N}$, and $x$ and $y$ any elements of a commutative ring with unity.\n\n\\begin{lemma}\\label{finding solutions of a given order}\nFor an element $a$ of $\\mathcal{P}(X_{m})$ of order $k$, there are $\\binom{k}{p}2^{m-k}$ elements $b$ such that $a\\wedge b$ has order $p\\leq k$, and $\\binom{m-k}{q}2^{k}$ elements $c$ such that $a\\vee c$ has order $q\\geq k$.\n\\end{lemma}\n\n\\begin{proof}\nIf $a\\wedge b$ has order $p$, this means $b$ has $p$ elements in common with $a$, and since there are $k$ elements in $a$, we obtain $\\binom{k}{p}$ possible values for $a\\cap b$; but $a^{c}\\cap b$ can be any subset of $a^{c}$, of which there exist $2^{m-k}$ given $a^{c}$ has $m-k$ elements. Combining the two values, we obtain $\\binom{k}{p}2^{m-k}$ solutions.\n\nNow, if $a\\vee c$ has order $q$, $a^{c}\\cap c$ has $q-k$ elements that $a$ doesn't, and since there are $m-k$ elements in $a^{c}$, this gives us $\\binom{m-k}{q}$ values for $a^{c}\\cap c$; but $a\\cap c$ can be any subset of $a$, giving us $2^{k}$ values for $a\\cap c$ since $a$ has $k$ elements. This leads to the total $\\binom{m-k}{q}2^{k}$ solutions.\n\\end{proof}\n\nNotice that, by adapting the proof of the previous lemma, we may show that, for an $a$ of order $k$, there are $\\binom{k}{p}$ elements $b$ such that $a\\vee b=1$ and $a\\wedge b$ has order $p\\leq k$. This is because: $a^{c}\\cap b$ must equal $a^{c}$, so that $a\\vee b=1$; and $a\\cap b$ must have $p$ elements, and since $a$ has $k$ elements, this gives us $\\binom{k}{p}$ solutions.\n\n\\begin{lemma}\\label{a case of the binomial theorem}\nFor $m\\in\\mathbb{N}$ and $p\\leq m$,\n\\[\\sum_{j=p}^{m}\\binom{j}{p}\\binom{m}{j}x^{m-j}=\\binom{m}{p}(x+1)^{m-p}.\\]\n\\end{lemma}\n\n\\begin{proof}\nWe see that, from the definition of the binomial coefficients and the binomial theorem,\n\\[\\sum_{j=p}^{m}\\binom{j}{p}\\binom{m}{j}x^{m-j}=\\sum_{j=p}^{m}\\frac{j!}{p!(j-p)!}\\frac{m!}{j!(m-j)!}x^{m-j}=\\]\n\\[\\sum_{j=p}^{m}\\frac{m!}{p!(j-p)!(m-j)!}x^{m-j}=\\sum_{j=p}^{m}\\frac{1}{p!}\\frac{m!}{(j-p)!(m-j)!}x^{m-j}=\\]\n\\[\\sum_{j=p}^{m}\\frac{m!}{p!(m-p)!}\\frac{(m-p)!}{(j-p)!(m-j)!}x^{m-j}=\\]\n\\[\\binom{m}{p}\\sum_{i=0}^{m-p}\\frac{(m-p)!}{((i+p)-p)!(m-(i+p))!}x^{m-(i+p)}=\\]\n\\[\\binom{m}{p}\\sum_{i=0}^{m-p}\\frac{(m-p)!}{i!((m-p)-i)!}x^{(m-p)-i}=\\binom{m}{p}\\sum_{i=0}^{m-p}\\binom{m-p}{i}x^{(m-p)-i}1^{i}=\\]\n\\[\\binom{m}{p}(x+1)^{m-p}.\\]\n\\end{proof}\n\n\n\\begin{lemma}\\label{counting certain snapshots}\nFor any $n\\geq 0$, if $\\mathcal{B}$ has $2^{m}$ elements then $B_{n}^{\\mathcal{B}}$ has precisely $\\binom{m}{p}(n+1)^{m-p}$ elements $(a_{1}, \\dotsc , a_{n+1})$ such that $\\bigwedge_{i=1}^{n+1}a_{i}$ has order $p\\leq m$.\n\\end{lemma}\n\n\\begin{proof}\nLet $(a,b)$ be a pair in $B_{1}^{\\mathcal{B}}$, meaning $a\\vee b=1$, such that $a\\wedge b$ has order $p$. Then, by the commentary after Lemma \\ref{finding solutions of a given order}, if $a$ has order $k\\geq p$, there are $\\binom{k}{p}$ possible values for $b$, and if $k\\leq p$ there are none. Since there are $\\binom{m}{k}$ elements $a$ with order $k$, this gives us the number of pairs satisfying $a\\vee b=1$ and $a\\wedge b$ having order $p$ to be\n\\[\\binom{p}{p}\\binom{m}{p}+\\binom{p+1}{p}\\binom{m}{p+1}+\\cdots+\\binom{m}{p}\\binom{m}{m},\\]\nwhich equals $\\binom{m}{p}2^{m-p}$ from Lemma \\ref{a case of the binomial theorem}, once we use $x=1$.\n\nNow, suppose the described conditions hold for $B_{n}^{\\mathcal{B}}$: there are, by induction hypothesis, $\\binom{m}{k}(n+1)^{m-k}$ elements $(a_{1}, \\dotsc , a_{n+1})$ such that $\\bigwedge_{i=1}^{n+1}a_{i}$ is of order $k$, and then there are $\\binom{k}{p}$ possible values for $a_{n+3}$, such that $(a_{1}, \\dotsc , a_{n+2})$ is in $B_{n+1}^{\\mathcal{B}}$ (meaning $a_{n+2}\\vee\\bigwedge_{i=1}^{n+1}a_{i}=1$) and $\\bigwedge_{i=1}^{n+2}a_{i}$ has order $p$, what gives the total\n\\[\\binom{p}{p}\\binom{m}{p}(n+1)^{m-p}+\\cdots+\\binom{m}{p}\\binom{m}{m}(n+1)^{m-m},\\]\nwhich adds to $\\binom{m}{p}(n+2)^{m-p}$ once we apply $x=n+1$ to Lemma \\ref{a case of the binomial theorem}, ending our proof.\n\\end{proof}\n\n\\begin{theorem}\nIf $\\mathcal{B}$ has $2^{m}$ elements, $B_{n}^{\\mathcal{B}}$ has $(n+2)^{m}$ elements.\n\\end{theorem}\n\n\\begin{proof}\nBy Lemma \\ref{counting certain snapshots}, there are $\\binom{m}{0}(n+1)^{m-0}$ elements $(a_{1}, \\dotsc , a_{n+1})$ in $B_{n}^{\\mathcal{B}}$ such that $\\bigwedge_{i=1}^{n+1}a_{i}$ has order $0$, $\\binom{m}{1}(n+1)^{m-1}$ elements such that $\\bigwedge_{i=1}^{n+1}a_{i}$ has order $1$ and so on, adding up to a total of\n\\[\\binom{m}{0}(n+1)^{m-0}+\\binom{m}{1}(n+1)^{m-1}+\\cdots+\\binom{m}{m}(n+1)^{m-m}=\\]\n\\[\\sum_{p=0}^{m}\\binom{m}{p}(n+1)^{m-p}1^{p}=((n+1)+1)^{m}=(n+2)^{m}.\\]\n\\end{proof}\n\nNotice, furthermore, that $D_{1}^{\\mathcal{B}}$ has $2^{m}$ elements whenever $\\mathcal{B}$ has $2^{m}$ elements: this happens since $(a, b)\\in D_{1}^{\\mathcal{B}}$ whenever $a=1$ and $a\\vee b=1$, meaning $b$ can assume any value in $\\mathcal{B}$; and when $n\\geq 1$,\n\\[D_{n+1}^{\\mathcal{B}}=\\{(1, a_{1}, \\dotsc , a_{n+1})\\in B_{n+1}^{\\mathcal{B}}: (a_{1}, \\dotsc , a_{n+1})\\in B_{n}^{\\mathcal{B}}\\},\\]\nimplying $D_{n+1}^{\\mathcal{B}}$ has as many elements as $B_{n}^{\\mathcal{B}}$.\n\n\\begin{theorem}\nIf $\\mathcal{B}$ has $2^{m}$ elements:\n\\begin{enumerate}\n\\item $D_{n}^{\\mathcal{B}}$ has $(n+1)^{m}$ elements;\n\\item $Boo_{n}^{\\mathcal{B}}$ has $2^{m}$ elements.\n\\end{enumerate}\n\\end{theorem}\n\nA final, relevant, note is on the case that $\\mathcal{B}$ is infinite, and of cardinality $\\kappa$: since $B_{n}^{\\mathcal{B}}$ contains the set of Boolean elements $Boo_{n}^{\\mathcal{B}}$, isomorphic to $\\mathcal{B}$ itself, we can be sure that $B_{n}^{\\mathcal{B}}$ has at least cardinality $\\kappa$; however, since $B_{n}^{\\mathcal{B}}$ is a subset of $|\\mathcal{B}|^{n+1}$, which is too of cardinality $\\kappa$ given the assumption this is an infinite cardinal, we obtain that $B_{n}^{\\mathcal{B}}$ has precisely $\\kappa$ elements.\n\nTo provide one example of $B_{n}^{\\mathcal{B}}$'s complexity, let us take the four-valued Boolean algebra $\\mathcal{B}_{4}$ as the field of subsets of $\\{a,b\\}$, containing the elements $\\emptyset$ (which we shall denote by $0$), $\\{a\\}$, $\\{b\\}$, and $\\{a,b\\}$ (which we will denote by $1$). Then, $B_{1}^{\\mathcal{B}_{4}}$ has $9$ snapshots.\n\\begin{enumerate}\n\\item Designated and Boolean: $(1, 0)$.\n\\item Designated and not Boolean: $(1, \\{a\\})$, $(1, \\{b\\})$ and $(1, 1)$.\n\\item Undesignated and Boolean: $(0,1)$, $(\\{a\\}, \\{b\\})$ and $(\\{b\\}, \\{a\\})$.\n\\item Undesignated and not Boolean: $(\\{a\\}, 1)$ and $(\\{b\\}, 1)$.\n\\end{enumerate}\n\nStill on the Boolean algebra with four elements, $B_{2}^{\\mathcal{B}_{4}}$ has $16$ snapshots.\n\\begin{enumerate}\n\\item Designated and Boolean: $(1, 0, 1)$.\n\\item Designated and not Boolean: $(1, 1, 0)$, $(1, 1, \\{a\\})$, $(1, 1, \\{b\\})$, $(1, 1, 1)$, $(1, \\{a\\}, \\{b\\})$,\\\\ $(1, \\{b\\}, \\{a\\})$, $(1, \\{a\\}, 1)$ and $(1, \\{b\\}, 1)$.\n\\item Undesignated and Boolean: $(0, 1, 1)$, $(\\{a\\}, \\{b\\}, 1)$ and $(\\{b\\}, \\{a\\}, 1)$.\n\\item Undesignated and not Boolean: $(\\{a\\}, 1, 1)$, $(\\{a\\}, 1, \\{b\\})$, $(\\{b\\}, 1, 1)$ and $(\\{b\\}, 1, \\{a\\})$.\n\n\\end{enumerate}\n\nNow, consider the eight-valued Boolean algebra $\\mathcal{B}_{8}$, which we will take to be the Boolean algebra over the powerset of $\\{a, b, c\\}$. In that case, $B_{1}^{\\mathcal{B}_{8}}$ has $27$ elements:\n\n\\begin{enumerate}\n\\item Designated and Boolean: $(1, 0)$.\n\\item Designated and not Boolean: $(1, \\{a\\})$, $(1, \\{b\\})$, $(1, \\{c\\})$, $(1, \\{b, c\\})$, $(1, \\{a, c\\})$, $(1, \\{a, b\\})$ and $(1, 1)$.\n\\item Undesignated and Boolean: $(0, 1)$, $(\\{a\\}, \\{b, c\\})$, $(\\{b\\}, \\{a, c\\})$, $(\\{c\\}, \\{a, b\\})$, $(\\{b, c\\}, \\{a\\})$, $(\\{a, c\\}, \\{b\\})$ and $(\\{a, b\\}, \\{c\\})$.\n\\item Undesignated and not Boolean: $(\\{a\\}, 1)$, $(\\{b\\}, 1)$, $(\\{c\\}, 1)$, $(\\{b, c\\}, \\{a, c\\})$, $(\\{b, c\\}, \\{a, b\\})$, $(\\{b, c\\}, 1)$, $(\\{a, c\\}, \\{b, c\\})$, $(\\{a, c\\}, \\{a, b\\})$, $(\\{a, c\\}, 1)$, $(\\{a, b\\}, \\{b, c\\})$, $(\\{a, b\\}, \\{a, c\\})$\\\\ and $(\\{a, b\\}, 1)$.\n\\end{enumerate}\n\n\n\n\n\n\n\n\n\n\n\n\\section{Category of restricted swap structures for $C_{n}$}\n\nGiven a class $C$ of RNmatrices, how to make it into a category $\\mathcal{C}$? We believe there may exist more than one natural definition of what a morphism on $\\mathcal{C}$ should be, the applications desired for such an object dictating the best ones, but at least one definition seems to be more or less universal. To start defining it, remember: an RNmatrix is a triple $\\mathcal{M}=(\\mathcal{A}, D, \\mathcal{F})$, for $\\mathcal{A}$ a $\\Sigma$-multialgebra, $D$ a subset of its universe and $\\mathcal{F}$ a set of homomorphisms (of multialgebras) $\\nu:\\textbf{F}(\\Sigma, \\mathcal{V})\\rightarrow\\mathcal{A}$. So a morphism between RNmatrices $\\mathcal{M}=(\\mathcal{A}, D, \\mathcal{F})$ and $\\mathcal{M}^{\\prime}=(\\mathcal{A}^{\\prime}, D^{\\prime}, \\mathcal{F}^{\\prime})$, over the same signature $\\Sigma$, should be, at a minimum:\n\\begin{enumerate}\n\\item a homomorphism of multialgebras $h:\\mathcal{A}\\rightarrow\\mathcal{A}^{\\prime}$;\n\\item which preserves designated elements, meaning that $h(D)\\subseteq D^{\\prime}$;\n\\item and is absorbed by restricted valuations, meaning that for every $\\nu\\in\\mathcal{F}$, $h\\circ\\nu\\in\\mathcal{F}^{\\prime}$.\n\\begin{figure}[H]\n\\centering\n\\begin{tikzcd}\n \\mathcal{M} \\arrow{rr}{h} & & \\mathcal{M}^{\\prime}\\\\\n & \\textbf{F}(\\Sigma, \\mathcal{V}) \\arrow{ul}{\\nu} \\arrow{ur}{h\\circ\\nu} & \n \\end{tikzcd}\n \\caption*{If $\\nu\\in\\mathcal{F}$, $h\\circ\\nu$ must be in $\\mathcal{F}^{\\prime}$}\n\\end{figure}\n\\end{enumerate}\n\n\\begin{theorem}\\label{category of RNmatrices}\nA class of RNmatrices $C$, endowed with the morphisms defined above, becomes a category $\\mathcal{C}$.\n\\end{theorem}\n\n\\begin{proof}\nTake morphisms $g:\\mathcal{M}\\rightarrow\\mathcal{M}^{\\prime}$ and $h:\\mathcal{M}^{\\prime}\\rightarrow\\mathcal{M}^{\\prime\\prime}$ as described above, and first of all we will show that their composition, as a composition of functions, returns again such a morphism.\n\\begin{enumerate}\n\\item Since $g$ and $h$ are both $\\Sigma$-homomorphisms of multialgebras, respectively from $\\mathcal{A}$ to $\\mathcal{A}^{\\prime}$ and from $\\mathcal{A}^{\\prime}$ to $\\mathcal{A}^{\\prime\\prime}$, it is clear that $h\\circ g$ is a $\\Sigma$-homomorphism from $\\mathcal{A}$ to $\\mathcal{A}^{\\prime\\prime}$.\n\\item Given that $g(D)\\subseteq D^{\\prime}$ and $h(D^{\\prime})\\subseteq D^{\\prime\\prime}$, we have that $h\\circ g(D)=h(g(D))\\subseteq h(D^{\\prime})\\subseteq D^{\\prime\\prime}$, meaning that $h\\circ g$ preserves designated elements.\n\\item Finally, we have that for any $\\nu\\in\\mathcal{F}$ and any $\\nu^{\\prime}\\in\\mathcal{F}^{\\prime}$, $g\\circ\\nu\\in\\mathcal{F}^{\\prime}$ and $h\\circ\\nu^{\\prime}\\in\\mathcal{F}^{\\prime\\prime}$; thus, for any $\\nu\\in\\mathcal{F}$, $g\\circ \\nu\\in\\mathcal{F}^{\\prime}$ and so $(h\\circ g)\\circ \\nu=h\\circ(g\\circ \\nu)\\in\\mathcal{F}^{\\prime\\prime}$, proving $h\\circ g$ is absorbed by restricted valuations.\n\\end{enumerate}\n\nThe associativity of the composition of these morphisms comes from the associativity of the composition of functions. The identity morphisms are precisely the identity functions, meaning that given $\\mathcal{M}=(\\mathcal{A}, D, \\mathcal{F})$, the identity morphism in $\\mathcal{M}$ is the identity function on the universe of $\\mathcal{A}$, easily seem to be a homomorphism of multialgebras from $\\mathcal{A}$ to itself; that preserves designated elements; and is absorbed by restricted valuations.\n\\end{proof}\n\nNow, for a fixed da Costa's logic $C_{n}$, we take the class of RNmatrices $\\mathcal{RM}^{\\mathcal{B}}_{C_{n}}$ for $\\mathcal{B}$ a non-trivial Boolean algebra and construct its corresponding category $\\textbf{RSwap}_{C_{n}}$\\label{RSwapCn} as it was done in Theorem \\ref{category of RNmatrices} just above. To spell out our definition in this very important case, $\\textbf{RSwap}_{C_{n}}$ is the category: \n\\begin{enumerate}\n\\item with the class of (full) restricted swap structures $\\mathcal{RM}_{C_{n}}^{\\mathcal{B}}$, for all Boolean algebras $\\mathcal{B}$, as objects; \n\\item as morphisms between $\\mathcal{RM}_{C_{n}}^{\\mathcal{B}_{1}}$ and $\\mathcal{RM}_{C_{n}}^{\\mathcal{B}_{2}}$, all functions $\\varphi:B_{n}^{\\mathcal{B}_{1}}\\rightarrow B_{n}^{\\mathcal{B}_{2}}$ such that\n\\begin{enumerate}\n\\item $\\varphi$ is a $\\Sigma_{C}$-homomorphism between $\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{1}}$ and $\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{2}}$, seem as $\\Sigma_{C}$-multialgebras;\n\\item for all $d\\in D_{n}^{\\mathcal{B}_{1}}$, $\\varphi(d)$ is in $D_{n}^{\\mathcal{B}_{2}}$;\n\\item for any $\\nu:\\textbf{F}(\\Sigma_{C}, \\mathcal{V})\\rightarrow\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{1}}$ in $\\mathcal{F}_{C_{n}}^{\\mathcal{B}_{1}}$, $\\varphi\\circ\\nu:\\textbf{F}(\\Sigma_{C}, \\mathcal{V})\\rightarrow\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{2}}$ is in $\\mathcal{F}_{C_{n}}^{\\mathcal{B}_{2}}$.\n\\end{enumerate}\n\\end{enumerate}\n\n\\begin{figure}[H]\n\\centering\n\\begin{tikzcd}\n \\mathcal{A}_{C_{n}}^{\\mathcal{B}_{1}} \\arrow{rr}{\\varphi} & & \\mathcal{A}_{C_{n}}^{\\mathcal{B}_{2}}\\\\\n & \\textbf{F}(\\Sigma_{C}, \\mathcal{V}) \\arrow{ul}{\\nu} \\arrow{ur}{\\varphi\\circ\\nu} & \n \\end{tikzcd}\n \\caption*{If $\\nu\\in\\mathcal{F}_{C_{n}}^{\\mathcal{B}_{1}}$, $\\varphi\\circ\\nu$ must be in $\\mathcal{F}_{C_{n}}^{\\mathcal{B}_{2}}$}\n\\end{figure}\n\nOne important point to make is that demanding that a morphism of $\\textbf{RSwap}_{C_{n}}$ preserves designated elements is actually superfluous: we could prove this property from the fact that the morphisms in this category are absorbed by restricted valuations. However, the proof of this fact is somewhat involved, so it is easier to assume the preservation of designated elements as one of the defining characteristics of our morphisms.\n\nA first question regarding the definition of $\\textbf{RSwap}_{C_{n}}$ could be whether there are truly morphisms $\\varphi$ in it other than the identity ones.\n\n\\begin{proposition}\\label{homomorphisms of boolean algebras lead to morphisms}\nFor Boolean algebras $\\mathcal{B}_{1}$ and $\\mathcal{B}_{2}$ and a homomorphism $\\psi:\\mathcal{B}_{1}\\rightarrow\\mathcal{B}_{2}$ of Boolean algebras, $\\varphi:B_{n}^{\\mathcal{B}_{1}}\\rightarrow B_{n}^{\\mathcal{B}_{2}}$ such that, for every $z=(z_{1}, \\dotsc , z_{n+1})\\in B_{n}^{\\mathcal{B}_{1}}$,\n\\[\\varphi(z)_{i}=\\psi(z_{i}),\\quad\\text{for every}\\quad 1\\leq i\\leq n+1,\\]\nwhat we may write as $\\varphi(z)=(\\psi(z_{1}), \\dotsc , \\psi(z_{n+1}))$, is a morphism between $\\mathcal{RM}_{C_{n}}^{\\mathcal{B}_{1}}$ and $\\mathcal{RM}_{C_{n}}^{\\mathcal{B}_{2}}$ in $\\textbf{RSwap}_{C_{n}}$.\n\\end{proposition}\n\n\\begin{proof}\nNotice that if $z$ is a Boolean element of $\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{1}}$, then it is of the form $(a, {\\sim}a, 1, \\dotsc , 1)$, and $\\varphi(z)=(\\psi(a), \\psi({\\sim}a), \\psi(1), \\dotsc , \\psi(1))=(\\psi(a), {\\sim}\\psi(a), 1, \\dotsc , 1)$, given $\\psi$ is a homomorphism; and since $\\psi(a)$ is an element of $\\mathcal{B}_{2}$, we see $\\varphi(z)$ is a Boolean element of $\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{2}}$, meaning $\\varphi$ preserves Boolean elements. \n\nTake elements $w$ and $z$ in $B_{n}^{\\mathcal{B}_{1}}$. First, suppose $w$ and $z$ are not both Boolean elements and $u\\in w\\tilde{\\#}z$, meaning that $u_{1}=w_{1}\\# z_{1}$: then \n\\[\\varphi(u)=(\\psi(u_{1}), \\psi(u_{2}), \\dotsc , \\psi(u_{n+1}))\\quad\\text{equals}\\quad(\\psi(w_{1})\\#\\psi(z_{1}), \\psi(u_{2}), \\dotsc , \\psi(u_{n+1})),\\]\nsince $\\psi$ is a homomorphism, and since $\\varphi(w)_{1}=\\psi(w_{1})$ and $\\varphi(z)_{1}=\\psi(z_{1})$, we obtain that $\\varphi(u)\\in \\varphi(w)\\tilde{\\#}\\varphi(z)$. Now, if $w$ and $z$ are both Boolean elements and $u\\in w\\tilde{\\#}z$, then $u_{1}=w_{1}\\# z_{1}$ and $u$ is a Boolean element, and from what we saw above we find that $\\varphi(u)_{1}=\\varphi(w)_{1}\\#\\varphi(z)_{1}$ and that $\\varphi(w)$, $\\varphi(z)$ and $\\varphi(u)$ are all Boolean elements, meaning $\\varphi(u)\\in\\varphi(w)\\tilde{\\#}\\varphi(z)$.\n\nNow, for any $z\\in B_{n}^{\\mathcal{B}_{1}}$ and $w\\in\\tilde{\\neg}z$, $w_{1}=z_{2}$ and $w_{2}\\leq z_{1}$. Then $\\varphi(w)=(\\psi(w_{1}), \\dotsc , \\psi(w_{n+1}))$, and analogously for $\\varphi(z)$, meaning $\\varphi(w)_{1}=\\psi(w_{1})=\\psi(z_{2})=\\varphi(z)_{2}$ and $\\varphi(w)_{2}=\\psi(w_{2})\\leq \\psi(z_{1})=\\varphi(z)_{1}$,\\footnote{One should remember that homomorphisms of Boolean algebras preserve order: first of all, in a Boolean algebra $a\\leq b$ iff $a=a\\wedge b$ iff $b=a\\vee b$; given a homomorphism $\\psi:\\mathcal{B}_{1}\\rightarrow \\mathcal{B}_{2}$, if $a\\leq b$, $b=a\\vee b$ and so $\\psi(b)=\\psi(a\\vee b)=\\psi(a)\\vee\\psi(b)$, meaning that $\\psi(a)\\leq \\psi(b)$.}and therefore $\\varphi(w)\\in\\tilde{\\neg}\\varphi(z)$. With this, $\\varphi$ is a $\\Sigma_{\\textbf{C}}$-homomorphism between $\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{1}}$ and $\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{2}}$.\n\nGiven a designated element $z=(1, z_{2}, \\dotsc , z_{n+1})$ of $\\mathcal{RM}_{C_{n}}^{\\mathcal{B}_{1}}$, we get that $\\varphi(z)=(\\psi(1), \\psi(z_{2}), \\dotsc , \\psi(z_{n+1}))=(1, \\psi(z_{2}), \\dotsc , \\psi(z_{n+1}))$ is also designated, meaning that $\\varphi$ preserves designated elements.\n\nNow, a $\\Sigma_{\\textbf{C}}$-homomorphism $\\nu:\\textbf{F}(\\Sigma_{\\textbf{C}}, \\mathcal{V})\\rightarrow\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{1}}$ is in $\\mathcal{F}_{C_{n}}^{\\mathcal{B}_{1}}$ when, for all formulas $\\alpha$ and $\\beta$:\n\\begin{enumerate}\n\\item $\\nu(\\alpha\\wedge\\neg\\alpha)_{2}=\\nu(\\alpha)_{3}$;\n\\item $\\nu(\\alpha^{1})=(\\nu(\\alpha)_{3}, \\nu(\\alpha)_{1}\\wedge\\nu(\\alpha)_{2}, \\nu(\\alpha)_{4}, \\dotsc , \\nu(\\alpha)_{n+1}, {\\sim}(\\bigwedge_{i=1}^{n+1}\\nu(\\alpha)_{i}))$;\n\\item and $\\nu((\\alpha^{(n)}\\wedge\\beta^{(n)})\\rightarrow(\\alpha\\#\\beta)^{(n)})\\in D_{n}^{\\mathcal{B}_{1}}$, for $\\#\\in\\{\\vee, \\wedge, \\rightarrow\\}$.\n\\end{enumerate}\n\nQuite clearly $\\varphi\\circ\\nu:\\textbf{F}(\\Sigma_{\\textbf{C}}, \\mathcal{V})\\rightarrow\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{2}}$ is still a $\\Sigma_{\\textbf{C}}$-homomorphism.\n\n\\begin{enumerate}\n\\item We have $\\varphi(\\nu(\\alpha\\wedge\\neg\\alpha))_{2}=\\psi(\\nu(\\alpha\\wedge\\neg\\alpha)_{2})$, which equals, given $\\nu$ is in $\\mathcal{F}_{C_{n}}^{\\mathcal{B}_{1}}$, $\\psi(\\nu(\\alpha)_{3})=\\varphi(\\nu(\\alpha))_{3}$.\n\\item See that $\\varphi(\\nu(\\alpha^{1}))$ equals\n\\[\\big(\\psi(\\nu(\\alpha)_{3}), \\psi(\\nu(\\alpha)_{1}\\wedge\\nu(\\alpha)_{2}), \\psi(\\nu(\\alpha)_{4}), \\dotsc , \\psi(\\nu(\\alpha)_{n+1}), \\psi({\\sim}(\\bigwedge_{i=1}^{n+1}\\nu(\\alpha)_{i}))\\big)=\\]\n\\[\\big(\\psi(\\nu(\\alpha)_{3}), \\psi(\\nu(\\alpha)_{1})\\wedge\\psi(\\nu(\\alpha)_{2}), \\psi(\\nu(\\alpha)_{4}), \\dotsc , \\psi(\\nu(\\alpha)_{n+1}), {\\sim}(\\bigwedge_{i=1}^{n+1}\\psi(\\nu(\\alpha)_{i}))\\big)=\\]\n\\[\\big(\\varphi(\\nu(\\alpha))_{3}, \\varphi(\\nu(\\alpha))_{1}\\wedge\\varphi(\\nu(\\alpha))_{2}, \\varphi(\\nu(\\alpha))_{4}, \\dotsc , \\varphi(\\nu(\\alpha))_{n+1}, {\\sim}(\\bigwedge_{i=1}^{n+1}\\varphi(\\nu(\\alpha))_{i})\\big).\\]\n\\item Since, as we already proved, $\\varphi$ preserves designated elements, $\\nu((\\alpha^{(n)}\\wedge\\beta^{(n)})\\rightarrow(\\alpha\\#\\beta)^{(n)})\\in D_{n}^{\\mathcal{B}_{1}}$, for any $\\#\\in\\{\\vee, \\wedge, \\rightarrow\\}$, implies that $\\varphi(\\nu((\\alpha^{(n)}\\wedge\\beta^{(n)})\\rightarrow(\\alpha\\#\\beta)^{(n)}))\\in D_{n}^{\\mathcal{B}_{2}}$ for any $\\#\\in\\{\\vee, \\wedge, \\rightarrow\\}$.\n\\end{enumerate}\nThis finishes proving that $\\varphi\\circ\\nu$ lies in $\\mathcal{F}_{C_{n}}^{\\mathcal{B}_{2}}$.\n\n\\end{proof}\n\nNotice that, in the spirit of the last proposition, the identity morphism $Id_{\\mathcal{RM}_{C_{n}}^{\\mathcal{B}}}$ of $\\textbf{RSwap}_{C_{n}}$ on $\\mathcal{RM}_{C_{n}}^{\\mathcal{B}}$ can be written, for an arbitrary $z=(z_{1}, \\dotsc , z_{n+1})$ in $B_{n}^{\\mathcal{B}}$, as\n\\[Id_{\\mathcal{A}_{C_{n}}^{\\mathcal{B}}}(z)=(z_{1}, \\dotsc , z_{n+1})=(Id_{\\mathcal{B}}(z_{1}), \\dotsc , Id_{\\mathcal{B}}(z_{n+1})),\\]\nfor $Id_{\\mathcal{B}}:\\mathcal{B}\\rightarrow\\mathcal{B}$ the identity homomorphism on $\\mathcal{B}$, and therefore the identity morphisms in $\\textbf{RSwap}_{C_{n}}$ are, too, of the form described in the previous proposition.\n\nA second question regarding the definition of $\\textbf{RSwap}_{C_{n}}$ is whether there are morphisms in it other than those of the form described in Proposition \\ref{homomorphisms of boolean algebras lead to morphisms}. The answer, in this case, is no, and in the following section we prove exactly that.\n\n\n\n\n\n\n\n\n\n\\subsection{Morphisms of $\\textbf{RSwap}_{C_{n}}$}\n\nFor any function $\\varphi:B_{n}^{\\mathcal{B}_{1}}\\rightarrow B_{n}^{\\mathcal{B}_{2}}$ we will write, for any $z\\in B_{n}^{\\mathcal{B}_{1}}$, $\\varphi(z)$ as $(\\varphi_{1}(z), \\dotsc , \\varphi_{n+1}(z))$: here, $\\varphi_{i}$, for any $1\\leq i\\leq n+1$, is a function from $B_{n}^{\\mathcal{B}_{1}}$ to $\\mathcal{B}_{2}$ obtained by composing the $i$-th projection $\\pi_{i}$\\label{pii} of $B_{n}^{\\mathcal{B}_{1}}$ with $\\varphi$.\n\nWith this, we can say that $\\varphi$ is a $\\Sigma_{\\textbf{C}}$-homomorphism from $\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{1}}$ to $\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{2}}$ if, and only if, for all snapshots $w, z\\in B_{n}^{\\mathcal{B}_{1}}$, and $u\\in w\\tilde{\\#}z$ (for any $\\#\\in\\{\\vee, \\wedge, \\rightarrow\\}$) and $v\\in\\tilde{\\neg}z$, \n\\[\\varphi_{1}(u)=\\varphi_{1}(w)\\#\\varphi_{1}(z),\\quad\\text{and}\\quad\\varphi_{1}(v)=\\varphi_{2}(z)\\quad\\text{and}\\quad\\varphi_{2}(v)\\leq\\varphi_{1}(z),\\]\nwhat is equivalent to $\\varphi(u)\\in \\varphi(w)\\tilde{\\#}\\varphi(z)$ and $\\varphi(v)\\in\\tilde{\\neg}\\varphi(z)$ once one considers the definitions of $\\tilde{\\#}$ and $\\tilde{\\neg}$.\n\nSo, we would like to study the function $\\psi:\\mathcal{B}_{1}\\rightarrow\\mathcal{B}_{2}$, given by $\\psi(a)=$\\\\$\\varphi((a, {\\sim}a, 1, \\dotsc , 1))$ for any $a$ in $\\mathcal{B}_{1}$, in the case that $\\varphi$ is actually a $\\Sigma_{\\textbf{C}}$-homomorphism. Take a snapshot $z=(z_{1}, \\dotsc , z_{n+1})$ of $B_{n}^{\\mathcal{B}_{1}}$ and consider $z^{*}=(z_{1}, {\\sim}z_{1}, 1, \\dotsc , 1)$: then $\\psi(z_{1})=\\varphi(z^{*})$ by definition of $\\psi$. Furthermore, recalling that $t^{n}_{0}=(1, 1, 0, 1, \\dotsc , 1)$ is a non-Boolean snapshot of $B_{n}^{\\mathcal{B}}$, for any non-trivial Boolean algebra $\\mathcal{B}$, we find that\n\\[z\\tilde{\\wedge}t^{n}_{0}=z^{*}\\tilde{\\wedge}t^{n}_{0}=\\{w=(w_{1}, \\dotsc , w_{n+1})\\in B_{n}^{\\mathcal{B}_{1}}: w_{1}=z_{1}\\},\\]\nand thus $z, z^{*}\\in z\\tilde{\\wedge}t^{n}_{0}=z^{*}\\tilde{\\wedge}t^{n}_{0}$, since the first coordinates of both are precisely $z_{1}$; given $\\varphi$ is a homomorphism, $z\\in z\\tilde{\\wedge}t^{n}_{0}$ implies that $\\varphi_{1}(z)=\\varphi_{1}(z)\\wedge\\varphi_{1}(t^{n}_{0})$, while $z^{*}\\in z\\tilde{\\wedge}t^{n}_{0}$ implies that $\\varphi_{1}(z^{*})=\\varphi_{1}(z)\\wedge\\varphi_{1}(t^{n}_{0})$, leading us to $\\varphi_{1}(z)=\\varphi_{1}(z^{*})$ and then to $\\varphi_{1}(z)=\\psi(z_{1})$. In other words, the function $\\varphi_{1}$ depends exclusively on the first coordinate of a snapshot.\n\nFurthermore, for any $a, b\\in\\mathcal{B}_{1}$, it is true for the snapshots $(a, {\\sim}a, 1, \\dotsc , 1)$ and $(b, {\\sim}b, 1, \\dotsc , 1)$ in $B_{n}^{\\mathcal{B}_{1}}$ that \n\\[(a, {\\sim}a, 1, \\dotsc , 1)\\tilde{\\#}(b, {\\sim}b, 1, \\dotsc , 1)=\\{(a\\#b, {\\sim}a\\#b, 1, \\dotsc , 1)\\},\\]\nfor any $\\#\\in\\{\\vee, \\wedge, \\rightarrow\\}$. This way, taking into consideration that $\\varphi$ is a homomorphism and $\\varphi_{1}(z)=\\psi(z_{1})$, $\\psi(a\\#b)=\\psi(a)\\#\\psi(b)$. This means that $\\psi$ is almost a homomorphism, and we will prove further ahead that it actually is a homomorphism of Boolean algebras.\n\nFor now, let us also define the function $\\theta:\\mathcal{B}_{1}\\rightarrow\\mathcal{B}_{2}$ by $\\theta(a)=\\varphi_{2}(({\\sim}a, a, 1, \\dotsc , 1))$. For an arbitrary snapshot $z=(z_{1}, \\dotsc , z_{n+1})$ in $B_{n}^{\\mathcal{B}_{1}}$, consider \n\\[z^{*}=({\\sim}z_{2}, z_{2}, 1, \\dotsc , 1)\\quad\\text{and}\\quad z^{\\prime}=(z_{2}, {\\sim}z_{2}, 1, \\dotsc , 1),\\]\nand by definition of $\\theta$ we have that $\\varphi_{2}(z^{*})=\\theta(z_{2})$. Since $z$ is in $B_{n}^{\\mathcal{B}_{1}}$, we have that $z_{1}\\vee z_{2}=1$ and so ${\\sim}z_{2}\\leq z_{1}$, implying that $z^{\\prime}\\in\\tilde{\\neg}z$ and $z^{\\prime}\\in\\tilde{\\neg}z^{*}$. \\footnote{Actually $\\tilde{\\neg}z^{*}=\\{z^{\\prime}\\}$ and vice-versa, since both are Boolean snapshots with complementary first-coordinates.} Using, once again, that $\\varphi$ is a homomorphism, $\\varphi(z^{\\prime})\\in \\tilde{\\neg}\\varphi(z)$ and $\\varphi(z^{\\prime})\\in\\tilde{\\neg}\\varphi(z^{*})$, leading to $\\varphi_{1}(z^{\\prime})=\\varphi_{2}(z)$ and $\\varphi_{1}(z^{\\prime})=\\varphi_{2}(z^{*})$, \\textit{id est} $\\varphi_{2}(z)=\\varphi_{2}(z^{*})$. This leads us to $\\varphi_{2}(z)=\\theta(z_{2})$, meaning $\\varphi_{2}$ also depends exclusively of one coordinate, in this case the second.\n\nFinally, we may prove, with our only hypothesis being that $\\varphi$ is a homomorphism, that $\\psi=\\theta$. To see this, take any $a\\in\\mathcal{B}_{1}$, and the snapshots $z=(a, {\\sim}a, 1, \\dotsc , 1)$ and $z^{*}=({\\sim}a, a, 1, \\dotsc , 1)$ of $B_{n}^{\\mathcal{B}_{1}}$, with the property that $\\tilde{\\neg}z=\\{z^{*}\\}$ and $\\tilde{\\neg}z^{*}=\\{z\\}$. Then $\\varphi(z)\\in \\tilde{\\neg}\\varphi(z^{*})$, so $\\varphi_{1}(z)=\\varphi_{2}(z^{*})$ (and also $\\varphi_{2}(z)\\leq \\varphi_{2}(z^{*})$) and therefore $\\psi(a)=\\theta(a)$, using that $\\varphi_{1}(z)=\\psi(a)$ and $\\varphi_{2}(z^{*})=\\theta(a)$. We may summarize the results so far in the following theorem.\n\n\\begin{theorem}\nIf $\\varphi:\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{1}}\\rightarrow\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{2}}$ is a $\\Sigma_{\\textbf{C}}$-homomorphism, there exists a function $\\psi:\\mathcal{B}_{1}\\rightarrow\\mathcal{B}_{2}$ such that \n\\[\\varphi_{1}(z)=\\psi(z_{1})\\quad\\text{and}\\quad\\varphi_{2}(z)=\\psi(z_{2}),\\]\nfor all snapshots $z=(z_{1}, \\dotsc , z_{n+1})$ of $B_{n}^{\\mathcal{B}_{1}}$, and which satisfies, for all $a, b\\in \\mathcal{B}_{1}$ and $\\#\\in\\{\\vee, \\wedge, \\rightarrow\\}$, $\\psi(a\\#b)=\\psi(a)\\#\\psi(b)$.\n\\end{theorem}\n\nNow we will demand that $\\varphi:B_{n}^{\\mathcal{B}_{1}}\\rightarrow B_{n}^{\\mathcal{B}_{2}}$ be, not only a homomorphism from $\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{1}}$ to $\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{1}}$, but a morphism of the category $\\textbf{RSwap}_{C_{n}}$ from $\\mathcal{RM}_{C_{n}}^{\\mathcal{B}_{1}}$ to $\\mathcal{RM}_{C_{n}}^{\\mathcal{B}_{2}}$, so it also preserves designated elements and is absorbed by restricted valuations.\n\nSo, take a snapshot $z=(1, z_{2}, \\dotsc , z_{n+1})\\in D_{n}^{\\mathcal{B}_{1}}$: since we must have, by hypothesis over $\\varphi$, $\\varphi(z)\\in D_{n}^{\\mathcal{B}_{2}}$, and \n\\[\\varphi(z)=(\\varphi_{1}(z), \\dotsc , \\varphi_{n+1}(z))=(\\psi(1), \\psi(z_{2}), \\varphi_{3}(z), \\dotsc , \\varphi_{n+1}(z)),\\]\nwe obtain that $\\psi(1)=1$. Furthermore, for any formula $\\alpha$ in the language of $C_{n}$ and valuation $\\nu$ in $\\mathcal{F}_{C_{n}}^{\\mathcal{B}}$, Proposition \\ref{Value of strong negation} shows that $\\nu(\\alpha\\wedge\\neg\\alpha\\wedge\\alpha^{(n)})=F_{n}$; since $\\varphi$ is absorbed by restricted valuations, for any $\\nu\\in\\mathcal{F}_{C_{n}}^{\\mathcal{B}_{1}}$ we have $\\varphi\\circ\\nu\\in\\mathcal{F}_{C_{n}}^{\\mathcal{B}_{2}}$ and so, for any formula $\\alpha$ of $C_{n}$\n\\[F_{n}=\\varphi\\circ\\nu(\\alpha\\wedge\\neg\\alpha\\wedge\\alpha^{(n)})=\\varphi(F_{n})=(\\psi(0), \\psi(1), \\varphi_{3}(F_{n}), \\dotsc , \\varphi_{n+1}(F_{n})),\\]\nmeaning that we also have $\\psi(0)=0$. Since, for any $a, b\\in\\mathcal{B}_{1}$ and $\\#\\in\\{\\vee, \\wedge, \\rightarrow\\}$, $\\psi(a\\#b)=\\psi(a)\\#\\psi(b)$, $\\psi(1)=1$ and $\\psi(0)=0$, we actually have that $\\psi$ is a homomorphism of Boolean algebras, as we had previously promised: given that ${\\sim}a=a\\rightarrow 0$, we have \n\\[\\psi({\\sim}a)=\\psi(a\\rightarrow 0)=\\psi(a)\\rightarrow\\psi(0)={\\sim}\\psi(a).\\]\n\nBut the fact that $\\varphi$ is absorbed by restricted valuations allows to prove something even stronger: remember that, for any $\\nu\\in\\mathcal{F}^{\\mathcal{B}}_{C_{n}}$ and formula $\\alpha$ of $C_{n}$, $\\nu(\\alpha^{k})_{1}=\\nu(\\alpha)_{k+2}$ for $1\\leq k\\leq n-1$, what we proved back in Lemma \\ref{Finding the coordinates} for $2\\leq k\\leq n-1$ and is obvious by the definition of $\\mathcal{F}_{C_{n}}^{\\mathcal{B}}$ for $k=1$. Given a $z$ in $B_{n}^{\\mathcal{B}_{1}}$, for any propositional variable $p$ we can find a restricted valuation in $\\mathcal{F}_{C_{n}}^{\\mathcal{B}_{1}}$ with $\\nu(p)=z$, by proceeding as in Section \\ref{RM is not trivial} if necessary; so $\\nu(p^{k})_{1}=z_{k+2}$.\n\nSince $\\varphi\\circ\\nu$ must be in $\\mathcal{F}_{C_{n}}^{\\mathcal{B}_{2}}$, we have that\n\\[\\varphi\\circ\\nu(p^{k})_{1}=\\varphi\\circ\\nu(p)_{k+2}=\\varphi(z)_{k+2}=\\varphi_{k+2}(z),\\]\nand at the same time,\n\\[\\varphi\\circ\\nu(p^{k})_{1}=\\psi(\\nu(p^{k})_{1})=\\psi(z_{k+2}),\\]\nimplying that $\\varphi_{k+2}(z)=\\psi(z_{k+2})$, for every $1\\leq k\\leq n-1$. We get then the following theorem.\n\n\\begin{theorem}\\label{Classifying morphisms in RSwap}\nIf $\\varphi:\\mathcal{RM}_{C_{n}}^{\\mathcal{B}_{1}}\\rightarrow\\mathcal{RM}_{C_{n}}^{\\mathcal{B}_{2}}$ is a morphism of $\\textbf{RSwap}_{C_{n}}$, there exists a homomorphism of Boolean algebras $\\psi:\\mathcal{B}_{1}\\rightarrow\\mathcal{B}_{2}$ such that, for all snapshots $z\\in B_{n}^{\\mathcal{B}_{1}}$,\n\\[\\varphi_{i}(z)=\\psi(z_{i}),\\quad\\text{for all}\\quad1\\leq i\\leq n+1.\\]\n\\end{theorem}\n\n\n\n\n\n\n\n\n\n\n\\subsection{$\\textbf{BA}$ and $\\textbf{RSwap}_{C_{n}}$ are isomorphic}\\label{BA and RSwap}\n\n\\begin{proposition}\nThe set $Boo_{n}^{\\mathcal{B}}$ of Boolean elements of $\\mathcal{A}_{C_{n}}^{\\mathcal{B}}$ is a Boolean algebra isomorphic to $\\mathcal{B}$ when we define, for $\\#\\in\\{\\vee, \\wedge, \\rightarrow\\}$, and $a$ and $b$ elements of $\\mathcal{B}$,\n\\[(a, {\\sim}a, 1, \\dotsc , 1)\\#(b, {\\sim}b, 1, \\dotsc , 1)=(a\\#b, {\\sim}(a\\#b), 1, \\dotsc , 1),\\]\n\\[{\\sim}(a, {\\sim}a, 1, \\dotsc , 1)=({\\sim}a, a, 1, \\dotsc , 1),\\]\n$\\top=(1, 0, 1, \\dotsc , 1)$ and $\\bot=(0, 1, 1, \\dotsc , 1)$\n\\end{proposition}\n\n\\begin{proof}\nConsider the map $\\rho:B\\rightarrow Boo_{n}^{\\mathcal{B}}$, for $B$ the universe of $\\mathcal{B}$, given by $\\rho(a)=$\\\\$(a, {\\sim}a, 1, \\dotsc , 1)$. Then, for $a, b\\in B$:\n\\begin{enumerate}\n\\item for $\\#\\in\\{\\vee, \\wedge, \\rightarrow\\}$, \n\\[\\rho(a\\#b)=(a\\#b, {\\sim}(a\\#b), 1, \\dotsc , 1)=(a, {\\sim}a, 1, \\dotsc , 1)\\#(b, {\\sim}b, 1, \\dotsc , 1)=\\rho(a)\\#\\rho(b);\\]\n\\item $\\rho({\\sim}a)=({\\sim}a, {\\sim}{\\sim}a, 1, \\dotsc , 1)=({\\sim}a, a, 1, \\dotsc , 1)={\\sim}(a, {\\sim}a, 1, \\dotsc , 1)={\\sim}\\rho(a)$;\n\\item $\\rho(\\top)=(1, 0, 1, \\dotsc , 1)=\\top$;\n\\item $\\rho(\\bot)=(0, 1, 1, \\dotsc , 1)=\\bot$.\n\\end{enumerate}\nFurthermore, $\\rho$ is clearly both injective and surjective, being therefore an isomorphism.\n\\end{proof}\n\nNotice that the operations we give to $Boo_{n}^{\\mathcal{B}}$ are the natural choice of Boolean algebra structure for this set, taking into consideration the multioperations on $\\mathcal{A}_{C_{n}}^{\\mathcal{B}}$, since for elements $a$ and $b$ of $\\mathcal{B}$, and $\\#\\in\\{\\vee, \\wedge, \\rightarrow\\}$, \n\\[\\rho(a)\\tilde{\\#}\\rho(b)=\\{\\rho(a)\\#\\rho(b)\\}\\quad\\text{and}\\quad\\tilde{\\neg}\\rho(a)=\\{{\\sim}\\rho(a)\\}.\\]\n\nConsider now the category $\\textbf{BA}$\\label{BA} of non-degenerate Boolean algebras\\index{Boolean algebra, Non-degenerate} (that is, Boolean algebras where $0\\neq 1$), with homomorphisms of Boolean algebras as morphisms, and the functors $\\mathcal{A}_{n}:\\textbf{BA}\\rightarrow\\textbf{RSwap}_{C_{n}}$\\label{An}, which takes\n\\begin{enumerate}\n\\item a Boolean algebra $\\mathcal{B}$ to $\\mathcal{A}_{n}\\mathcal{B}=\\mathcal{A}_{C_{n}}^{\\mathcal{B}}$;\n\\item a homomorphism of Boolean algebras $\\psi:\\mathcal{B}_{1}\\rightarrow\\mathcal{B}_{2}$ to the morphism $\\mathcal{A}_{n}\\psi:\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{1}}\\rightarrow\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{2}}$ such that, for $z=(z_{1}, \\dotsc , z_{n+1})\\in B_{n}^{\\mathcal{B}_{1}}$, $\\mathcal{A}_{n}\\psi(z)_{i}=\\psi(z_{i})$, for every $i\\in\\{1, \\dotsc , n+1\\}$;\n\\end{enumerate}\nand $\\textbf{Boo}_{n}$\\label{tBoon}, which takes\n\\begin{enumerate}\n\\item the restricted swap structure $\\mathcal{A}_{C_{n}}^{\\mathcal{B}}$ to the Boolean algebra $\\mathcal{B}$ (or, what we saw to be equivalent, $Boo_{n}^{\\mathcal{B}}$ with its natural Boolean algebra structure);\n\\item a morphism $\\varphi:\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{1}}\\rightarrow\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{2}}$ to the homomorphism of Boolean algebras $\\textbf{Boo}_{n}\\varphi:\\mathcal{B}_{1}\\rightarrow\\mathcal{B}_{2}$ such that, for $a$ an element of $\\mathcal{B}_{1}$, $\\textbf{Boo}_{n}\\varphi(a)=\\varphi((a, {\\sim}a, 1, \\dotsc , 1))_{1}$.\n\\end{enumerate}\n\n\\begin{proposition}\nAs described, $\\mathcal{A}_{n}$ is, indeed, a functor.\n\\end{proposition}\n\n\\begin{proof}\nAs we proved in Theorem \\ref{Classifying morphisms in RSwap}, for a homomorphism of Boolean algebras $\\psi:\\mathcal{B}_{1}\\rightarrow\\mathcal{B}_{2}$, $\\mathcal{A}_{n}\\psi:B_{n}^{\\mathcal{B}_{1}}\\rightarrow B_{n}^{\\mathcal{B}_{2}}$ such that, for an arbitrary $z=(z_{1}, \\dotsc , z_{n+1})\\in B_{n}^{\\mathcal{B}_{1}}$, $\\mathcal{A}_{n}\\psi(z)_{i}=\\psi(z_{i})$, for $1\\leq i\\leq n+1$, is indeed a morphism in $\\textbf{RSwap}_{C_{n}}$.\n\nNow, for a second homomorphism of Boolean algebras $\\theta:\\mathcal{B}_{2}\\rightarrow\\mathcal{B}_{3}$, one has\n\\[\\mathcal{A}_{n}(\\theta\\circ\\psi)(z)=(\\theta\\circ\\psi(z_{1}), \\dotsc , \\theta\\circ\\psi(z_{n+1}))=\\mathcal{A}_{n}\\theta((\\psi(z_{1}), \\dotsc , \\psi(z_{n+1}))=\\]\n\\[\\mathcal{A}_{n}\\theta\\circ\\mathcal{A}_{n}\\psi(z);\\]\nfurthermore, for the identity $Id_{\\mathcal{B}}:\\mathcal{B}\\rightarrow\\mathcal{B}$ of $\\mathcal{B}$, $\\mathcal{A}_{n}Id_{\\mathcal{B}}:\\mathcal{A}_{C_{n}}^{\\mathcal{B}}\\rightarrow\\mathcal{A}_{C_{n}}^{\\mathcal{B}}$ satisfies that, for $z\\in B_{n}^{\\mathcal{B}}$, \n\\[\\mathcal{A}_{n}Id_{\\mathcal{B}}(z)=(Id_{\\mathcal{B}}(z_{1}), \\dotsc , Id_{\\mathcal{B}}(z_{n+1}))=(z_{1}, \\dotsc , z_{n+1})=z,\\]\nand therefore is precisely the identity on $\\mathcal{A}_{C_{n}}^{\\mathcal{B}}$.\n\\end{proof}\n\n\\begin{proposition}\nAs described, $\\textbf{Boo}_{n}$ is, indeed, a functor.\n\\end{proposition}\n\n\\begin{proof}\nSince, for any morphism $\\varphi:\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{1}}\\rightarrow\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{2}}$, there exists a homomorphism $\\psi:\\mathcal{B}_{1}\\rightarrow\\mathcal{B}_{2}$ such that, for any $z=(z_{1}, \\dotsc , z_{n+1})\\in B_{n}^{\\mathcal{B}_{1}}$, $\\varphi(z)_{i}=\\psi(z_{i})$, for any $1\\leq i\\leq n+1$, one sees that \n\\[\\textbf{Boo}_{n}\\varphi(a)=\\varphi((a, {\\sim}a, 1, \\dotsc , 1))_{1}=\\psi(a),\\]\nand so $\\textbf{Boo}_{n}\\varphi$ is indeed a homomorphism of Boolean algebras.\n\nFor a second morphism $\\eta:\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{2}}\\rightarrow \\mathcal{A}_{C_{n}}^{\\mathcal{B}_{3}}$ and a homomorphism $\\theta:\\mathcal{B}_{2}\\rightarrow\\mathcal{B}_{3}$ such that, for every $w\\in B_{n}^{\\mathcal{B}_{2}}$, $\\eta(w)_{i}=\\theta(w_{i})$, consider an element $a$ of $\\mathcal{B}_{1}$: then\n\\[\\textbf{Boo}_{n}(\\eta\\circ\\varphi)(a)=\\eta(\\varphi((a, {\\sim}a, 1, \\dotsc , 1)))_{1}=\\eta_{1}(\\varphi((a, {\\sim}a, 1, \\dotsc , 1)))=\\]\n\\[\\theta(\\varphi((a, {\\sim}a, 1, \\dotsc , 1))_{1})=\\theta(\\varphi_{1}((a, {\\sim}a, 1, \\dotsc , 1)))=\\theta(\\psi(a))=\\theta(\\textbf{Boo}_{n}\\varphi(a))=\\]\n\\[\\textbf{Boo}_{n}\\eta\\circ\\textbf{Boo}_{n}\\varphi(a).\\]\n\nFinally, consider the identity homomorphism $Id_{\\mathcal{A}_{C_{n}}^{\\mathcal{B}}}:\\mathcal{A}_{C_{n}}^{\\mathcal{B}}\\rightarrow\\mathcal{A}_{C_{n}}^{\\mathcal{B}}$: for an $a$ in $\\mathcal{B}$,\n\\[\\textbf{Boo}_{n}Id_{\\mathcal{A}_{C_{n}}^{\\mathcal{B}}}(a)=Id_{\\mathcal{A}_{C_{n}}^{\\mathcal{B}}}((a, {\\sim}a, 1, \\dotsc , 1))_{1}=a,\\]\nproving $\\textbf{Boo}_{n}Id_{\\mathcal{A}_{C_{n}}^{\\mathcal{B}}}$ is the identity on $\\mathcal{B}$.\n\\end{proof}\n\n\\begin{theorem}\n$\\textbf{Boo}_{n}\\circ\\mathcal{A}_{n}=Id_{\\textbf{BA}}$.\n\\end{theorem}\n\n\\begin{proof}\nA Boolean algebra $\\mathcal{B}$ is taken by $\\mathcal{A}_{n}$ into $\\mathcal{A}_{C_{n}}^{\\mathcal{B}}$, and $\\mathcal{A}_{C_{n}}^{\\mathcal{B}}$ is taken to $\\mathcal{B}$ again by $\\textbf{Boo}_{n}$. This means $\\textbf{Boo}_{n}\\circ\\mathcal{A}_{n}$ is the identity on objects.\n\nNow, for Boolean algebras $\\mathcal{B}_{1}$ and $\\mathcal{B}_{2}$, a homomorphism $\\psi:\\mathcal{B}_{1}\\rightarrow\\mathcal{B}_{2}$, and an element $a$ of $\\mathcal{B}_{1}$, let $\\varphi=\\mathcal{A}_{n}\\psi$:\n\\[(\\textbf{Boo}_{n}\\circ\\mathcal{A}_{n})\\psi(a)=\\textbf{Boo}_{n}\\varphi(a)=\\varphi((a, {\\sim}a, 1, \\dotsc , 1))_{1}=\\psi(a),\\]\nimplying $\\textbf{Boo}_{n}\\circ\\mathcal{A}_{n}$ is also identical when applied to morphisms.\n\\end{proof}\n\n\\begin{theorem}\n$\\mathcal{A}_{n}\\circ\\textbf{Boo}_{n}=Id_{\\textbf{RSwap}_{C_{n}}}$.\n\\end{theorem}\n\n\\begin{proof}\n$\\mathcal{A}_{C_{n}}^{\\mathcal{B}}$ is taken by $\\textbf{Boo}_{n}$ to $\\mathcal{B}$, which is taken back to $\\mathcal{A}_{C_{n}}^{\\mathcal{B}}$ by $\\mathcal{A}_{n}$, giving us the identity on objects.\n\nNow, for restricted swap structures $\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{1}}$ and $\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{2}}$, a morphism $\\varphi:\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{1}}\\rightarrow\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{2}}$ and an element $z=(z_{1}, \\dotsc , z_{n+1})\\in B_{n}^{\\mathcal{B}_{1}}$, let us denote $\\textbf{Boo}_{n}\\varphi$ by $\\psi$: \n\\[\\mathcal{A}_{n}\\circ\\textbf{Boo}_{n}\\varphi(z)=\\mathcal{A}_{n}\\psi(z)=(\\psi(z_{1}), \\dotsc , \\psi(z_{n+1}));\\]\nsince $\\textbf{Boo}_{n}\\varphi=\\psi$, for any $a$ in $\\mathcal{B}_{1}$, $\\varphi((a, {\\sim}a, 1, \\dotsc , 1))_{1}=\\psi(a)$. We also know that there exists a homomorphism $\\theta:\\mathcal{B}_{1}\\rightarrow\\mathcal{B}_{2}$ such that, for any $z\\in B_{n}^{\\mathcal{B}_{1}}$, $\\varphi(z)_{i}=\\theta(z_{i})$ for $1\\leq i\\leq n+1$.\n\nFrom the fact that $\\theta(a)=\\varphi((a, {\\sim}a, 1, \\dotsc , 1))_{1}=\\psi(a)$, we obtain $\\theta=\\psi$, and so\n\\[\\varphi(z)=(\\psi(z_{1}), \\dotsc , \\psi(z_{n+1}))=\\mathcal{A}_{n}\\circ\\textbf{Boo}_{n}\\varphi(z),\\]\nimplying finally that $\\mathcal{A}_{n}\\circ\\textbf{Boo}_{n}$ maintains morphisms fixed as well.\n\\end{proof}\n\nNow that we have $\\textbf{RSwap}_{C_{n}}$ is isomorphic to $\\textbf{BA}$, we may use this isomorphism to bring some results from the latter category into the former: of course, $\\textbf{BA}$ is one of the better known categories, so there are many such results. Let us start by reminding that every atomic, complete Boolean algebra is isomorphic to some power of the two-valued Boolean algebra $\\textbf{2}$; in particular, an atomic, complete Boolean algebra with $\\kappa$ atoms is isomorphic to precisely $\\textbf{2}^{\\kappa}$, where $\\kappa$ may be a finite or infinite cardinal.\n\nSince every finite Boolean algebra is forcibly both atomic and complete, and from the results of Section \\ref{Counting snapshots} $\\mathcal{RM}_{C_{n}}^{\\mathcal{B}}$ is finite if and only if $\\mathcal{B}$ is finite, we find a first representation result in $\\textbf{RSwap}_{C_{n}}$.\n\n\\begin{corollary}\nEvery finite $\\mathcal{RM}_{C_{n}}^{\\mathcal{B}}$ is isomorphic to a power of $\\mathcal{RM}_{C_{n}}$.\n\\end{corollary}\n\nAnother representation theorem on Boolean algebras, a stronger one, states that every Boolean algebra is isomorphic to a field of sets, that is, a Boolean subalgebra of the powerset Boolean algebra of a given set; since every powerset is a complete and atomic Boolean algebra, and is therefore isomorphic to some power $\\textbf{2}^{\\kappa}$ of $\\textbf{2}$, it follows that every Boolean algebra is isomorphic to a subalgebra of a power of $\\textbf{2}$.\n\nNow, we still haven't defined what it means for an RNmatrix to be a \"subRNmatrix\" of another, although we could merely translate what this means from $\\textbf{BA}$ into $\\textbf{RSwap}_{C_{n}}$; we chose to give a more general approach instead. So, given RNmatrices $\\mathcal{M}=(\\mathcal{A}, D, \\mathcal{F})$ and $\\mathcal{M}^{\\prime}=(\\mathcal{A}^{\\prime}, D^{\\prime}, \\mathcal{F}^{\\prime})$ over the same signature, we say that $\\mathcal{M}^{\\prime}$ is a subRNmatrix\\index{SubRNmatrix} of $\\mathcal{M}$ if:\n\\begin{enumerate}\n\\item $\\mathcal{A}^{\\prime}$ is a submultialgebra of $\\mathcal{A}$;\n\\item $D^{\\prime}$ is a subset of $D$, and\n\\item $\\{j\\circ\\nu: \\nu\\in\\mathcal{F}^{\\prime}\\}\\subseteq \\mathcal{F}$, for $j$ the inclusion of the universe of $\\mathcal{A}^{\\prime}$ into the universe of $\\mathcal{A}$.\n\\end{enumerate}\nIn other words, $\\mathcal{M}^{\\prime}$ is a subRNmatrix of $\\mathcal{M}$ if $j$ is a morphism on the underlying category of RNmatrices.\n\nNow, we state that in $\\textbf{RSwap}_{C_{n}}$ the concept just described of a subRNmatrix corresponds exactly to what one would obtain by translating the notion of being a subalgebra from $\\textbf{BA}$, that is, $\\mathcal{RM}_{C_{n}}^{\\mathcal{B}_{1}}$ is a subRNmatrix of $\\mathcal{RM}_{C_{n}}^{\\mathcal{B}_{2}}$ if and only if $\\mathcal{B}_{1}$ is a subalgebra of $\\mathcal{B}_{2}$.\n\nIn one direction, by assuming that $\\mathcal{RM}_{C_{n}}^{\\mathcal{B}_{1}}$ is a subRNmatrix of $\\mathcal{RM}_{C_{n}}^{\\mathcal{B}_{2}}$, we have that $\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{1}}$ is a submultialgebra of $\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{2}}$, so for every $a\\in\\mathcal{B}_{1}$, given that $(a, {\\sim}a, 1, \\dotsc , 1)$ is an element of $B_{n}^{\\mathcal{B}_{1}}$, we have that it must also be an element of $B_{n}^{\\mathcal{B}_{2}}$, and therefore $a\\in\\mathcal{B}_{2}$. Since, for any $a, b\\in\\mathcal{B}_{1}$ and $\\#\\in\\{\\vee, \\wedge, \\rightarrow\\}$, \n\\[(a, {\\sim}a, 1, \\dotsc , 1)\\tilde{\\#}(b, {\\sim}b, 1, \\dotsc , 1)=\\{(a\\#b, {\\sim}(a\\#b), 1, \\dotsc , 1)\\]\nand $\\tilde{\\neg}(a, {\\sim}a, 1, \\dotsc , 1)=\\{({\\sim}a, a, 1, \\dotsc , 1)\\}$, and given that the operations $\\tilde{\\#}$ and $\\tilde{\\neg}$ in $\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{1}}$ are the same as those in $\\mathcal{A}_{C_{n}}^{\\mathcal{B}_{2}}$, we obtain that the operations in $\\mathcal{B}_{1}$ are the same as those in $\\mathcal{B}_{2}$, proving the former is a Boolean subalgebra of the latter.\n\nReciprocally, suppose $\\mathcal{B}_{1}$ is a Boolean subalgebra of $\\mathcal{B}_{2}$: given a snapshot $z=(z_{1}, \\dotsc , z_{n+1})\\in B_{n}^{\\mathcal{B}_{1}}$, we have that, fist of all, $z_{1}, \\dotsc , z_{n+1}\\in \\mathcal{B}_{1}$, implying that $z_{1}, \\dotsc , z_{n+1}\\in \\mathcal{B}_{2}$; second, since $(\\bigwedge_{i=1}^{k}z_{i})\\vee z_{k+1}=1$, for all $1\\leq k\\leq n$, in $\\mathcal{B}_{1}$, and since the operations in $\\mathcal{B}_{2}$, when restricted to the universe of $\\mathcal{B}_{1}$, coincide to the operations in $\\mathcal{B}_{2}$, we obtain that $(\\bigwedge_{i=1}^{k}z_{i})\\vee z_{k+1}=1$, now in $\\mathcal{B}_{2}$. With this, $z$ is a snapshot of $B_{n}^{\\mathcal{B}_{2}}$.\n\nSo we can consider the inclusion $j:B_{n}^{\\mathcal{B}_{1}}\\rightarrow B_{n}^{\\mathcal{B}_{2}}$, and it is easy to prove that it is a morphism of $\\textbf{RSwap}_{C_{n}}$: after all, it can be written as $j(z)=(i(z_{1}), \\dotsc , i(z_{n+1}))$ for $i:\\mathcal{B}_{1}\\rightarrow\\mathcal{B}_{2}$ the inclusion homomorphism, and $z=(z_{1}, \\dotsc , z_{n+1})$ an arbitrary snapshot in $B_{n}^{\\mathcal{B}_{1}}$.\n\n\\begin{corollary}\nEvery $\\mathcal{RM}_{C_{n}}^{\\mathcal{B}}$ is a subRNmatrix of a power of $\\mathcal{RM}_{C_{n}}$.\n\\end{corollary}\n\n\\newpage\n\\printbibliography[segment=\\therefsegment,heading=subbibliography]\n\\end{refsegment}\n\n\n\n\n\n\\begin{refsegment}\n\\defbibfilter{notother}{not segment=\\therefsegment}\n\\setcounter{chapter}{6}\n\\chapter{Logics of formal incompatibility}\\label{Chapter7}\\label{Chapter 7}\n\nIn classical logics, a formula and its negation are not compatible, in the sense that having both $\\alpha$ and ${\\sim} \\alpha$ to be true trivialize whatever argument we are working over.\n\nWhen dealing with logics of formal inconsistency, this is no longer true: we can have $\\alpha$ and its negation $\\neg\\alpha$ without trivializing our logic, as long as $\\alpha$ is inconsistent, that is, $\\circ\\alpha$ is not true, when we have at our disposal the consistency connective \"$\\circ$\".\n\nTo formalize such a notion of incompatibility, we will consider a binary connective that, when connecting formulas $\\alpha$ and $\\beta$, will stand intuitively for \"$\\alpha$ is incompatible with $\\beta$\". When choosing a symbol for such connective, one somewhat natural choice is the Sheffer's stroke\\index{Sheffer's stroke}: in classical propositional logic, the Sheffer's stroke may be defined from the usual connectives as\n\\[\\alpha\\Uparrow\\beta={\\sim}(\\alpha\\wedge\\beta),\\]\nand of course having $\\alpha\\Uparrow\\beta$ to hold, along with $\\alpha$ and $\\beta$, trivializes an argument. The basic axiom we will expect a system for incompatibility to satisfy will be\\index{Incompatibility}\\label{uparrow}\n\\[(\\alpha\\uparrow\\beta)\\rightarrow(\\alpha\\rightarrow(\\beta\\rightarrow\\gamma)),\\]\nfor any formula $\\gamma$, or more generally, if we don't have a deduction meta-theorem,\n\\[\\alpha\\uparrow\\beta, \\alpha, \\beta\\vdash_{\\mathscr{L}}\\gamma;\\]\nin words, that means that having $\\alpha$ and $\\beta$ to be true while having $\\alpha$ and $\\beta$ to be incompatible imply that our logic is trivial.\n\nReferring back to paraconsistency, one sees consistency may be characterized as incompatibility: $\\alpha$ is consistent if, and only if, is incompatible with $\\neg\\alpha$. But, given a logic dealing with incompatibility, we are also tempted to say that any formula $\\beta$ which is incompatible with a given $\\alpha$ is a negation of $\\alpha$, which leads to a notion of consistency and back again to logics having inconsistency in their scope.\n\nBefore we go any further, it is important to formally define with which structures we are working. We define the signature $\\Sigma_{\\bI}$\\label{SigmabI} by: $(\\Sigma_{\\bI})_{2}=\\{\\vee, \\wedge, \\rightarrow, \\uparrow\\}$, and $(\\Sigma_{\\bI})_{n}=\\emptyset$ for $n\\neq 2$.\n\n\\begin{definition}\\label{Logics of formal incompatibility}\nA logic $\\mathcal{L}$ is said to be a logic of formal incompatibility\\index{Logic of formal incompatibility}, or in short $\\textbf{LIp}$\\label{LIp}, if it contains a family of formulas $\\Uuparrow(p,q)$, exactly on the variables $p$ and $q$, such that there exist formulas $\\varphi$, $\\phi$ and $\\psi$ for which\n\\[\\varphi, \\phi\\not\\vdash\\psi;\\]\nformulas $\\alpha$, $\\beta$ and $\\gamma$ such that:\n\\begin{enumerate}\n\\item $\\Uuparrow(\\alpha, \\beta), \\alpha\\not\\vdash\\gamma$ and\n\\item $\\Uuparrow(\\alpha, \\beta), \\beta\\not\\vdash\\gamma$,\n\\end{enumerate}\nwhere $\\Uuparrow(\\alpha, \\beta)$ is the set obtained by replacing, in each formula of $\\Uuparrow(p,q)$, $p$ by $\\alpha$ and $q$ by $\\beta$; and, for all formulas $\\gamma$, $\\theta$ and $\\omega$,\n\\[\\Uuparrow(\\gamma, \\theta), \\gamma, \\theta\\vdash\\omega.\\]\n\\end{definition}\n\nMore often than not, we want $\\Uuparrow(p,q)$ to be composed of only one formula, given by a primitive, binary connective evaluated on $p$ and $q$ for which we will use the infix notation, that is, $p\\uparrow q$.\n\nNow, one can ask oneself when incompatibility has an interpretation in natural logics, and perhaps the most canonical example would be the one of limited resources\\index{Limited resources}. Suppose we are given a basis of true statements, $\\Gamma$, and formulas $\\varphi$ that must be tested against $\\Gamma$ in the most resource-efficient way possible, in regards to both time and amount of information recorded.\n\nNow, if $\\Gamma\\vdash {\\sim} \\varphi$, we must discard $\\varphi$: but, if the proof of this implication is unbelievably long, we waste precious time. One could, to avoid spending time unnecessarily, have a vast table pre-establishing if $\\varphi$ follows from $\\Gamma$ or not, but this would, of course, be memory-consuming, and in a way also time-consuming: without a strong algorithm to check the list, this could prove to be a very long search to identify $\\varphi$ or ${\\sim} \\varphi$ on said list.\n\nThe solution could lie halfway between those two approaches: one could have a small list of statements that do not follow from $\\Gamma$, which are incompatible with $\\Gamma$, and, given a $\\varphi$, attempt to prove or disprove the given formula, knowing a few shortcuts given by our list of incompatible statements.\n\nThis is very common not only in computer science, but also in science in general, mathematics, and logic: one does not completely prove a statement starting from the axioms, but rather accept a few results, derived from previous work, as true, and therefore accept their negations as incompatible with whatever result is being searched for, and proceed from there; it is also useful, in this context, to consider compatibility\\index{Compatibility} of two formulas $\\alpha$ and $\\beta$ rather than incompatibility, defined trivially as $\\alpha\\downarrow\\beta={\\sim}(\\alpha\\uparrow\\beta)$\\label{downarrow}.\n\nAnother useful application of incompatibility is when dealing with partial information\\index{Partial information}: suppose one must test $\\varphi$ against a basis of true statements $\\Delta$, without having direct access to $\\Delta$ but rather $\\Gamma\\subset \\Delta$; in this case, knowing some key incompatibilities between $\\varphi$ and $\\Delta$ or $\\varphi$ and the complement of $\\Delta$ may help deriving $\\Delta\\vdash \\varphi$ or $\\Delta\\vdash{\\sim} \\varphi$ while only using directly $\\Gamma$.\n\nPerhaps more importantly, incompatibility has an obvious connection to probability, specifically independence (for what is to come, any reference in probability logic is sufficient, such as \\cite{ProbabilityLogic}): consider a sample space $X$, that is, the set of all things that may occur in ones analysis; if, to give one example, we were to look at the outcome of flipping a coin, our sample space could be $X=\\{heads, tails\\}$. The next step, when dealing with probabilities, is taking a $\\sigma$-algebra\\footnote{A $\\sigma$-algebra $\\mathcal{A}$ on a set $X$ is a collection of subsets of $X$ containing $X$ itself and closed under complements and countable unions, meaning that if $A\\in \\mathcal{A}$ and $\\{A_{n}\\}_{n\\in\\mathbb{N}}\\subseteq\\mathcal{A}$, then $X\\setminus A$ and $\\bigcup_{n\\in\\mathbb{N}}A_{n}$ are both in $\\mathcal{A}$.} $\\mathcal{A}$ of subsets of $X$, most commonly the whole powerset of $X$. Finally, we also need a probability measure $P$ on $\\mathcal{A}$ to obtain a probability space $(X, \\mathcal{A}, P)$, that is, a function from $\\mathcal{A}$ into the interval $[0,1]$ satisfying $P(X)=1$ and \n\\[P(\\bigcup_{n\\in\\mathbb{N}}A_{n})=\\sum_{n\\in\\mathbb{N}}P(A_{n}),\\]\nfor $\\{A_{n}\\}_{n\\in\\mathbb{N}}$ a family of pairwise disjoint sets of $\\mathcal{A}$. Intuitively, $P$ tells us the probability of an event (\\textit{i. e.} element of $\\mathcal{A}$) happening, and $\\mathcal{A}$ tells us to which events a probability can actually be assigned. We say two events $A$ and $B$ are mutually exclusive whenever \n\\[\\text{$P(A\\cap B)=0$ or, what is equivalent, $P(A\\cup B)=P(A)+P(B)$;}\\]\nintuitively, mutually exclusive events are those events that cannot both occur in an experiment. Then, if a propositional variable $p$ is associated to an event $A$, and a second propositional variable is associated to an event $B$, it is very natural to interpret the incompatibility $p\\uparrow q$ of $p$ and $q$ as the mutual exclusivity of the events $A$ and $B$, in which case the axiom $(\\alpha\\uparrow\\beta)\\rightarrow(\\alpha\\rightarrow (\\beta\\rightarrow\\gamma))$ clearly models the fact that both events $A$ and $B$ can not simultaneously occur if the two are independent.\n\nThis seems like a specially fruitful approach to apply to probability logics, systems used in the formal study of probability theory and statistics, as well as in Bayesian perspectives on epistemology of science. But, still on the field of probability theory, we can think of an alternative interpretation of incompatibility: two events $A$ and $B$ are independent if \n\\[P(A\\cap B)=P(A)P(B);\\]\nindependent events naturally arise from the fact that one can usually only derive more information about the probability of a complex event by knowing that its constituent events are pairwise independent, sometimes an even stronger independence condition being necessary. One could, then, for variables $p$ and $q$ standing for, respectively, events $A$ and $B$, interpret $p\\uparrow q$ as the independence of $A$ and $B$, what makes sense at first glance due to the binary nature of the two concepts. However, in this interpretation, the axiom $(\\alpha\\uparrow\\beta)\\rightarrow(\\alpha\\rightarrow (\\beta\\rightarrow\\gamma))$ is no longer the most desirable one, at least in a naive interpretation of implication, given that two independent events can, and often do, simultaneously happen.\n\nMost of the research developed in this chapter can be found as a preprint in \\cite{Frominconsistency}.\n\n\n\n\n\\section{The logic $\\bI$}\\label{Defining bI}\n\nOur simplest $\\textbf{LIp}$, which we shall denote by $\\bI$\\label{bI}, satisfies the axiom schemata of the positive fragment of classical propositional logic,\n\n\\begin{enumerate}\n\\item[\\textbf{Ax\\: 1}] $\\alpha\\rightarrow(\\beta\\rightarrow\\alpha)$;\n\\item[\\textbf{Ax\\: 2}] $\\big(\\alpha\\rightarrow (\\beta\\rightarrow \\gamma)\\big)\\rightarrow\\big((\\alpha\\rightarrow\\beta)\\rightarrow(\\alpha\\rightarrow\\gamma)\\big)$;\n\\item[\\textbf{Ax\\: 3}] $\\alpha\\rightarrow\\big(\\beta\\rightarrow(\\alpha\\wedge\\beta)\\big)$;\n\\item[\\textbf{Ax\\: 4}] $(\\alpha\\wedge\\beta)\\rightarrow \\alpha$;\n\\item[\\textbf{Ax\\: 5}] $(\\alpha\\wedge\\beta)\\rightarrow \\beta$;\n\\item[\\textbf{Ax\\: 6}] $\\alpha\\rightarrow(\\alpha\\vee\\beta)$;\n\\item[\\textbf{Ax\\: 7}] $\\beta\\rightarrow(\\alpha\\vee\\beta)$;\n\\item[\\textbf{Ax\\: 8}] $(\\alpha\\rightarrow\\gamma)\\rightarrow\\Big((\\beta\\rightarrow\\gamma)\\rightarrow \\big((\\alpha\\vee\\beta)\\rightarrow\\gamma\\big)\\Big)$;\n\\item[$\\textbf{Ax\\: 9}^{*}$] $(\\alpha\\rightarrow \\beta)\\vee\\alpha$,\n\\end{enumerate}\nplus\\label{Ip}\n\\[\\tag{\\textbf{Ip}}(\\alpha\\uparrow\\beta)\\rightarrow(\\alpha\\rightarrow(\\beta\\rightarrow \\gamma))\\]\nand\\label{Comm}\n\\[\\tag{\\textbf{Comm}}(\\alpha\\uparrow\\beta)\\rightarrow(\\beta\\uparrow\\alpha),\\]\nand follows the inference rule of Modus Ponens.\n\nCompare it with the classical definition of $\\textbf{mbC}$ and its extensions: they are logics over $\\Sigma_{\\textbf{LFI}}$, which is simply $\\Sigma_{\\bI}$ once we exchange $\\uparrow$ for $\\circ$ and add a negation, whose Hilbert calculus consists at least of the axiom schemata for the positive fragment of propositional logic, excluded middle and the schema\n\\[\\tag{\\textbf{bc1}}\\circ\\alpha\\rightarrow(\\alpha\\rightarrow({\\sim} \\alpha\\rightarrow \\beta)),\\]\nwith Modus Ponens as inference rule. The main difference here is the presence of $\\textbf{Comm}$, standing for the commutativity\\index{Commutativity of incompatibility} of the connective $\\uparrow$. For convenience, we will denote the logic $\\bI$ without $\\textbf{Comm}$ by $\\bI^{-}$\\label{bI-}.\n\n\\begin{example}\nWe now present a model of $\\bI$: take the signature $\\Sigma_{\\bI}$ and let $\\textbf{CPL}$ be the classical propositional logic; we make $\\textbf{CPL}$ into a model $\\textbf{CPL}_{\\uparrow}$ of $\\bI$ by defining \n\\[\\alpha\\uparrow\\beta\\quad\\text{to be}\\quad\\alpha\\Uparrow\\beta={\\sim}(\\alpha\\wedge\\beta),\\]\nthat is, $\\uparrow$ is the classically defined Sheffer's stroke. Clearly $\\textbf{CPL}_{\\uparrow}$ satisfies the axioms of the positive fragment of classical propositional logic and Modus Ponens, remaining for us to show that it also validates $\\textbf{Ip}$ and $\\textbf{Comm}$.\n\nThis is easy since, if we have $\\alpha\\uparrow\\beta$, $\\alpha$ and $\\beta$, and $\\alpha\\uparrow\\beta$ means that ${\\sim}(\\alpha\\wedge\\beta)$, we have, by use of the deduction meta-theorem for classical propositional logic, that $(\\alpha\\wedge\\beta)\\wedge{\\sim}(\\alpha\\wedge\\beta)$, which by the explosivity of negation in $\\textbf{CPL}$ implies any $\\gamma$.\n\nAnd since the conjunction is commutative in $\\textbf{CPL}$, ${\\sim}(\\alpha\\wedge\\beta)\\rightarrow{\\sim}(\\beta\\wedge\\alpha)$ or, what is equivalent, $(\\alpha\\uparrow\\beta)\\rightarrow(\\beta\\uparrow\\alpha)$, proving $\\textbf{Comm}$ is valid in $\\textbf{CPL}_{\\uparrow}$.\n\\end{example}\n\n\n\n\\subsection{Bottom and top elements, and classical negation}\n\nBefore anything else, we wish to show that $\\bI$ has a formula equivalent to a bottom, and how we can use this to define a classical negation on it. So, for any two formulas $\\alpha$ and $\\beta$ in the language of $\\bI$, consider\n\\[\\bot_{\\alpha\\beta}=\\alpha\\wedge(\\beta\\wedge(\\alpha\\uparrow\\beta)).\\]\n\n\\begin{lemma}\\label{Deduction meta-theorem for bI}\n\\begin{enumerate}\n\\item For a set $\\Gamma\\cup\\{\\alpha, \\beta\\}$ of formulas in $\\bI$, we have that $\\Gamma, \\alpha\\vdash_{\\bI}\\beta$ if and only if $\\Gamma\\vdash_{\\bI}\\alpha\\rightarrow\\beta$ (this is known as the deduction meta-theorem).\n\\item If $\\Gamma\\vdash_{\\bI}\\alpha\\rightarrow\\beta$ and $\\Gamma\\vdash_{\\bI}\\beta\\rightarrow\\gamma$, then $\\Gamma\\vdash_{\\bI}\\alpha\\rightarrow\\gamma$.\n\\item For a set $\\Gamma\\cup\\{\\alpha, \\beta, \\varphi\\}$ of formulas in $\\bI$, we have that, if $\\Gamma, \\alpha\\vdash_{\\bI}\\varphi$ and $\\Gamma, \\beta\\vdash_{\\bI}\\varphi$, then $\\Gamma, \\alpha\\vee\\beta\\vdash_{\\bI}\\varphi$ (this is know as a proof by cases).\n\\end{enumerate}\n\\end{lemma}\n\n\\begin{proof}\n\\begin{enumerate}\n\\item The proof of this result for the positive fragment of $\\textbf{CPL}$ can be found in \\cite{Men-IntLog} (notice the proof does not use the negation), Proposition $1.9$, and the result extends to $\\bI$ since this last logic has the axiom schemata of the previous one; but we make a point of proving this theorem given its importance.\n\nSuppose first that $\\Gamma, \\alpha\\vdash_{\\bI}\\beta$ and let $\\varphi_{1}, \\dotsc , \\varphi_{n}=\\beta$ be a demonstration of $\\beta$ from $\\Gamma\\cup\\{\\alpha\\}$; we show by induction that, for every $1\\leq i\\leq n$, $\\Gamma\\vdash_{\\bI}\\alpha\\rightarrow\\varphi_{i}$, and therefore $\\Gamma\\vdash_{\\bI}\\alpha\\rightarrow\\beta$ when we take $i=n$. \n\nThe case $i=1$ follows from the general case, being an instance of an axiom or a premise, so let us assume that the result holds for $\\varphi_{1}$ through $\\varphi_{i}$, and prove it for $\\varphi_{i+1}$; then $\\varphi_{i+1}$ can be one of the following.\n\n\\begin{enumerate}\n\\item An instance of an axiom: in this case, since $\\varphi_{i+1}$ and $\\varphi_{i+1}\\rightarrow(\\alpha\\rightarrow \\varphi_{i+1})$ are instances of axioms, the second formula of $\\textbf{Ax\\: 1}$, we obtain by Modus Ponens that $\\vdash_{\\bI}\\alpha\\rightarrow\\varphi_{i+1}$, and so $\\Gamma\\vdash_{\\bI}\\alpha\\rightarrow\\varphi_{i+1}$ since $\\bI$ is tarskian.\n\\item A premise: $\\varphi_{i+1}\\rightarrow(\\alpha\\rightarrow \\varphi_{i+1})$ certainly remains an instance of $\\textbf{Ax\\: 1}$, and through Modus Ponens $\\varphi_{i+1}\\vdash_{\\bI}\\alpha\\rightarrow\\varphi_{i+1}$, thus leaving us with $\\Gamma\\vdash_{\\bI}\\alpha\\rightarrow\\varphi_{i+1}$.\n\\item A conclusion of an inference rule: since we have only Modus Ponens as inference rule, this means that there exist $1\\leq j2$. This allows us to define a \\index{Fidel structure for $\\bI$}Fidel structure, presented as a $\\Sigma_{\\bI}^{\\textbf{CPL}}$-multialgebra, for $\\bI$ to be any $\\Sigma_{\\bI}^{\\textbf{CPL}}$-multialgebra $\\mathcal{A}=(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma_{\\bI}^{\\textbf{CPL}}})$ such that:\n\\begin{enumerate}\n\\item $(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma^{\\textbf{CPL}}})$ is a Boolean algebra;\\footnote{Here, a distinction is important: the operations corresponding to the symbols $\\bot$, $\\top$, $\\sim$, $\\vee$, $\\wedge$ and $\\rightarrow$ are deterministic, and will be treated as such; meanwhile, $\\uparrow$ will be, at most times, non-deterministic.}\n\\item for all $a, b\\in A$ and $c\\in\\uparrow_{\\mathcal{A}}(a,b)$, \n\\[\\wedge_{\\mathcal{A}}(a, \\wedge_{\\mathcal{A}}(b,c))=\\bot_{\\mathcal{A}};\\]\n\\item for all $a, b\\in A$, $\\uparrow_{\\mathcal{A}}(a, b)=\\uparrow_{\\mathcal{A}}(b,a)$.\\footnote{Technically, this last condition could be replaced by, for all $a, b\\in A$, $\\uparrow_{\\mathcal{A}}(a,b)\\cap\\uparrow_{\\mathcal{A}}(b,a)\\neq\\emptyset$, or simply erased from our definition of Fidel structures for $\\bI$; it only guarantees that, for every such Fidel structure, any function from the variables $\\mathcal{V}$ into $A$ may be extended to a restricted valuation. However, given the condition $\\nu(\\alpha\\uparrow\\beta)=\\nu(\\beta\\uparrow\\alpha)$ we will ask of these restricted valuations, dropping $\\uparrow_{\\mathcal{A}}(a, b)=\\uparrow_{\\mathcal{A}}(b,a)$ from the definition of a Fidel structure would only mean that some Fidel structure would not help in establishing the validity of an argument; the class of all Fidel structures, however, would still be sound and complete.}\n\\end{enumerate}\n\nFor simplicity, we will drop the indexes $\\mathcal{A}$ and use the standard infix notation.\n\nGiven a Fidel structure $\\mathcal{A}$, presented as a $\\Sigma_{\\bI}^{\\textbf{CPL}}$-multialgebra, for $\\bI$, a valuation over $\\mathcal{A}$ is a function $\\nu:F(\\Sigma_{\\bI}, \\mathcal{V})\\rightarrow A$ such that:\n\\begin{enumerate}\n\\item $\\nu(\\alpha\\#\\beta)=\\nu(\\alpha)\\#\\nu(\\beta)$, for $\\#\\in\\{\\vee, \\wedge, \\rightarrow\\}$;\n\\item $\\nu(\\alpha\\uparrow\\beta)\\in\\nu(\\alpha)\\uparrow\\nu(\\beta)$.\n\\end{enumerate}\n\nNotice that $\\nu$ is a valuation for $\\mathcal{A}$ if and only if it is a $\\Sigma_{\\bI}$-homomorphism, between $\\textbf{F}(\\Sigma_{\\bI}, \\mathcal{V})$ and $(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma_{\\bI}})$. \n\nFor every Fidel structure $\\mathcal{A}$ for $\\bI$, we will consider the restricted Nmatrix\\\\ $(\\mathcal{A}, \\{\\top\\}, \\mathcal{F}_{\\mathcal{A}})$, where $\\mathcal{F}_{\\mathcal{A}}$ is the set of valuations $\\nu:\\textbf{F}(\\Sigma_{\\bI}, \\mathcal{V})\\rightarrow\\mathcal{A}$ such that \n\\[\\nu(\\alpha\\uparrow\\beta)=\\nu(\\beta\\uparrow\\alpha),\\]\nfor any two formulas $\\alpha$ and $\\beta$ in $F(\\Sigma_{\\bI}, \\mathcal{V})$. If $\\Gamma$ proves $\\varphi$ according to such restricted Nmatrices, we will write $\\Gamma\\Vdash_{\\mathcal{F}}^{\\bI}\\varphi$\\label{VdashFbI}: as one could suspect, $\\Gamma\\vdash_{\\bI}\\varphi$ if and only if $\\Gamma\\Vdash_{\\mathcal{F}}^{\\bI}\\varphi$.\n\n\\begin{proposition}\nEach restricted Nmatrix $(\\mathcal{A}, \\{\\top\\}, \\mathcal{F}_{\\mathcal{A}})$, as described above, is structural.\n\\end{proposition}\n\n\\begin{proof}\nTake any $\\Sigma_{\\bI}$-homomorphism $\\sigma:\\textbf{F}(\\Sigma_{\\bI}, \\mathcal{V})\\rightarrow\\textbf{F}(\\Sigma_{\\bI}, \\mathcal{V})$ and an element $\\nu\\in\\mathcal{F}_{\\mathcal{A}}$, which is by definition of $\\mathcal{F}_{\\mathcal{A}}$ a $\\Sigma_{\\bI}$-homomorphism $\\nu:\\textbf{F}(\\Sigma_{\\bI}, \\mathcal{V})\\rightarrow \\mathcal{A}$ such that, for any two formulas $\\alpha$ and $\\beta$, $\\nu(\\alpha\\uparrow\\beta)=\\nu(\\beta\\uparrow\\alpha)$.\n\nClearly $\\nu\\circ\\sigma:\\textbf{F}(\\Sigma_{\\bI}, \\mathcal{V})\\rightarrow \\mathcal{A}$ is a $\\Sigma_{\\bI}$-homomorphism, so we only need to show that, for any two formulas $\\alpha$ and $\\beta$, $\\nu\\circ\\sigma(\\alpha\\uparrow\\beta)=\\nu\\circ\\sigma(\\beta\\uparrow\\alpha)$; and since \n\\[\\nu\\circ\\sigma(\\alpha\\uparrow\\beta)=\\nu(\\sigma(\\alpha\\uparrow\\beta))=\\nu(\\sigma(\\alpha)\\uparrow\\sigma(\\beta))=\\nu(\\sigma(\\beta)\\uparrow\\sigma(\\alpha))=\\nu(\\sigma(\\beta\\uparrow\\alpha))=\\nu\\circ\\sigma(\\beta\\uparrow\\alpha),\\]\nwe finish the proof.\n\\end{proof}\n\nIf we simply consider the class $\\mathbb{M}$ of Nmatrices $(\\mathcal{A},\\{\\top\\})$, where $\\mathcal{A}$ is a Fidel structure for $\\bI$, it is not hard to prove that $\\Gamma\\vDash_{\\mathbb{M}}\\varphi$ if and only if $\\Gamma\\vdash_{\\bI^{-}}\\varphi$.\n\n\n\\subsubsection{Soundness}\n\nFirst of all, we state that for any instance of an axiom $\\alpha$ of the positive fragment of classical propositional logic in $\\bI$, $\\Vdash_{\\mathcal{F}}^{\\bI}\\alpha$ (meaning that $\\emptyset\\Vdash_{\\mathcal{F}}^{\\bI}\\alpha$): this is true because Boolean algebras model classical propositional logic. To see one example, take an instance $\\alpha\\rightarrow(\\beta\\rightarrow\\alpha)$ of $\\textbf{Ax\\: 1}$: we have that\n\\[\\nu(\\alpha\\rightarrow(\\beta\\rightarrow\\alpha))=\\nu(\\alpha)\\rightarrow(\\nu(\\beta)\\rightarrow\\nu(\\alpha)),\\]\nand remembering that in a Boolean algebra, $x\\rightarrow y={\\sim}x\\vee y$, this equals ${\\sim}\\nu(\\alpha)\\vee({\\sim}\\nu(\\beta)\\vee\\nu(\\alpha))$; using that $\\vee$ is commutative and associative, in this order, this equals\n\\[({\\sim}\\nu(\\alpha)\\vee\\nu(\\alpha))\\vee{\\sim}\\nu(\\beta)=\\top\\vee{\\sim}\\nu(\\beta)=\\top,\\]\nand since for any valuation $\\nu$ we have that $\\nu(\\alpha\\rightarrow(\\beta\\rightarrow\\alpha))=\\top$, we find $\\Vdash_{\\mathcal{F}}^{\\bI}\\alpha\\rightarrow(\\beta\\rightarrow\\alpha)$.\n\nNow we take an instance $(\\alpha\\uparrow\\beta)\\rightarrow(\\alpha\\rightarrow(\\beta\\rightarrow\\gamma))$ of axiom $\\textbf{Ip}$; using again that $x\\rightarrow y={\\sim}x\\vee y$, the image under a valuation $\\nu$ of this formula is \n\\[{\\sim}\\nu(\\alpha\\uparrow\\beta)\\vee({\\sim}\\nu(\\alpha)\\vee({\\sim}\\nu(\\beta)\\vee\\nu(\\gamma)))=({\\sim}\\nu(\\alpha\\uparrow\\beta)\\vee({\\sim}\\nu(\\alpha)\\vee{\\sim}\\nu(\\beta)))\\vee\\nu(\\gamma);\\]\nusing that ${\\sim}x\\vee{\\sim}y={\\sim}(x\\wedge y)$ in a Boolean algebra, one of the De Morgan's laws, we get this expression equals\n\\[{\\sim}(\\nu(\\alpha\\uparrow\\beta)\\wedge(\\nu(\\alpha)\\wedge\\nu(\\beta)))\\vee\\nu(\\gamma),\\]\nand from the requirements made on the multioperation $\\uparrow$ in the definition of Fidel structures for $\\bI$, we get that $\\nu(\\alpha\\uparrow\\beta)\\wedge(\\nu(\\alpha)\\wedge\\nu(\\beta))=\\bot$, and therefore the whole expression simplifies to\n\\[{\\sim}\\bot\\vee\\nu(\\gamma)=\\top\\vee\\nu(\\gamma)=\\top.\\]\n\nFinally, take an instance $(\\alpha\\uparrow\\beta)\\rightarrow(\\beta\\uparrow\\alpha)$ of $\\textbf{Comm}$, and for any restricted Nmatrix $(\\mathcal{A},\\{\\top\\},\\mathcal{F}_{\\mathcal{A}})$ for $\\bI$, its image under a $\\nu\\in\\mathcal{F}_{\\mathcal{A}}$ is $\\nu(\\alpha\\uparrow\\beta)\\rightarrow\\nu(\\beta\\uparrow\\alpha)$. Since $\\nu\\in\\mathcal{F}_{\\mathcal{A}}$, we have that $\\nu(\\alpha\\uparrow\\beta)=\\nu(\\beta\\uparrow\\alpha)$, and therefore $\\nu((\\alpha\\uparrow\\beta)\\rightarrow(\\beta\\uparrow\\alpha))$ equals $\\nu(\\alpha\\uparrow\\beta)\\rightarrow\\nu(\\alpha\\uparrow\\beta)$, which is always $\\top$, meaning \"$\\Vdash_{\\mathcal{F}}^{\\bI}$\" models all axioms of $\\bI$.\n\n\\begin{theorem}\\label{Soundness of Fidel structures for bI}\nGiven formulas $\\Gamma\\cup\\{\\varphi\\}$ of $\\bI$, if $\\Gamma\\vdash_{\\bI}\\varphi$ then $\\Gamma\\Vdash_{\\mathcal{F}}^{\\bI}\\varphi$.\n\\end{theorem}\n\n\\begin{proof}\nLet $\\alpha_{1}, \\dotsc , \\alpha_{n}$ be a demonstration of $\\varphi$ from $\\Gamma$, with $\\alpha_{n}=\\varphi$, and let $\\mathcal{A}$ be a Fidel structure, represented as a $\\Sigma_{\\bI}$-multialgebra, for $\\bI$, with $\\nu$ a valuation for $\\bI$ over $\\mathcal{A}$ in $\\mathcal{F}_{\\mathcal{A}}$.\n\nWe aim to show that, if $\\nu(\\Gamma)\\subseteq\\{\\top\\}$, then $\\nu(\\alpha_{i})=\\top$ for every $i\\in\\{1, \\dotsc , n\\}$. We will proceed by induction, assuming that the result already holds for all formulas prior to $\\alpha_{i}$ (notice that $\\alpha_{0}$ is either an axiom or a premise). Then we have that $\\alpha_{i}$ is either:\n\\begin{enumerate}\n\\item an instance of an axiom, when we already proved $\\Vdash_{\\mathcal{F}}^{\\bI}\\alpha_{i}$, and therefore $\\nu(\\alpha_{i})=\\top$;\n\\item a premise, and by our hypothesis that $\\nu(\\Gamma)\\subseteq\\{\\top\\}$ we find $\\nu(\\alpha_{i})=\\top$;\n\\item there exist $j,k2$.\n\nSo, it becomes easy to define Fidel structures for $\\nbI$: \\index{Fidel structure for $\\nbI$}a Fidel structure for $\\nbI$ is any $\\Sigma_{\\nbI}^{\\textbf{CPL}}$-multialgebra $\\mathcal{A}=(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma_{\\nbI}^{\\textbf{CPL}}})$ such that:\n\\begin{enumerate}\n\\item $(A,\\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma^{\\textbf{CPL}}})$ is a Boolean algebra;\n\\item for all $a\\in A$ and $b\\in \\neg _{\\mathcal{A}}(a)$, $a\\vee b=\\top$;\n\\item for all $a,b\\in A$ and $c\\in\\uparrow_{\\mathcal{A}}(a,b)$, $a\\wedge(b\\wedge c)=\\bot$;\n\\item for all $a, b\\in A$, $a\\uparrow b=b\\uparrow a$.\n\\end{enumerate}\n\nAs we did before, we shall drop the index $\\mathcal{A}$ and use the infix notation.\n \nGiven a Fidel structure $\\mathcal{A}$, presented as a $\\Sigma_{\\nbI}^{\\textbf{CPL}}$-multialgebra, for $\\nbI$, a valuation for $\\mathcal{A}$ is any $\\Sigma_{\\nbI}$-homomorphism $\\nu:\\textbf{F}(\\Sigma_{\\nbI}, \\mathcal{V})\\rightarrow \\mathcal{A}$; and, for every Fidel structure $\\mathcal{A}$ for $\\nbI$, we will consider the restricted Nmatrix $(\\mathcal{A}, \\{\\top\\}, \\mathcal{F}_{\\mathcal{A}})$, where $\\mathcal{F}_{\\mathcal{A}}$ is the set of valuations $\\nu:\\textbf{F}(\\Sigma_{\\nbI}, \\mathcal{V})\\rightarrow\\mathcal{A}$ such that \n\\[\\nu(\\alpha\\uparrow\\beta)=\\nu(\\beta\\uparrow\\alpha),\\]\nfor any two formulas $\\alpha$ and $\\beta$ in $F(\\Sigma_{\\nbI}, \\mathcal{V})$; it is clear how these RNmatrices are structural. If $\\Gamma$ proves $\\varphi$ according to such restricted Nmatrices, we will write\\label{VdashFnbI} $\\Gamma\\Vdash_{\\mathcal{F}}^{\\nbI}\\varphi$.\n\nIt is easy to see that if $\\phi$ is an instance of an axiom of $\\bI$, $\\Vdash_{\\mathcal{F}}^{\\nbI}\\phi$, and if $\\Vdash_{\\mathcal{F}}^{\\nbI}\\alpha$ and $\\Vdash_{\\mathcal{F}}^{\\nbI}\\alpha\\rightarrow\\beta$ then $\\Vdash_{\\mathcal{F}}^{\\nbI}\\beta$: we would like to prove that, for any formula $\\alpha$ of $\\nbI$, we also have \n\\[\\Vdash_{\\mathcal{F}}^{\\nbI}\\alpha\\vee\\neg\\alpha,\\]\nimplying that \"$\\Vdash_{\\mathcal{F}}^{\\nbI}$\" models the axiom schemata and rules of inference of $\\nbI$.\n\nSo, for any Fidel structure $\\mathcal{A}$ for $\\nbI$ and valuation $\\nu$ over $\\mathcal{A}$, we have that $\\nu(\\alpha\\vee\\neg \\alpha)=\\nu(\\alpha)\\vee\\nu(\\neg \\alpha)$; now, $\\nu(\\neg \\alpha)\\in\\neg\\nu(\\alpha)$, and therefore, from the conditions demanded of a Fidel structure for $\\nbI$, $\\nu(\\alpha)\\vee\\nu(\\neg \\alpha)=\\top$, what finishes the desired proof.\n\n\\begin{theorem}\nGiven formulas $\\Gamma\\cup\\{\\varphi\\}$ of $\\nbI$, if $\\Gamma\\vdash_{\\nbI}\\varphi$ then $\\Gamma\\Vdash_{\\mathcal{F}}^{\\nbI}\\varphi$.\n\\end{theorem}\n\nTo prove the reciprocal result, we define the equivalence relation between formulas of $\\nbI$ such that, for a fixed set of formulas $\\Gamma$, \\label{equivnbI}$\\alpha\\equiv_{\\Gamma}^{\\nbI}\\beta$ if and only if $\\Gamma\\vdash_{\\nbI}\\alpha\\rightarrow\\beta$ and $\\Gamma\\vdash_{\\nbI}\\beta\\rightarrow\\alpha$; this relation is also a congruence with respect to the connectives in $\\{\\vee, \\wedge,\\rightarrow\\}$, allowing us to make $A^{\\nbI}_{\\Gamma}=F(\\Sigma_{\\nbI},\\mathcal{V})\/\\equiv_{\\Gamma}^{\\nbI}$ into the universe of a Boolean algebra. By defining, for classes of formulas $[\\alpha]$ and $[\\beta]$,\n\\[\\neg[\\alpha]=\\{[\\neg \\varphi]\\ :\\ \\varphi\\in[\\alpha]\\}\\]\nand\n\\[[\\alpha]\\uparrow[\\beta]=\\{[\\varphi\\uparrow\\psi]\\ :\\ \\varphi\\in[\\alpha], \\psi\\in[\\beta]\\},\\]\nwe can prove \\label{AnbIGamma}$\\mathcal{A}^{\\nbI}_{\\Gamma}=(A^{\\nbI}_{\\Gamma},\\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma_{\\nbI}^{\\textbf{CPL}}})$ is a Fidel structure, presented as a $\\Sigma_{\\nbI}^{\\textbf{CPL}}$-multialgebra, for $\\nbI$, which we shall call the Lindenbaum-Tarski Fidel structure of $\\nbI$ associated to $\\Gamma$.\n\nThe proof that, for $[\\gamma]\\in [\\alpha]\\uparrow[\\beta]$, we have $[\\alpha]\\wedge([\\beta]\\wedge[\\gamma])=\\bot$, and that $[\\alpha]\\uparrow[\\beta]=[\\beta]\\uparrow[\\alpha]$ are exactly the same as those for the same facts in $\\bI$ (found in Section \\ref{Completeness fo Fidel structures for bI}), so we will rather focus on proving that, for any class of formulas $[\\alpha]$ and any $[\\beta]\\in \\neg[\\alpha]$, $[\\alpha]\\vee[\\beta]=\\top$: by definition, there exists $\\varphi\\in\\alpha$ such that $[\\beta]=[\\neg\\varphi]$.\n\nThen $[\\alpha]\\vee[\\beta]=[\\alpha\\vee\\neg \\varphi]$: very clearly $\\Gamma\\vdash_{\\nbI}\\alpha\\vee\\neg \\varphi\\rightarrow\\top_{\\alpha}$, remaining for us to show that $\\Gamma\\vdash_{\\nbI}\\top_{\\alpha}\\rightarrow\\alpha\\vee\\neg \\varphi$ or, better yet, that $\\Gamma\\vdash_{\\nbI}\\alpha\\vee\\neg \\varphi$, by one application of the deduction meta-theorem and the fact that a top element is always a tautology.\n\nWe know $\\varphi\\vee\\neg \\varphi$ is an instance of $\\textbf{Ax\\: 11}^{*}$ and $\\Gamma\\vdash_{\\nbI}\\varphi\\rightarrow\\alpha$:\n\\begin{enumerate}\n\\item $\\Gamma,\\varphi\\vdash_{\\nbI}\\alpha$ by the deduction meta-theorem, and from the instance $\\alpha\\rightarrow\\alpha\\vee\\neg \\varphi$ of $\\textbf{Ax\\: 6}$ and one application of Modus Ponens we get that $\\Gamma,\\varphi\\vdash_{\\nbI}\\alpha\\vee\\neg \\varphi$;\n\\item $\\Gamma,\\neg \\varphi\\vdash_{\\nbI}\\neg \\varphi$, and from the instance $\\neg \\varphi\\rightarrow\\alpha\\vee\\neg \\varphi$ of $\\textbf{Ax\\: 7}$ and the deduction meta-theorem we get that $\\Gamma,\\neg \\varphi\\vdash_{\\nbI}\\alpha\\vee\\neg \\varphi$;\n\\end{enumerate}\nfrom this, we get that $\\Gamma, \\varphi\\vee\\neg \\varphi\\vdash_{\\nbI}\\alpha\\vee\\neg \\varphi$ by a proof by cases, and since $\\varphi\\vee\\neg \\varphi$ is an instance of an axiom, $\\Gamma\\vdash_{\\nbI}\\alpha\\vee\\neg \\varphi$, as we wanted to prove.\n\n\\begin{theorem}\nGiven formulas $\\Gamma\\cup\\{\\varphi\\}$ of $\\nbI$, if $\\Gamma\\Vdash_{\\mathcal{F}}^{\\nbI}\\varphi$ then $\\Gamma\\vdash_{\\nbI}\\varphi$.\n\\end{theorem}\n\n\n\n\n\n\\subsection{Decision method}\n\nWe take the Boolean algebra $\\textbf{2}$ once again and define the unary multioperation $\\neg:\\{0,1\\}\\rightarrow\\mathcal{P}(\\{0,1\\})\\setminus\\{\\emptyset\\}$ trough $\\neg 0=\\{1\\}$ and $\\neg 1=\\{0,1\\}$,\n\n\\begin{figure}[H]\n\\centering\n\\begin{tabular}{|l|r|}\n\\hline\n& $\\neg $\\\\ \\hline\n$0$ & $\\{1\\}$\\\\ \\hline\n$1$ & $\\{0,1\\}$\\\\\\hline\n\\end{tabular}\n\\caption*{Table for our Paraconsistent Negation}\n\\end{figure}\n\nand, as we did in Section \\ref{Decision Method for bI}, we define $1\\uparrow 1=\\{0\\}$ and $x\\uparrow y=\\{0,1\\}$ for any other values of $x$ and $y$. By adding the previous operations to $\\textbf{2}$, we transform it into a $\\Sigma_{\\nbI}^{\\textbf{CPL}}$-multialgebra\\label{2nbI} $\\textbf{2}_{\\nbI}$ with a finite universe $\\{0,1\\}$. It is easy to prove that:\n\\begin{enumerate}\n\\item $(\\{0,1\\}, \\{\\sigma_{\\textbf{2}_{\\nbI}}\\}_{\\sigma\\in\\Sigma^{\\textbf{CPL}}})$ is a Boolean algebra, since it equals $\\textbf{2}$;\n\\item for any $x,y\\in\\{0,1\\}$ and $z\\in x\\uparrow y$, $x\\wedge(y\\wedge z)=0$;\n\\item for any $x, y\\in\\{0,1\\}$, $x\\uparrow y=y\\uparrow x$.\n\\end{enumerate}\n\nIt remains to be shown, to prove that $\\textbf{2}_{\\nbI}$ is a Fidel structure presented as a $\\Sigma_{\\nbI}^{\\textbf{CPL}}$-multialgebra for $\\nbI$, that for any $x\\in \\{0,1\\}$ and $y\\in \\neg x$, $x\\vee y=1$. If $x=0$, we must have $y=1$, and then $x\\vee y=0\\vee 1=1$; if $x=1$, we have that, either $y=0$, when $x\\vee y=1\\vee 0=1$, or $y=1$, when $x\\vee y=1\\vee1=1$, so we are done\n\n\\begin{theorem}\n$\\nu:F(\\Sigma_{\\nbI},\\mathcal{V})\\rightarrow\\{0,1\\}$ is a bivaluation for $\\nbI$ if, and only if, it is a $\\Sigma_{\\nbI}$-homomorphism from $\\textbf{F}(\\Sigma_{\\nbI}, \\mathcal{V})$ to $\\textbf{2}_{\\nbI}$ which lies in $\\mathcal{F}_{\\textbf{2}_{\\nbI}}$.\n\\end{theorem}\n\nAs expected, we will denote the restricted Nmatrix $(\\textbf{2}_{\\nbI}, \\{1\\}, \\mathcal{F}_{\\textbf{2}_{\\nbI}})$ simply by $\\mathbb{2}_{\\nbI}$.\n\n\\begin{theorem}\nGiven formulas $\\Gamma\\cup\\{\\varphi\\}$ of $\\nbI$, $\\Gamma\\vDash_{\\nbI}\\varphi$ if and only if $\\Gamma\\vDash_{\\mathbb{2}_{\\nbI}}\\varphi$.\n\\end{theorem}\n\n\nSo, take the Nmatrix $(\\textbf{2}_{\\nbI}, \\{1\\})$ subjacent to the RNmatrix $(\\textbf{2}_{\\nbI}, \\{1\\}, \\mathcal{F}_{\\textbf{2}_{\\nbI}})$: if one writes the row-branching truth table for the said Nmatrix to test a formula $\\varphi$ of $\\nbI$ and simply erases the rows where $\\alpha\\uparrow\\beta$ and $\\beta\\uparrow\\alpha$ are given different values, we get a row-branching, row-eliminating truth table that decides the validity, in a finite number of steps, of $\\varphi$ in $\\nbI$. Essentially, this is the same decision method as the one presented for $\\bI$ in Section \\ref{Decision Method for bI}, with an extra multi-operation standing for negation.\n\n\n\\subsection{Another decision method}\n\nGiven the similitude between $\\bI$ and $\\nbI$, one should hope that a tableau calculus for $\\nbI$, that we will name $\\mathbb{T}_{\\nbI}$, could be easily obtained from $\\mathbb{T}_{\\bI}$ of Section \\ref{Tableaux for bI}: this is, in fact, the case. It is sufficient to add one new rule governing the case in which $\\neg\\varphi$ is false, \n\\[\\frac{\\textsf{0}(\\neg\\varphi)}{\\textsf{1}(\\varphi)}\\]\nand change the conditions for a branch to be complete: a branch $\\theta$ is complete, now in $\\mathbb{T}_{\\nbI}$, when it contains, for every $\\textsf{L}(\\gamma)$ in $\\theta$ neither of the form $\\textsf{0}(\\varphi\\uparrow\\psi)$ or $\\textsf{1}(\\neg\\varphi)$ and with $\\gamma$ not a variable, all the labeled formulas of one of the branches in the rule headed by $\\textsf{L}(\\gamma)$.\n\n\n\\section{Collapsing axioms}\n\n\n\\subsection{$\\textbf{ci}^{\\uparrow}$}\\label{ciuparrow}\n\nNow that we have in our signature a paraconsistent negation, we are capable of doing to the axioms schemata $\\textbf{ci}$ and $\\textbf{cl}$ much the same we did to $\\textbf{ciw}$ when we transformed it into $\\textbf{ciw}^{\\uparrow}$: consider\\label{ciuparrow}\n\\[\\tag{$\\textbf{ci}^{\\uparrow}$} \\neg(\\alpha\\uparrow\\beta)\\rightarrow(\\alpha\\wedge\\beta)\\]\nand the logic $\\nbI\\textbf{ci}^{\\uparrow}$ obtained from $\\nbI$ by addition of $\\textbf{ci}^{\\uparrow}$; we shall prove, once again, that $\\nbI\\textbf{ci}^{\\uparrow}$ collapses back to $\\textbf{CPL}$, and $\\alpha\\uparrow\\beta$ is equivalent to $\\alpha\\wedge\\beta\\rightarrow\\bot_{\\alpha\\beta}$.\n\nIt is clear how $\\alpha\\uparrow\\beta\\vdash_{\\nbI\\textbf{ci}^{\\uparrow}}\\alpha\\wedge\\beta\\rightarrow\\bot_{\\alpha\\beta}$, since such a deduction could be made in $\\bI$ itself, so let us concentrate in proving that $\\alpha\\wedge\\beta\\rightarrow\\bot_{\\alpha\\beta}\\vdash_{\\nbI\\textbf{ci}^{\\uparrow}}\\alpha\\uparrow\\beta$.\n\nIt is also clear that \n\\[\\alpha\\uparrow\\beta, \\alpha\\wedge\\beta\\rightarrow\\bot_{\\alpha\\beta}\\vdash_{\\nbI\\textbf{ci}^{\\uparrow}}\\alpha\\uparrow\\beta,\\]\nbut from the instance $\\neg(\\alpha\\uparrow\\beta)\\rightarrow(\\alpha\\wedge\\beta)$ of $\\textbf{ci}^{\\uparrow}$ plus Modus Ponens we can similarly get \n\\[\\neg(\\alpha\\uparrow\\beta), \\alpha\\wedge\\beta\\rightarrow\\bot_{\\alpha\\beta}\\vdash_{\\nbI\\textbf{ci}^{\\uparrow}}\\alpha\\uparrow\\beta,\\]\nand using a proof by cases we get\n\\[(\\alpha\\uparrow\\beta)\\vee\\neg(\\alpha\\uparrow\\beta), \\alpha\\wedge\\beta\\rightarrow\\bot_{\\alpha\\beta}\\vdash_{\\nbI\\textbf{ci}^{\\uparrow}}\\alpha\\uparrow\\beta.\\]\nSince $(\\alpha\\uparrow\\beta)\\vee\\neg(\\alpha\\uparrow\\beta)$ is an instance of $\\textbf{Ax\\: 11}^{*}$, we get the desired result.\n\nBut we can prove an even nicer result, that implies the previous: an instance of $\\textbf{ci}^{\\uparrow}$ actually implies its corresponding instance of $\\textbf{ciw}^{\\uparrow}$ in $\\nbI$, that is,\n\\[\\neg(\\alpha\\uparrow\\beta)\\rightarrow(\\alpha\\wedge\\beta)\\vdash_{\\nbI}(\\alpha\\uparrow\\beta)\\vee(\\alpha\\wedge\\beta).\\]\nWe have that\n\\[\\neg(\\alpha\\uparrow\\beta),\\neg(\\alpha\\uparrow\\beta)\\rightarrow(\\alpha\\wedge\\beta)\\vdash_{\\nbI}(\\alpha\\uparrow\\beta)\\vee(\\alpha\\wedge\\beta)\\]\nby an application of Modus Ponens and remembering $(\\alpha\\wedge\\beta)\\rightarrow[(\\alpha\\uparrow\\beta)\\vee(\\alpha\\wedge\\beta)]$ is an instance of $\\textbf{Ax\\: 7}$; furthermore, we trivially find\n\\[\\alpha\\uparrow\\beta,\\neg(\\alpha\\uparrow\\beta)\\rightarrow(\\alpha\\wedge\\beta)\\vdash_{\\nbI}(\\alpha\\uparrow\\beta)\\vee(\\alpha\\wedge\\beta)\\]\nby remembering $(\\alpha\\uparrow\\beta)\\rightarrow[(\\alpha\\uparrow\\beta)\\vee(\\alpha\\wedge\\beta)]$ is an instance of $\\textbf{Ax\\: 6}$. By a proof by cases, \n\\[(\\alpha\\uparrow\\beta)\\vee\\neg(\\alpha\\uparrow\\beta),\\neg(\\alpha\\uparrow\\beta)\\rightarrow(\\alpha\\wedge\\beta)\\vdash_{\\nbI}(\\alpha\\uparrow\\beta)\\vee(\\alpha\\wedge\\beta),\\]\nand since $(\\alpha\\uparrow\\beta)\\vee\\neg(\\alpha\\uparrow\\beta)$ is an instance of $\\textbf{Ax\\: 11}^{*}$, we derive the aforementioned result.\n\n\n\\subsection{$\\textbf{cl}^{\\uparrow}$}\\label{cluparrow}\n\nAs one should expect, the axiom schema\\label{cluparrow}\n\\[\\tag{$\\textbf{cl}^{\\uparrow}$} \\neg(\\alpha\\wedge\\beta)\\rightarrow(\\alpha\\uparrow\\beta)\\]\nwill collapse its corresponding logic $\\nbI\\textbf{cl}^{\\uparrow}$ back to $\\textbf{CPL}$, with $\\alpha\\uparrow\\beta$ standing for $\\alpha\\wedge\\beta\\rightarrow\\bot_{\\alpha\\beta}$.\n\nTo see that $\\alpha\\wedge\\beta\\rightarrow\\bot_{\\alpha\\beta}\\vdash_{\\nbI\\textbf{cl}^{\\uparrow}}\\alpha\\uparrow\\beta$, we begin by noticing\n\\[\\alpha\\wedge\\beta, \\alpha\\wedge\\beta\\rightarrow\\bot_{\\alpha\\beta}\\vdash_{\\nbI\\textbf{cl}^{\\uparrow}}\\alpha\\uparrow\\beta,\\]\nby use of Modus Ponens and the fact $\\bot_{\\alpha\\beta}$ behaves like a bottom element; at the same time,\n\\[\\neg(\\alpha\\wedge\\beta), \\alpha\\wedge\\beta\\rightarrow\\bot_{\\alpha\\beta}\\vdash_{\\nbI\\textbf{cl}^{\\uparrow}}\\alpha\\uparrow\\beta,\\]\ngiven the instance $ \\neg(\\alpha\\wedge\\beta)\\rightarrow(\\alpha\\uparrow\\beta)$ of $\\textbf{cl}^{\\uparrow}$. With a proof by cases,\n\\[(\\alpha\\wedge\\beta)\\vee\\neg(\\alpha\\wedge\\beta), \\alpha\\wedge\\beta\\rightarrow\\bot_{\\alpha\\beta}\\vdash_{\\nbI\\textbf{cl}^{\\uparrow}}\\alpha\\uparrow\\beta,\\]\nand since $(\\alpha\\wedge\\beta)\\vee\\neg(\\alpha\\wedge\\beta)$ is an instance of $\\textbf{Ax\\: 11}^{*}$ we obtain the desired result.\n\nHowever, we can again prove a stronger result: an instance of $\\textbf{cl}^{\\uparrow}$ can prove, in $\\nbI$, the corresponding instance of $\\textbf{ciw}^{\\uparrow}$, what proves, as a special case, the previous collapse. To summarize, we wish to prove\n\\[\\neg(\\alpha\\wedge\\beta)\\rightarrow(\\alpha\\uparrow\\beta)\\vdash_{\\nbI}(\\alpha\\uparrow\\beta)\\vee(\\alpha\\wedge\\beta).\\]\nThis is pretty simple:\n\\[\\alpha\\wedge\\beta, \\neg(\\alpha\\wedge\\beta)\\rightarrow(\\alpha\\uparrow\\beta)\\vdash_{\\nbI}(\\alpha\\uparrow\\beta)\\vee(\\alpha\\wedge\\beta),\\]\nby the instance $(\\alpha\\wedge\\beta)\\rightarrow[(\\alpha\\uparrow\\beta)\\vee(\\alpha\\wedge\\beta)]$ of $\\textbf{Ax\\: 7}$, and\n\\[\\neg(\\alpha\\wedge\\beta), \\neg(\\alpha\\wedge\\beta)\\rightarrow(\\alpha\\uparrow\\beta)\\vdash_{\\nbI}(\\alpha\\uparrow\\beta)\\vee(\\alpha\\wedge\\beta)\\]\nby Modus Ponens and the instance $(\\alpha\\uparrow\\beta)\\rightarrow[(\\alpha\\uparrow\\beta)\\vee(\\alpha\\wedge\\beta)]$ of $\\textbf{Ax\\: 6}$; with a proof by cases, \n\\[(\\alpha\\wedge\\beta)\\vee\\neg(\\alpha\\wedge\\beta), \\neg(\\alpha\\wedge\\beta)\\rightarrow(\\alpha\\uparrow\\beta)\\vdash_{\\nbI}(\\alpha\\uparrow\\beta)\\vee(\\alpha\\wedge\\beta),\\]\nand since $(\\alpha\\wedge\\beta)\\vee\\neg(\\alpha\\wedge\\beta)$ is an instance of $\\textbf{Ax\\: 11}^{*}$, the proof is done.\n\n\n\n\n\n\n\n\n\n\n\n\n\n\\section{The logics $\\nbIciw$, $\\nbIci$ and $\\nbIcl$}\n\nOne notices that one of the basic axiom schema, $\\textbf{Ip}$, of $\\bI$ is\n\\[(\\alpha\\uparrow\\beta)\\rightarrow(\\alpha\\rightarrow(\\beta\\rightarrow\\gamma)),\\]\nwhich can be seem as the basic axiom of $\\textbf{mbC}$, $\\textbf{bc1}$,\n\\[\\circ\\alpha\\rightarrow(\\alpha\\rightarrow(\\neg\\alpha\\rightarrow\\beta)),\\]\nwith $\\neg\\alpha$ replaced by $\\beta$ and $\\circ\\alpha$, that is, that $\\alpha$ is consistent, replaced by $\\alpha\\uparrow\\beta$, that is, that $\\alpha$ is incompatible with $\\beta$.\n\nAn attempt to make something similar to the axiom $\\textbf{ciw}$,\n\\[\\circ\\alpha\\vee(\\alpha\\wedge\\neg\\alpha),\\]\nis not successful, as the axiom $\\textbf{ciw}^{\\uparrow}$\n\\[(\\alpha\\uparrow\\beta)\\vee(\\alpha\\wedge\\beta)\\]\ncollapses the incompatibility operator back to its classical interpretation. However, in $\\nbI$, where we find at our disposal a non-classical negation, we can take an adaptation of the axiom $\\textbf{ciw}$ to the language $\\Sigma_{\\nbI}$, instead of the generalization $\\textbf{ciw}^{\\uparrow}$, and study the resulting logic. We will do the same with the axioms $\\textbf{ci}$ and $\\textbf{cl}$, producing three logics that mix paraconsistency and incompatibility.\n\nSo, over the signature $\\Sigma_{\\nbI}$, we consider the systems:\n\\begin{enumerate}\n\\item $\\nbIciw$\\label{nbIciw}, obtained from $\\nbI$ by addition of the axiom schema\\label{ciw*}\n\\[\\tag{$\\textbf{ciw}^{*}$} (\\alpha\\uparrow\\neg\\alpha)\\vee(\\alpha\\wedge\\neg\\alpha);\\]\n\\item $\\nbIci$\\label{nbIci}, obtained from $\\nbI$ by adding the axiom schema\\label{ci*}\n\\[\\tag{$\\textbf{ci}^{*}$}\\neg(\\alpha\\uparrow\\neg\\alpha)\\rightarrow(\\alpha\\wedge\\neg\\alpha);\\]\n\\item $\\nbIcl$\\label{nbIcl}, obtained from $\\nbI$ by adding\\label{cl*}\n\\[\\tag{$\\textbf{cl}^{*}$}\\neg(\\alpha\\wedge\\neg\\alpha)\\rightarrow(\\alpha\\uparrow\\neg\\alpha).\\]\n\\end{enumerate}\n\n\n\\subsection{Bivaluations}\n\n\\begin{definition}\\label{definition of Bivaluation for nbIciw, nbIci, ...}\nA \\index{Bivaluation for $\\nbIciw$}\\index{Bivaluation for $\\nbIci$}\\index{Bivaluation for $\\nbIcl$}bivaluation for $\\mathcal{L}\\in\\{\\nbIciw, \\nbIci, \\nbIcl\\}$ is a valuation for $\\nbI$ satisfying in addition that:\n\\begin{enumerate}\n\\item $\\nu(\\alpha\\uparrow\\neg\\alpha)=1$ if and only if $\\nu(\\alpha)=0$ or $\\nu(\\neg\\alpha)=0$;\n\\item \\begin{enumerate}\n\\item if $\\mathcal{L}=\\nbIci$ and $\\nu(\\neg(\\alpha\\uparrow\\neg\\alpha))=1$, then $\\nu(\\alpha)=\\nu(\\neg\\alpha)=1$;\n\\item if $\\mathcal{L}=\\nbIcl$ and $\\nu(\\neg(\\alpha\\wedge\\neg\\alpha))=1$, then $\\nu(\\alpha\\uparrow\\neg\\alpha)=1$.\n\\end{enumerate}\n\\end{enumerate}\n\\end{definition}\n\nAs usual, given a set of formulas $\\Gamma\\cup\\{\\varphi\\}$ on the signature $\\Sigma_{\\nbI}$, we say $\\Gamma$ proves $\\varphi$ according to bivaluations for $\\mathcal{L}\\in\\{\\nbIciw, \\nbIci, \\nbIcl\\}$, and write \\label{vDashnbIciw}\\label{vDashnbIci}\\label{vDashnbIcl}$\\Gamma\\vDash_{\\mathcal{L}}\\varphi$, if, for every bivaluation $\\nu$ for $\\mathcal{L}$ such that $\\nu(\\Gamma)\\subseteq\\{1\\}$, one has $\\nu(\\varphi)=1$.\n\nSince bivaluations for $\\mathcal{L}\\in\\{\\nbIciw, \\nbIci, \\nbIcl\\}$ are bivaluations for $\\nbI$ satisfying some additional property, it is easy to see that \"$\\vDash_{\\mathcal{L}}$\" models all those axiom schemata of $\\nbI$, plus Modus Ponens, its only inference rule; we wish now to prove that \"$\\vDash_{\\nbIciw}$\" also models $\\textbf{ciw}^{*}$, and an analogous result holds for the other logics.\n\n\\begin{enumerate}\n\\item Take an instance $(\\alpha\\uparrow\\neg\\alpha)\\vee(\\alpha\\wedge\\neg\\alpha)$ of $\\textbf{ciw}^{*}$ and a bivaluation $\\nu$ for $\\nbIciw$, and we see that \n\\[\\nu((\\alpha\\uparrow\\neg\\alpha)\\vee(\\alpha\\wedge\\neg\\alpha))=0\\]\nif and only if $\\nu(\\alpha\\uparrow\\neg\\alpha)=0$ and $\\nu(\\alpha\\wedge\\neg\\alpha)=0$; the first of these equalities holds if and only $\\nu(\\alpha)=1$ and $\\nu(\\neg\\alpha)=1$, what would imply that $\\nu(\\alpha\\wedge\\neg\\alpha)=1$, reaching a contradiction.\n\n\\item Take a bivaluation $\\nu$ for $\\nbIci$ and suppose \n\\[\\nu(\\neg(\\alpha\\uparrow\\neg\\alpha)\\rightarrow(\\alpha\\wedge\\neg\\alpha))=0,\\]\nwhat happens if and only if $\\nu(\\neg(\\alpha\\uparrow\\neg\\alpha))=1$ but $\\nu(\\alpha\\wedge\\neg\\alpha)=0$; from the first equality, $\\nu(\\alpha)=\\nu(\\neg\\alpha)=1$, what again contradicts $\\nu(\\alpha\\wedge\\neg\\alpha)=0$.\n\n\\item Finally, we take a bivaluation $\\nu$ for $\\nbIcl$ and assume it is possible to have\n\\[\\nu(\\neg(\\alpha\\wedge\\neg\\alpha)\\rightarrow(\\alpha\\uparrow\\neg\\alpha))=0.\\]\nThis is equivalent to having $\\nu(\\neg(\\alpha\\wedge\\neg\\alpha))=1$ and $\\nu(\\alpha\\uparrow\\neg\\alpha)=0$, but the first equality implies $\\nu(\\alpha\\uparrow\\neg\\alpha)=1$, which is evidently contradictory.\n\\end{enumerate}\n\n\\begin{theorem}\nGiven formulas $\\Gamma\\cup\\{\\varphi\\}$ over the signature $\\Sigma_{\\nbI}$ and $\\mathcal{L}\\in\\{\\nbIciw, \\nbIci, \\nbIcl\\}$, if $\\Gamma\\vdash_{\\mathcal{L}}\\varphi$ then $\\Gamma\\vDash_{\\mathcal{L}}\\varphi$.\n\\end{theorem}\n\nTo prove the reciprocal, take a set of formulas $\\Gamma$ maximal with respect to not proving $\\varphi$ in $\\mathcal{L}\\in\\{\\nbIciw, \\nbIci, \\nbIcl\\}$ which is also closed (again, with respect to $\\mathcal{L}$) and non-trivial. We define $\\nu:F(\\Sigma_{\\nbI},\\mathcal{V})\\rightarrow\\{0,1\\}$ such that $\\nu(\\gamma)=1$ if and only if $\\gamma\\in\\Gamma$, and wish to prove that $\\nu$ is a bivaluation for $\\mathcal{L}$.\n\nIt is clear $\\Gamma$ does not prove $\\varphi$ in $\\nbI$, since $\\mathcal{L}$ extends $\\nbI$, and also that $\\Gamma$ is closed in $\\nbI$, given it is closed according to $\\mathcal{L}$ and this logic is stronger than the previous; this implies $\\nu$ is at least a bivaluation for $\\nbI$, according to the definition found in Section \\ref{Bivaluations for nbI}.\n\n\\begin{enumerate}\n\\item For any of the logics $\\mathcal{L}\\in\\{\\nbIciw, \\nbIci, \\nbIcl\\}$, if $\\nu(\\alpha\\uparrow\\neg\\alpha)=1$, suppose $\\nu(\\alpha)=1$ and $\\nu(\\neg\\alpha)=1$: by definition of $\\nu$, this means $\\alpha\\uparrow\\neg\\alpha, \\alpha, \\neg\\alpha\\in\\Gamma$, and since $\\Gamma$ is closed and \n\\[(\\alpha\\uparrow\\neg\\alpha)\\rightarrow(\\alpha\\rightarrow(\\neg\\alpha\\rightarrow\\varphi))\\]\nis an instance of $\\textbf{Ip}$, we get that $\\varphi\\in\\Gamma$, what is a contradiction, We must then have either $\\nu(\\alpha)=0$ or $\\nu(\\neg\\alpha)=0$.\n\nSuppose, reciprocally, that $\\nu(\\alpha)=0$ or $\\nu(\\neg\\alpha)=0$, and for a proof by contradiction, let us take $\\nu(\\alpha\\uparrow\\neg\\alpha)=0$, meaning $\\alpha\\uparrow\\neg\\alpha\\notin\\Gamma$ and either $\\alpha\\notin\\Gamma$ or $\\neg\\alpha\\notin\\Gamma$, implying $\\alpha\\wedge\\neg\\alpha\\notin\\Gamma$. By the maximality of $\\Gamma$, one has $\\Gamma, \\alpha\\uparrow\\neg\\alpha\\vdash_{\\mathcal{L}}\\varphi$ and $\\Gamma,\\alpha\\wedge\\neg\\alpha\\vdash_{\\mathcal{L}}\\varphi$, and by a proof by cases\n\\[\\Gamma, (\\alpha\\uparrow\\neg\\alpha)\\vee(\\alpha\\wedge\\neg\\alpha)\\vdash_{\\mathcal{L}}\\varphi.\\]\nSince $(\\alpha\\uparrow\\neg\\alpha)\\vee(\\alpha\\wedge\\neg\\alpha)$ is an instance of $\\textbf{ciw}^{*}$, if $\\mathcal{L}=\\nbIciw$ we get $\\Gamma\\vdash_{\\nbIciw}\\varphi$, a contradiction; for the other two logics, given that, from Sections \\ref{ciuparrow} and \\ref{cluparrow},\n\\[\\neg(\\alpha\\uparrow\\beta)\\rightarrow(\\alpha\\wedge\\beta)\\vdash_{\\nbI}(\\alpha\\uparrow\\beta)\\vee(\\alpha\\wedge\\beta)\\]\nand\n\\[\\neg(\\alpha\\wedge\\beta)\\rightarrow(\\alpha\\uparrow\\beta)\\vdash_{\\nbI}(\\alpha\\uparrow\\beta)\\vee(\\alpha\\wedge\\beta),\\]\nby replacing $\\beta$ with $\\neg\\alpha$ we obtain that we still have $\\Gamma\\vdash_{\\mathcal{L}}\\varphi$, so still a contradiction. This means we must have $\\nu(\\alpha\\uparrow\\neg\\alpha)=1$.\n\n\\item If $\\mathcal{L}=\\nbIci$ and $\\nu(\\neg(\\alpha\\uparrow\\neg\\alpha))=1$, this means $\\neg(\\alpha\\uparrow\\neg\\alpha)\\in\\Gamma$; since $\\neg(\\alpha\\uparrow\\neg\\alpha)\\rightarrow(\\alpha\\wedge\\neg\\alpha)$ is an instance of $\\textbf{ci}^{*}$ and $\\Gamma$ is closed, $\\alpha\\wedge\\neg\\alpha\\in\\Gamma$, implying that $\\alpha, \\neg\\alpha\\in\\Gamma$ and therefore $\\nu(\\alpha)=1$ and $\\nu(\\neg\\alpha)=1$.\n\n\\item If $\\mathcal{L}=\\nbIcl$ and $\\nu(\\neg(\\alpha\\wedge\\neg\\alpha))=1$, we have $\\neg(\\alpha\\wedge\\neg\\alpha)\\in\\Gamma$; since $\\neg(\\alpha\\uparrow\\neg\\alpha)\\rightarrow(\\alpha\\wedge\\neg\\alpha)$ is an instance of $\\textbf{cl}^{*}$ and $\\Gamma$ is closed, $\\alpha\\uparrow\\neg\\alpha\\in\\Gamma$, and therefore $\\nu(\\alpha\\uparrow\\neg\\alpha)=1$.\n\\end{enumerate}\n\n\\begin{theorem}\nGiven formulas $\\Gamma\\cup\\{\\varphi\\}$ over the signature $\\Sigma_{\\nbI}$ and $\\mathcal{L}\\in\\{\\nbIciw, \\nbIci, \\nbIcl\\}$, if $\\Gamma\\vDash_{\\mathcal{L}}\\varphi$ then $\\Gamma\\vdash_{\\mathcal{L}}\\varphi$.\n\\end{theorem}\n\nIn the spirit of the arguments just used, since\n\\[\\neg(\\alpha\\uparrow\\beta)\\rightarrow(\\alpha\\wedge\\beta)\\vdash_{\\nbI}(\\alpha\\uparrow\\beta)\\vee(\\alpha\\wedge\\beta)\\]\nand\n\\[\\neg(\\alpha\\wedge\\beta)\\rightarrow(\\alpha\\uparrow\\beta)\\vdash_{\\nbI}(\\alpha\\uparrow\\beta)\\vee(\\alpha\\wedge\\beta),\\]\nby replacing $\\beta$ with $\\neg\\alpha$ we discover both $\\nbIci$ and $\\nbIcl$ are extensions of $\\nbIciw$, and with the aid of bivaluations, we can prove even more.\n\n\\begin{proposition}\n\\begin{enumerate}\n\\item $\\nbIciw$ can not prove $\\textbf{ci}^{*}$, and therefore $\\nbIci$ is strictly stronger than $\\nbIciw$;\n\\item $\\nbIciw$ can not prove $\\textbf{cl}^{*}$, and therefore $\\nbIcl$ is strictly stronger than $\\nbIciw$.\n\\end{enumerate}\n\\end{proposition}\n\n\\begin{proof}\n\\begin{enumerate}\n\\item Take a bivaluation $\\nu$ for $\\nbIciw$ such that $\\nu(\\alpha)=0$, $\\nu(\\neg\\alpha)=0$ (and therefore $\\nu(\\alpha\\wedge\\neg\\alpha)=0$), $\\nu(\\alpha\\uparrow\\neg\\alpha)=1$ and $\\nu(\\neg(\\alpha\\uparrow\\neg\\alpha))=1$: then\n\\[\\nu(\\neg(\\alpha\\uparrow\\neg\\alpha)\\rightarrow(\\alpha\\wedge\\neg\\alpha))=0.\\]\n\\item Take a bivaluation $\\nu$ for $\\nbIciw$ such that $\\nu(\\alpha)=1$, $\\nu(\\neg\\alpha)=1$ (and therefore $\\nu(\\alpha\\wedge\\neg\\alpha)=1$), $\\nu(\\alpha\\uparrow\\neg\\alpha)=0$ and $\\nu(\\neg(\\alpha\\wedge\\neg\\alpha))=1$: then \n\\[\\nu(\\neg(\\alpha\\wedge\\neg\\alpha)\\rightarrow(\\alpha\\uparrow\\neg\\alpha))=0.\\]\n\\end{enumerate}\n\\end{proof}\n\n\\begin{theorem}\n\\begin{enumerate}\n\\item The formula $(\\alpha\\uparrow\\neg\\alpha)\\uparrow\\neg(\\alpha\\uparrow\\neg\\alpha)$ is a tautology of $\\nbIci$;\n\\item $\\nbIci$ is obtained from $\\nbIciw$ by adding the axiom schema\\label{cc*}\n\\[\\tag{$\\textbf{cc}^{*}$} (\\alpha\\uparrow\\neg\\alpha)\\uparrow\\neg(\\alpha\\uparrow\\neg\\alpha).\\]\n\\end{enumerate}\n\\end{theorem}\n\n\\begin{proof}\n\\begin{enumerate}\n\\item It is clear that\n\\[(\\alpha\\uparrow\\neg\\alpha)\\uparrow\\neg(\\alpha\\uparrow\\neg\\alpha)\\vdash_{\\nbIci}(\\alpha\\uparrow\\neg\\alpha)\\uparrow\\neg(\\alpha\\uparrow\\neg\\alpha),\\]\nbut from the instance\n\\[\\neg[(\\alpha\\uparrow\\neg\\alpha)\\uparrow\\neg(\\alpha\\uparrow\\neg\\alpha)]\\rightarrow[(\\alpha\\uparrow\\neg\\alpha)\\wedge\\neg(\\alpha\\uparrow\\neg\\alpha)]\\]\nof $\\textbf{ci}^{*}$ we see that\n\\[\\neg[(\\alpha\\uparrow\\neg\\alpha)\\uparrow\\neg(\\alpha\\uparrow\\neg\\alpha)]\\vdash_{\\nbIci}(\\alpha\\uparrow\\neg\\alpha)\\wedge\\neg(\\alpha\\uparrow\\neg\\alpha).\\]\n$(\\alpha\\uparrow\\neg\\alpha)\\wedge\\neg(\\alpha\\uparrow\\neg\\alpha)$ easily deduces both $\\alpha\\uparrow\\neg\\alpha$ and $\\neg(\\alpha\\uparrow\\neg\\alpha)$, and from this last formula and the instance $\\neg(\\alpha\\uparrow\\neg\\alpha)\\rightarrow(\\alpha\\wedge\\neg\\alpha)$ of $\\textbf{ci}^{*}$, we get $\\neg[(\\alpha\\uparrow\\neg\\alpha)\\uparrow\\neg(\\alpha\\uparrow\\neg\\alpha)]$ deduces $\\alpha\\uparrow\\neg\\alpha$, $\\alpha$ and $\\neg\\alpha$: from the instance of $\\textbf{Ip}$\n\\[(\\alpha\\uparrow\\neg\\alpha)\\rightarrow\\big[\\alpha\\rightarrow\\big[\\neg\\alpha\\rightarrow\\big((\\alpha\\uparrow\\neg\\alpha)\\uparrow\\neg(\\alpha\\uparrow\\neg\\alpha)\\big)\\big]\\big],\\]\nthis means \n\\[\\neg[(\\alpha\\uparrow\\neg\\alpha)\\uparrow\\neg(\\alpha\\uparrow\\neg\\alpha)]\\vdash_{\\nbIci}(\\alpha\\uparrow\\neg\\alpha)\\uparrow\\neg(\\alpha\\uparrow\\neg\\alpha),\\]\nand by a proof by cases \n\\[[(\\alpha\\uparrow\\neg\\alpha)\\uparrow\\neg(\\alpha\\uparrow\\neg\\alpha)]\\vee\\neg[(\\alpha\\uparrow\\neg\\alpha)\\uparrow\\neg(\\alpha\\uparrow\\neg\\alpha)]\\vdash_{\\nbIci}(\\alpha\\uparrow\\neg\\alpha)\\uparrow\\neg(\\alpha\\uparrow\\neg\\alpha).\\]\nSince the antecedent in this argument is an instance of $\\textbf{Ax\\: 11}^{*}$, we obtain that $(\\alpha\\uparrow\\neg\\alpha)\\uparrow\\neg(\\alpha\\uparrow\\neg\\alpha)$ is a tautology.\n\n\\item So we have that $\\nbIci$ proves any instance of $\\textbf{ciw}^{*}$ and $\\textbf{cc}^{*}$, being therefore stronger than the logic obtained from $\\nbIciw$ by addition of $\\textbf{cc}^{*}$, which we will briefly call $\\nbIciw+$. To prove that $\\nbIciw+$ is as strong as $\\nbIci$, and therefore that both are equal, we only need to prove that any instance of $\\textbf{ci}^{*}$ is a tautology in $\\nbIciw+$, or what is equivalent, that\n\\[\\neg(\\alpha\\uparrow\\neg\\alpha)\\vdash_{\\nbIciw+}\\alpha\\wedge\\neg\\alpha.\\]\nObviously\n\\[\\alpha\\wedge\\neg\\alpha, \\neg(\\alpha\\uparrow\\neg\\alpha)\\vdash_{\\nbIciw+}\\alpha\\wedge\\neg\\alpha;\\]\nnow, from the instance\n\\[[(\\alpha\\uparrow\\neg\\alpha)\\uparrow\\neg(\\alpha\\uparrow\\neg\\alpha)]\\rightarrow\\big[(\\alpha\\uparrow\\neg\\alpha)\\rightarrow\\big[\\neg(\\alpha\\uparrow\\neg\\alpha)\\rightarrow(\\alpha\\wedge\\neg\\alpha)\\big]\\big]\\]\nof $\\textbf{Ip}$ and the fact $(\\alpha\\uparrow\\neg\\alpha)\\uparrow\\neg(\\alpha\\uparrow\\neg\\alpha)$ is an instance of $\\textbf{cc}^{*}$, we can see by applying the deduction meta-theorem as needed that\n\\[\\alpha\\uparrow\\neg\\alpha, \\neg(\\alpha\\uparrow\\neg\\alpha)\\vdash_{\\nbIciw+}\\alpha\\wedge\\neg\\alpha.\\]\n\nBy a proof by cases,\n\\[(\\alpha\\uparrow\\neg\\alpha)\\vee(\\alpha\\wedge\\neg\\alpha), \\neg(\\alpha\\uparrow\\neg\\alpha)\\vdash_{\\nbIciw+}\\alpha\\wedge\\neg\\alpha,\\]\nand since $(\\alpha\\uparrow\\neg\\alpha)\\vee(\\alpha\\wedge\\neg\\alpha)$ is an instance of $\\textbf{ciw}^{*}$, we discover that $\\nbIci$ is $\\nbIciw+$, as we wanted to show.\n\\end{enumerate}\n\\end{proof}\n\n\n\n\n\n\n\n\\subsection{Fidel structures}\n\nA \\index{Fidel structure for $\\nbIciw$}\\index{Fidel structure for $\\nbIci$}\\index{Fidel structure for $\\nbIcl$}Fidel structure, presented as a $\\Sigma_{\\nbI}^{\\textbf{CPL}}$-multialgebra, for $\\mathcal{L}\\in\\{\\nbIciw, \\nbIci, \\nbIcl\\}$ is any $\\Sigma_{\\nbI}^{\\textbf{CPL}}$-multialgebra $\\mathcal{A}=(A,\\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma_{\\nbI}^{\\textbf{CPL}}})$ that is a Fidel structure for $\\nbI$ and satisfies, additionally:\n\\begin{enumerate}\n\\item if $b\\in\\neg a$, ${\\sim}(a\\wedge b)\\in a\\uparrow b$;\n\\item\\begin{enumerate}\n\\item if $\\mathcal{L}=\\nbIci$ and $b\\in\\neg a$, $a\\wedge b\\in\\neg{\\sim}(a\\wedge b)$;\n\\item if $\\mathcal{L}=\\nbIcl$ and $b\\in\\neg a$, ${\\sim}(a\\wedge b)\\in\\neg(a\\wedge b)$.\n\\end{enumerate}\n\\end{enumerate}\n\n\nA valuation for a Fidel structure $\\mathcal{A}$, presented as a $\\Sigma_{\\nbI}^{\\textbf{CPL}}$-multialgebra, for $\\mathcal{L}\\in\\{\\nbIciw, \\nbIci, \\nbIcl\\}$, will still be any $\\Sigma_{\\nbI}$-homomorphism $\\nu:\\textbf{F}(\\Sigma_{\\nbI},\\mathcal{V})\\rightarrow\\mathcal{A}$; and now we will consider the restricted Nmatrices $(\\mathcal{A},\\{\\top\\},\\mathcal{F}_{\\mathcal{A}}^{\\mathcal{L}})$, where $\\mathcal{A}$ is a Fidel structure for $\\mathcal{L}$ and $\\mathcal{F}_{\\mathcal{A}}^{\\mathcal{L}}$ is the set of valuations $\\nu:\\textbf{F}(\\Sigma_{\\nbI},\\mathcal{V})\\rightarrow\\mathcal{A}$ such that, for any formulas $\\alpha$ and $\\beta$ over the signature $\\Sigma_{\\nbI}$:\n\\begin{enumerate}\n\\item $\\nu(\\alpha\\uparrow\\beta)=\\nu(\\beta\\uparrow\\alpha)$;\n\\item $\\nu(\\alpha\\uparrow\\neg\\alpha)={\\sim}(\\nu(\\alpha)\\wedge\\nu(\\neg\\alpha))$;\n\\item \\begin{enumerate}\n\\item if $\\mathcal{L}=\\nbIci$, $\\nu(\\neg(\\alpha\\uparrow\\neg\\alpha))=\\nu(\\alpha)\\wedge\\nu(\\neg\\alpha)$;\n\\item if $\\mathcal{L}=\\nbIcl$, $\\nu(\\alpha\\uparrow\\neg\\alpha)=\\nu(\\neg(\\alpha\\wedge\\neg\\alpha))$.\n\\end{enumerate}\n\\end{enumerate}\n\n\\begin{proposition}\nThe RNmatrices $(\\mathcal{A}, \\{\\top\\}, \\mathcal{F}_{\\mathcal{A}}^{\\mathcal{L}})$ as described above, for $\\mathcal{L}\\in$\\\\$\\{\\nbIciw, \\nbIci, \\nbIcl\\}$, are structural.\n\\end{proposition}\n\n\\begin{proof}\nLet $\\sigma:\\textbf{F}(\\Sigma_{\\nbI},\\mathcal{V})\\rightarrow\\textbf{F}(\\Sigma_{\\nbI},\\mathcal{V})$ be a $\\Sigma_{\\nbI}$-homomorphism and $\\nu\\in\\mathcal{F}_{\\mathcal{A}}^{\\mathcal{L}}$: it is clear that $\\nu\\circ\\sigma: \\textbf{F}(\\Sigma_{\\nbI},\\mathcal{V})\\rightarrow\\mathcal{A}$ is always a $\\Sigma_{\\nbI}$-homomorphism, satisfying additionally that for all formulas $\\alpha$ and $\\beta$ on the signature $\\Sigma_{\\nbI}$, $\\nu\\circ\\sigma(\\alpha\\uparrow\\beta)=\\nu\\circ\\sigma(\\beta\\uparrow\\alpha)$.\n\n\\begin{enumerate}\n\\item We must prove that $\\nu\\circ\\sigma(\\alpha\\uparrow\\neg\\alpha)={\\sim}(\\nu\\circ\\sigma(\\alpha)\\wedge\\nu\\circ\\sigma(\\neg\\alpha))$, what is easy since\n\\[\\nu\\circ\\sigma(\\alpha\\uparrow\\neg\\alpha)=\\nu(\\sigma(\\alpha\\uparrow\\neg\\alpha))=\\nu(\\sigma(\\alpha)\\uparrow\\sigma(\\neg\\alpha))=\\nu(\\sigma(\\alpha)\\uparrow\\neg\\sigma(\\alpha))=\\]\n\\[{\\sim}\\big(\\nu(\\sigma(\\alpha))\\wedge\\neg\\nu(\\sigma(\\alpha))\\big)={\\sim}(\\nu\\circ\\sigma(\\alpha)\\wedge\\nu\\circ\\sigma(\\neg\\alpha)).\\]\n\n\\item\\begin{enumerate}\n\\item If $\\mathcal{L}=\\nbIci$, it remains for us to show that $\\nu\\circ\\sigma(\\neg(\\alpha\\uparrow\\neg\\alpha))=\\nu\\circ\\sigma(\\alpha)\\wedge\\nu\\circ\\sigma(\\neg\\alpha)$, and we see that\n\\[\\nu\\circ\\sigma(\\neg(\\alpha\\uparrow\\neg\\alpha))=\\nu\\big(\\neg(\\sigma(\\alpha)\\uparrow\\neg\\sigma(\\alpha))\\big)=\\nu(\\sigma(\\alpha))\\wedge\\nu(\\neg\\sigma(\\alpha))=\\nu\\circ\\sigma(\\alpha)\\wedge\\nu\\circ\\sigma(\\neg\\alpha).\\]\n\n\\item If $\\mathcal{L}=\\nbIcl$, we only need to prove that $\\nu\\circ\\sigma(\\alpha\\uparrow\\neg\\alpha)=\\nu\\circ\\sigma(\\neg(\\alpha\\wedge\\neg\\alpha))$:\n\\[\\nu\\circ\\sigma(\\alpha\\uparrow\\neg\\alpha)=\\nu(\\sigma(\\alpha)\\uparrow\\neg\\sigma(\\alpha))=\\nu\\big(\\neg(\\sigma(\\alpha)\\wedge\\neg\\sigma(\\alpha))\\big)=\\nu\\circ\\sigma(\\neg(\\alpha\\wedge\\neg\\alpha)).\\]\n\\end{enumerate}\n\\end{enumerate}\n\\end{proof}\n\nGiven formulas $\\Gamma\\cup\\{\\varphi\\}$ over the signature $\\Sigma_{\\nbI}$, if $\\Gamma$ proves $\\varphi$ according to such restricted Nmatrices we will write \\label{VdashFnbIciw}\\label{VdashFnbIci}\\label{VdashFnbIcl}$\\Gamma\\Vdash_{\\mathcal{F}}^{\\mathcal{L}}\\varphi$, and say that $\\Gamma$ proves $\\varphi$ according to Fidel structures for $\\mathcal{L}$.\n\nIt is easy to see how \"$\\Vdash_{\\mathcal{F}}^{\\mathcal{L}}$\" models the axiom schemata of $\\nbI$ and its rules of inference, but we can also prove \"$\\Vdash_{\\mathcal{F}}^{\\nbIciw}$\" proves any instance of $\\textbf{ciw}^{*}$, \"$\\Vdash_{\\mathcal{F}}^{\\nbIci}$\" proves any instance of $\\textbf{ci}^{*}$ and \"$\\Vdash_{\\mathcal{F}}^{\\nbIcl}$\" proves any instance of $\\textbf{cl}^{*}$, so take a formula $\\alpha$ over the signature $\\Sigma_{\\nbI}$.\n\n\\begin{enumerate}\n\\item Given an instance $(\\alpha\\uparrow\\neg\\alpha)\\vee(\\alpha\\wedge\\neg\\alpha)$ of $\\textbf{ciw}^{*}$, we have for a Fidel structure $\\mathcal{A}$ for $\\nbIciw$ and a $\\nu\\in \\mathcal{F}_{\\mathcal{A}}^{\\nbIciw}$ that\n\\[\\nu((\\alpha\\uparrow\\neg\\alpha)\\vee(\\alpha\\wedge\\neg\\alpha))=\\nu(\\alpha\\uparrow\\neg\\alpha)\\vee\\nu(\\alpha\\wedge\\neg\\alpha)=\\]\n\\[{\\sim}(\\nu(\\alpha)\\wedge\\nu(\\neg\\alpha))\\vee(\\nu(\\alpha)\\wedge\\nu(\\neg\\alpha))=({\\sim}\\nu(\\alpha)\\vee{\\sim}\\nu(\\neg\\alpha))\\vee(\\nu(\\alpha)\\wedge\\nu(\\neg\\alpha))=\\]\n\\[\\big[({\\sim}\\nu(\\alpha)\\vee{\\sim}\\nu(\\neg\\alpha))\\vee\\nu(\\alpha)\\big]\\wedge\\big[({\\sim}\\nu(\\alpha)\\vee{\\sim}\\nu(\\neg\\alpha))\\vee\\nu(\\neg\\alpha)\\big]=\\]\n\\[[{\\sim}\\nu(\\neg\\alpha)\\vee\\top]\\wedge[{\\sim}\\nu(\\alpha)\\vee\\top]=\\top\\wedge\\top=\\top.\\]\n\n\\item Take an instance $\\neg(\\alpha\\uparrow\\neg\\alpha)\\rightarrow(\\alpha\\wedge\\neg\\alpha)$ of $\\textbf{ci}^{*}$, a Fidel structure $\\mathcal{A}$ for $\\nbIci$ and a $\\nu\\in\\mathcal{F}_{\\mathcal{A}}^{\\nbIci}$:\n\\[\\nu(\\neg(\\alpha\\uparrow\\neg\\alpha)\\rightarrow(\\alpha\\wedge\\neg\\alpha))=\\nu(\\neg(\\alpha\\uparrow\\neg\\alpha))\\rightarrow\\nu(\\alpha\\wedge\\neg\\alpha)=\\]\n\\[\\big[\\nu(\\alpha)\\wedge\\nu(\\neg\\alpha)\\big]\\rightarrow\\big[\\nu(\\alpha)\\wedge\\nu(\\neg\\alpha)\\big]=\\top.\\]\n\n\\item For $\\neg(\\alpha\\wedge\\neg\\alpha)\\rightarrow(\\alpha\\uparrow\\neg\\alpha)$ an instance of $\\textbf{cl}^{*}$, a Fidel structure $\\mathcal{A}$ for $\\nbIcl$ and a $\\nu\\in\\mathcal{F}_{\\mathcal{A}}^{\\nbIcl}$,\n\\[\\nu(\\neg(\\alpha\\wedge\\neg\\alpha)\\rightarrow(\\alpha\\uparrow\\neg\\alpha))=\\nu(\\neg(\\alpha\\wedge\\neg\\alpha))\\rightarrow\\nu(\\alpha\\uparrow\\neg\\alpha)=\\nu(\\alpha\\uparrow\\neg\\alpha)\\rightarrow\\nu(\\alpha\\uparrow\\neg\\alpha)=\\top.\\]\n\\end{enumerate}\n\n\\begin{theorem}\nGiven formulas $\\Gamma\\cup\\{\\varphi\\}$ over the signature $\\Sigma_{\\nbI}$, for any $\\mathcal{L}\\in$\\\\$\\{\\nbIciw, \\nbIci, \\nbIcl\\}$ we have that, if $\\Gamma\\vdash_{\\mathcal{L}}\\varphi$, then $\\Gamma\\vDash_{\\mathcal{F}}^{\\mathcal{L}}\\varphi$.\n\\end{theorem}\n\nReciprocally, we define, for a logic $\\mathcal{L}\\in\\{\\nbIciw,\\nbIci,\\nbIcl\\}$ and a set of formulas $\\Gamma$ over the signature $\\Sigma_{\\nbI}$, the equivalence relation such that, for formulas $\\alpha$ and $\\beta$ still over the signature $\\Sigma_{\\nbI}$, $\\alpha\\equiv^{\\mathcal{L}}_{\\Gamma}\\beta$ if and only if $\\Gamma\\vdash_{\\mathcal{L}}\\alpha\\rightarrow\\beta$ and $\\Gamma\\vdash_{\\mathcal{L}}\\beta\\rightarrow\\alpha$.\n\nAs we did several times before, the well-defined quotient $A^{\\mathcal{L}}_{\\Gamma}=F(\\Sigma_{\\nbI}, \\mathcal{V})\/\\equiv^{\\mathcal{L}}_{\\Gamma}$ is made into a Boolean algebra, and by defining, for classes of formulas $[\\alpha]$ and $[\\beta]$, $\\neg\\alpha=$\\\\$\\{[\\neg\\varphi]\\ :\\ \\varphi\\in[\\alpha]\\}$ and\n\\begin{enumerate}\n\\item if $[\\beta]\\in \\neg[\\alpha]$,\n\\[[\\alpha]\\uparrow[\\beta]=[{\\sim}(\\alpha\\wedge\\beta)];\\]\n\\item otherwise,\n\\[[\\alpha]\\uparrow[\\beta]=\\{[\\varphi\\uparrow\\psi]\\ :\\ \\varphi\\in[\\alpha], \\psi\\in[\\beta]\\};\\]\n\\end{enumerate}\nwe shall prove $\\mathcal{A}^{\\mathcal{L}}_{\\Gamma}=(A^{\\mathcal{L}}_{\\Gamma},\\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in \\Sigma_{\\nbI}^{\\textbf{CPL}}})$ is a Fidel structure for $\\mathcal{L}$, which we shall call the Lindenbaum-Tarski Fidel structure of $\\mathcal{L}$ associated to $\\Gamma$.\n\nQuite clearly $(A^{\\mathcal{L}}_{\\Gamma},\\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma^{\\textbf{CPL}}})$ is a Boolean algebra; concerning the paraconsistent negation, for every $[\\beta]\\in\\neg[\\alpha]$, there exists $\\varphi\\equiv^{\\mathcal{L}}_{\\Gamma}\\alpha$ such that $\\beta\\equiv^{\\mathcal{L}}_{\\Gamma}\\neg\\varphi$ and then $[\\alpha]\\vee[\\beta]=[\\alpha\\vee\\beta]=[\\varphi\\vee\\neg\\varphi]=\\top$; regarding the incompatibility connective, we begin by proving the operation is well-defined in the case that $[\\beta]\\in\\neg[\\alpha]$.\n\nIf $\\varphi\\in[\\alpha]$ and $\\psi\\in[\\beta]$, we have that $\\alpha\\equiv^{\\mathcal{L}}_{\\Gamma}\\varphi$ and $\\beta\\equiv^{\\mathcal{L}}_{\\Gamma}\\psi$, implying that ${\\sim}(\\alpha\\wedge\\beta)\\equiv^{\\mathcal{L}}_{\\Gamma}{\\sim}(\\varphi\\wedge\\psi)$ given \"$\\equiv^{\\mathcal{L}}_{\\Gamma}$\" is a congruence for the connectives in $\\{\\vee, \\wedge, \\rightarrow, {\\sim}\\}$, and therefore $[\\alpha]\\uparrow[\\beta]=[\\varphi]\\uparrow[\\psi]$. In the case of $[\\alpha]\\uparrow[\\beta]$ with $[\\beta]\\not\\in\\neg[\\alpha$, we use the same reasoning that worked for $\\bI$ in Section \\ref{Completeness fo Fidel structures for bI}.\n\nTo prove $\\mathcal{A}^{\\mathcal{L}}_{\\Gamma}$ is a Fidel structure for $\\nbI$, we are yet to prove that, if $[\\beta]\\in\\neg[\\alpha]$, for every value $[\\gamma]$ in $[\\alpha]\\uparrow[\\beta]$ (in this case, there is only one), we have $[\\alpha]\\wedge([\\beta]\\wedge[\\gamma])=\\bot$, the result being clear in the case that $[\\beta]\\notin\\neg[\\alpha]$. Since $\\gamma\\equiv^{\\mathcal{L}}_{\\Gamma}{\\sim}(\\alpha\\wedge\\beta)$,\n\\[\\alpha\\wedge(\\beta\\wedge\\gamma)\\equiv^{\\mathcal{L}}_{\\Gamma}\\alpha\\wedge(\\beta\\wedge{\\sim}(\\alpha\\wedge\\beta))\\equiv^{\\mathcal{L}}_{\\Gamma}\\bot,\\]\nand the result holds; the case in which $[\\alpha]\\in\\neg[\\beta]$ is analogous.\n\nIt only remains to be shown that $\\mathcal{A}^{\\mathcal{L}}_{\\Gamma}$ is a Fidel structure for $\\mathcal{L}$.\n\\begin{lemma}\\label{some equivalences in extensions of nbI}\n\\begin{enumerate}\n\\item For any logic $\\mathcal{L}\\in\\{\\nbIciw, \\nbIci, \\nbIcl\\}$, $\\alpha\\uparrow\\neg\\alpha$ and ${\\sim}(\\alpha\\wedge\\neg\\alpha)$ are equivalent.\n\\item $\\neg(\\alpha\\uparrow\\neg\\alpha)$ and $\\alpha\\wedge\\neg\\alpha$ are equivalent in $\\nbIci$.\n\\item $\\neg(\\alpha\\wedge\\neg\\alpha)$ and $\\alpha\\uparrow\\neg\\alpha$ are equivalent in $\\nbIcl$.\n\\end{enumerate}\n\\end{lemma}\n\n\\begin{proof}\n\\begin{enumerate}\n\\item For simplicity, we denote $\\bot_{\\alpha\\wedge\\neg\\alpha,\\alpha\\wedge\\neg\\alpha}$ simply by $\\bot$. Applying the deduction meta-theorem, we must prove $\\alpha\\uparrow\\neg\\alpha, \\alpha\\wedge\\neg\\alpha\\vdash_{\\mathcal{L}}\\bot$ and $(\\alpha\\wedge\\neg\\alpha)\\rightarrow\\bot\\vdash_{\\mathcal{L}}\\alpha\\uparrow\\neg\\alpha$.\n\nThe first implication is obvious, since $\\alpha\\uparrow\\neg\\alpha$ together with $\\alpha\\wedge\\neg\\alpha$ is exactly $\\bot_{\\alpha,\\neg\\alpha}$, and all bottom elements are equivalent to each other. Reciprocally, quite obviously\n\\[(\\alpha\\wedge\\neg\\alpha)\\rightarrow\\bot, \\alpha\\uparrow\\neg\\alpha\\vdash_{\\mathcal{L}}\\alpha\\uparrow\\neg\\alpha,\\]\nand\n\\[(\\alpha\\wedge\\neg\\alpha)\\rightarrow\\bot, \\alpha\\wedge\\neg\\alpha\\vdash_{\\mathcal{L}}\\alpha\\uparrow\\neg\\alpha\\]\nby Modus Ponens and the fact a bottom element implies any formula. With a proof by cases and the fact that $(\\alpha\\uparrow\\neg\\alpha)\\vee(\\alpha\\wedge\\neg\\alpha)$ is an instance of $\\textbf{ciw}^{*}$, which is a tautology in any of the logics $\\mathcal{L}$, we get the desired result.\n\n\\item We must prove $\\neg(\\alpha\\uparrow\\neg\\alpha)\\vdash_{\\nbIci}\\alpha\\wedge\\neg\\alpha$ and vice-versa, being the first direction a clear application of $\\textbf{ci}^{*}$. Reciprocally, we have \n\\[\\alpha\\wedge\\neg\\alpha, \\neg(\\alpha\\uparrow\\neg\\alpha)\\vdash_{\\nbIci}\\neg(\\alpha\\uparrow\\neg\\alpha)\\]\nwith easy, while \n\\[\\alpha\\wedge\\neg\\alpha, \\alpha\\uparrow\\neg\\alpha\\vdash_{\\nbIci}\\neg(\\alpha\\uparrow\\neg\\alpha)\\]\nfollows from the fact $\\alpha\\wedge\\neg\\alpha$ implies both $\\alpha$ and $\\neg\\alpha$, followed by an application of $\\textbf{Ip}$. With a proof by cases and the fact $(\\alpha\\uparrow\\neg\\alpha)\\vee\\neg(\\alpha\\uparrow\\neg\\alpha)$ is an instance of $\\textbf{Ax\\: 11}^{*}$, the result follows.\n\n\\item Now, we must prove $\\neg(\\alpha\\wedge\\neg\\alpha)\\vdash_{\\nbIcl}\\alpha\\uparrow\\neg\\alpha$ and its reciprocal, being the first implication a direct application of $\\textbf{cl}^{*}$. Reciprocally, \n\\[\\alpha\\uparrow\\neg\\alpha, \\neg(\\alpha\\wedge\\neg\\alpha)\\vdash_{\\nbIcl}\\neg(\\alpha\\wedge\\neg\\alpha)\\]\nand, from the fact $\\alpha\\wedge\\neg\\alpha$ implies both $\\alpha$ and $\\neg\\alpha$, and by applying $\\textbf{Ip}$,\n\\[\\alpha\\uparrow\\neg\\alpha, \\alpha\\wedge\\neg\\alpha\\vdash_{\\nbIcl}\\neg(\\alpha\\wedge\\neg\\alpha);\\]\nthrough a proof by cases and the fact $(\\alpha\\wedge\\neg\\alpha)\\vee\\neg(\\alpha\\wedge\\neg\\alpha)$ is an instance of $\\textbf{Ax\\: 11}^{*}$, we conclude the proof.\n\\end{enumerate}\n\\end{proof}\n\n\\begin{enumerate}\n\\item By definition, if $[\\beta]\\in\\neg[\\alpha]$, then $[\\alpha]\\uparrow[\\beta]$ is single-valued and equal to ${\\sim}([\\alpha]\\wedge[\\beta])$.\n\\item \\begin{enumerate}\n\\item If $\\mathcal{L}=\\nbIci$ and $[\\beta]\\in\\neg[\\alpha]$, implying that there exists $\\phi\\equiv^{\\nbIci}_{\\Gamma}\\alpha$ such that $\\beta\\equiv^{\\nbIci}_{\\Gamma}\\neg\\phi$, by use of Lemma \\ref{some equivalences in extensions of nbI} we find\n\\[[{\\sim}(\\alpha\\wedge\\beta)]=[{\\sim}(\\phi\\wedge\\neg\\phi)]=[\\phi\\uparrow\\neg\\phi],\\]\nand therefore $\\neg[{\\sim}(\\alpha\\wedge\\beta)]$ contains $[\\neg(\\phi\\uparrow\\neg\\phi)]=[\\phi\\wedge\\neg\\phi]$, which equals $[\\alpha]\\wedge[\\beta]$, as we wanted to show.\n\\item If $\\mathcal{L}=\\nbIcl$ and $[\\beta]\\in\\neg[\\alpha]$, let $\\phi\\equiv^{\\nbIcl}_{\\Gamma}\\alpha$ be an element of $[\\alpha]$ such that $\\beta\\equiv^{\\nbIcl}_{\\Gamma}\\neg\\phi$, and then $[\\alpha\\wedge\\beta]=[\\phi\\wedge\\neg\\phi]$, while $[\\phi\\uparrow\\neg\\phi]=[\\neg(\\phi\\wedge\\neg\\phi)]\\in \\neg[\\phi\\wedge\\neg\\phi]$. \n\nSince $[\\phi\\uparrow\\neg\\phi]=[{\\sim}(\\phi\\wedge\\neg\\phi)]$ and ${\\sim}([\\phi]\\wedge[\\neg\\phi])={\\sim}([\\alpha]\\wedge[\\beta])$, the result is done: ${\\sim}([\\alpha]\\wedge[\\beta])\\in \\neg([\\alpha]\\wedge[\\beta])$.\n\\end{enumerate}\n\\end{enumerate}\n\n\\begin{theorem}\nGiven formulas $\\Gamma\\cup\\{\\varphi\\}$ over the signature $\\Sigma_{\\nbI}$, for any logic $\\mathcal{L}\\in$\\\\$\\{\\nbIciw, \\nbIci, \\nbIcl\\}$ we have that, if $\\Gamma\\Vdash_{\\mathcal{F}}^{\\mathcal{L}}\\varphi$, then $\\Gamma\\vdash_{\\mathcal{L}}\\varphi$.\n\\end{theorem}\n\n\n\n\n\n\n\n\n\\subsection{Decision method}\\label{Decision method for nbIciw, ...}\n\nWe take the Boolean algebra $\\textbf{2}$ and extend it to the $\\Sigma^{\\textbf{CPL}}_{\\textbf{nbI}}$-multialgebra $\\textbf{2}_{\\textbf{nbIciw}}$ with operations given by the tables below.\n\n\\begin{figure}[H]\n\\centering\n\\begin{minipage}[t]{4cm}\n\\centering\n\\begin{tabular}{|l|r|}\n\\hline\n& $\\neg$\\\\ \\hline\n$0$ & $\\{1\\}$\\\\ \\hline\n$1$ & $\\{0,1\\}$\\\\\\hline\n\\end{tabular}\n\\caption*{Negation}\n\\end{minipage}\n\\hspace{3cm}\n\\centering\n\\begin{minipage}[t]{4cm}\n\\centering\n\\begin{tabular}{|l|c|r|}\n\\hline\n$\\uparrow$ & $0$ & $1$\\\\ \\hline\n$0$ & $\\{0,1\\}$ & $\\{0, 1\\}$\\\\ \\hline\n$1$ & $\\{0, 1\\}$ & $\\{0\\}$\\\\\\hline\n\\end{tabular}\n\\caption*{Incompatibility}\n\\end{minipage}\n\\end{figure}\n\nIt is easy to see that:\n\\begin{enumerate}\n\\item $(\\{0,1\\}, \\{\\sigma_{\\textbf{2}_{\\textbf{nbIciw}}}\\}_{\\sigma\\in\\Sigma^{\\textbf{CPL}}})$ is a Boolean algebra (that is, $\\textbf{2}$);\n\\item for any $x,y\\in\\{0,1\\}$ and $z\\in x\\uparrow y$, $x\\wedge(y\\wedge z)=0$;\n\\item for any $x\\in \\{0,1\\}$ and $y\\in\\neg x$, $x\\vee y=1$, since in the case that $x=0$ we must have $y=1$.\n\\end{enumerate}\n\nSo $\\textbf{2}_{\\textbf{nbIciw}}$ is a Fidel structure for $\\textbf{nbI}$. But we can prove even more, that it is a Fidel structure for $\\mathcal{L}\\in\\{\\textbf{nbIciw}, \\textbf{nbIci}, \\textbf{nbIcl}\\}$.\n\n\\begin{enumerate}\n\\item We see that, for any values of $x$ and $y$ in $\\{0,1\\}$, ${\\sim}(x\\wedge y)\\in x\\uparrow y$ even if $y\\not\\in\\neg x$;\n\\item\n\\begin{enumerate}\n\\item suppose that, for some values $a$ and $b$ in $\\{0,1\\}$ such that $b\\in\\neg a$, $a\\wedge b$ is not in $\\neg{\\sim}(a\\wedge b)$, and since $\\neg 1=\\{0,1\\}$, we must then have ${\\sim}(a\\wedge b)=0$, that is, $a\\wedge b=1$ and therefore $a=b=1$; but then $\\neg{\\sim}(a\\wedge b)=\\{1\\}$, which contains exactly the value of $a\\wedge b=1$, leading to a contradiction; so, to summarize, for any values $a, b\\in\\{0,1\\}$ we have $a\\wedge b\\in\\neg{\\sim}(a\\wedge b)$;\n\\item suppose that, for some values $a$ and $b$ in $\\{0,1\\}$ such that $b\\in\\neg a$, ${\\sim}(a\\wedge b)$ is not in $\\neg(a\\wedge b)$, which means $a\\wedge b=0$ (since $\\neg 1=\\{0,1\\}$); but then ${\\sim}(a\\wedge b)=1$ and $\\neg(a\\wedge b)=\\{1\\}$, what is a contradiction; so, for any values $a,b\\in\\{0,1\\}$, ${\\sim}(a\\wedge b)\\in\\neg(a\\wedge b)$.\n\\end{enumerate}\n\\end{enumerate}\n\nFinally, we can define, for $\\mathcal{L}\\in\\{\\textbf{nbIciw}, \\textbf{nbIci}, \\textbf{nbIcl}\\}$, the restricted Nmatrix \n\\[\\mathbb{2}_{\\mathcal{L}}=(\\textbf{2}_{\\textbf{nbIciw}}, \\{1\\}, \\mathcal{F}_{\\textbf{2}_{\\textbf{nbIciw}}}^{\\mathcal{L}})\\]\nsuch that $\\mathcal{F}_{\\textbf{2}_{\\textbf{nbIciw}}}^{\\mathcal{L}}$ is the set of homomorphisms $\\nu:\\textbf{F}(\\Sigma_{\\textbf{nbI}}, \\mathcal{V})\\rightarrow \\textbf{2}_{\\textbf{nbIciw}}$ satisfying that, for any formulas $\\alpha$ and $\\beta$:\n\\begin{enumerate}\n\\item $\\nu(\\alpha\\uparrow\\beta)=\\nu(\\beta\\uparrow\\alpha)$;\n\\item $\\nu(\\alpha\\uparrow\\neg\\alpha)={\\sim}(\\nu(\\alpha)\\wedge\\nu(\\neg\\alpha))$;\n\\item\\begin{enumerate}\n\\item if $\\mathcal{L}=\\textbf{nbIci}$, $\\nu(\\neg(\\alpha\\uparrow\\neg\\alpha))=\\nu(\\alpha)\\wedge\\nu(\\neg\\alpha)$;\n\\item if $\\mathcal{L}=\\textbf{nbIcl}$, $\\nu(\\alpha\\uparrow\\neg\\alpha)=\\nu(\\neg(\\alpha\\wedge\\neg\\alpha))$.\n\\end{enumerate}\n\\end{enumerate}\nClearly such RNmatrices are structural.\n\n\\begin{theorem}\n$\\nu: F(\\Sigma_{\\textbf{nbI}},\\mathcal{V})\\rightarrow\\{0,1\\}$ is a bivaluation for $\\mathcal{L}\\in\\{\\textbf{nbIciw}, \\textbf{nbIci}, \\textbf{nbIcl}\\}$ if, and only if, it is a $\\Sigma_{\\textbf{nbI}}$-homomorphism from $\\textbf{F}(\\Sigma_{\\textbf{nbI}}, \\mathcal{V})$ to $\\textbf{2}_{\\textbf{nbIciw}}$ which lies in $\\mathcal{F}_{\\textbf{2}_{\\textbf{nbIciw}}}^{\\mathcal{L}}$.\n\\end{theorem}\n\n\\begin{theorem}\nGiven formulas $\\Gamma\\cup\\{\\varphi\\}$ of $\\mathcal{L}\\in\\{\\textbf{nbIciw}, \\textbf{nbIci}, \\textbf{nbIcl}\\}$, $\\Gamma\\vDash_{\\mathcal{L}}\\varphi$ if and only if $\\Gamma\\vDash_{\\mathbb{2}_{\\mathcal{L}}}\\varphi$.\n\\end{theorem}\n\n\nOur finite RNmatrices again lead to decision methods through row-branching, row-eliminating truth tables, and for all three of the logics $\\nbIciw$, $\\nbIci$ and $\\nbIcl$. We believe the subjacent Nmatrices are explicit enough, so let us focus on the conditions for a row to be erased.\n\\begin{enumerate}\n\\item If $\\alpha\\uparrow\\beta$ and $\\beta\\uparrow\\alpha$ have different values.\n\\item If $\\alpha\\uparrow\\neg\\alpha$ or $\\neg\\alpha\\uparrow\\alpha$ is $0$, and either $\\alpha$ or $\\neg\\alpha$ is also $0$.\n\\item \\begin{enumerate}\n\\item If the logic is $\\nbIci$: if $\\neg(\\alpha\\uparrow\\neg\\alpha)$ is $1$ but either $\\alpha$ or $\\neg\\alpha$ is $0$.\n\\item If the logic is $\\nbIcl$, $\\neg(\\alpha\\wedge\\neg\\alpha)$ is $1$ and: either $\\alpha\\uparrow\\neg\\alpha$ or $\\neg\\alpha\\uparrow\\alpha$ is $0$, or $\\alpha$ and $\\neg\\alpha$ are both $1$.\n\\end{enumerate}\n\\end{enumerate}\n\n\\subsection{Another decision method}\n\nWe can add rules to the tableau calculus $\\mathbb{T}_{\\nbI}$ inspired by the tables found in Section \\ref{Decision method for nbIciw, ...} to obtain tableau calculi $\\mathbb{T}_{\\nbIciw}$, $\\mathbb{T}_{\\nbIci}$ and $\\mathbb{T}_{\\nbIcl}$ capable of characterizing their respective logics. So, consider the following tableau rules:\n\n\n$$\n\\begin{array}{cp{1cm}cp{1cm}cp{1cm}c}\n\\displaystyle \\frac{\\textsf{0}(\\varphi\\uparrow\\neg\\varphi)}{\\begin{array}{c}\\textsf{1}(\\varphi) \\\\ \\textsf{1}(\\neg\\varphi)\\end{array}} & &\n\\displaystyle \\frac{\\textsf{0}(\\neg\\varphi\\uparrow\\varphi)}{\\begin{array}{c}\\textsf{1}(\\varphi) \\\\ \\textsf{1}(\\neg\\varphi)\\end{array}} & & \n\\displaystyle \\frac{\\textsf{1}(\\neg(\\varphi\\uparrow\\neg\\varphi))}{\\begin{array}{c}\\textsf{1}(\\varphi) \\\\ \\textsf{1}(\\neg\\varphi)\\end{array}} & & \n\\displaystyle \\frac{\\textsf{1}(\\neg(\\varphi\\wedge\\neg\\varphi))}{\\textsf{0}(\\varphi)\\mid\\textsf{0}(\\neg\\varphi)}\\\\[2mm]\n\\end{array}\n$$\n\nThen, by addition to $\\mathbb{T}_{\\nbI}$ of:\n\\begin{enumerate}\n\\item the two first rules, one obtains $\\mathbb{T}_{\\nbIciw}$;\n\\item the three first rules, one obtains $\\mathbb{T}_{\\nbIci}$;\n\\item the two first rules plus the fourth, one obtains $\\mathbb{T}_{\\nbIcl}$.\n\\end{enumerate}\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\\section{Our logics are not algebraizable by Blok and Pigozzi}\n\n\\subsection{$\\bI$ is not algebraizable by Blok and Pigozzi}\\label{bI is not algebraizable}\n\nWe wish to prove $\\bI$ is not algebraizable even by some quite expressive standards: specifically, we wish to show it is not \\index{Algebraizable}algebraizable according to Blok and Pigozzi \\cite{BlokPigozzi}. To accomplish that we will use \\index{Lewin}Lewin, Mikenberg and Schwarze's construction found in\\cite{Lewin}, and prove as they do that there is a model for $\\bI$ for which the Leibniz operator $\\Omega_{\\bI}$, that sends a $\\bI$-filter to its largest compatible congruence, is not a bijection; from Theorem $5.1$ of \\cite{BlokPigozzi}, this proves $\\bI$ is not algebraizable.\n\n\\begin{definition}\nGiven a signature $\\Sigma$ and a $\\Sigma$-algebra $\\mathcal{A}=(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma})$, a \\index{Congruence}congruence in $\\mathcal{A}$ is a relation $\\theta$ on $A\\times A$ such that, if $\\sigma\\in\\Sigma_{n}$ and $a_{1}, b_{1}, \\dotsc , a_{n}, b_{n}\\in A$ are elements such that $a_{1}\\theta b_{1}, \\dotsc , a_{n}\\theta b_{n}$, then\n\\[\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})\\theta\\sigma_{\\mathcal{A}}(b_{1}, \\dotsc , b_{n}).\\]\n\\end{definition}\n\n\\begin{definition}\nGiven a signature $\\Sigma$, a logic $\\mathcal{L}$ and a $\\Sigma$-algebra $\\mathcal{A}=(A, \\{\\sigma_{\\mathcal{A}}\\}_{\\sigma\\in\\Sigma})$, an \\index{Filter, $\\mathcal{L}$-}$\\mathcal{L}$-filter in $\\mathcal{A}$ is a subset $F\\subseteq A$ such that \n\\[\\Gamma\\vdash_{\\mathcal{L}}\\varphi\\quad\\text{implies}\\quad \\Gamma\\vDash_{(\\mathcal{A}, F)}\\varphi,\\]\nfor every set of formulas $\\Gamma\\cup\\{\\varphi\\}$ over $\\Sigma$.\n\\end{definition}\n\nThe \\index{Congruence, Largest compatible}largest compatible congruence $\\theta$ to a filter $F$ is the largest congruence such that, if $a\\theta b$ and $a\\in F$, then $b\\in F$.\n\nSo, over the signature $\\Sigma_{\\bI}$, for which $(\\Sigma_{\\bI})_{2}=\\{\\vee, \\wedge, \\rightarrow, \\uparrow\\}$ and $(\\Sigma_{\\bI})_{n}=\\emptyset$ for every $n\\neq2$, we consider the $\\Sigma_{\\bI}$-algebra $\\mathfrak{L}$ with universe $L=\\{u, 1, a, b, 0\\}$ and operations given by the tables bellow (all, with the exception of incompatibility, extracted from \\cite{Lewin}).\n\n\\begin{figure}[H]\n\\centering\n\\begin{minipage}[t]{5cm}\n\\centering\n\\begin{tabular}{l|ccccr}\n$\\vee$ & $u$ & $1$ & $a$ & $b$ & $0$ \\\\\\hline\n$u$ & $u$ & $u$ & $u$ & $u$ & $u$\\\\\n$1$ & $u$ & $1$ & $1$ & $1$ & $1$\\\\\n$a$ & $u$ & $1$ & $a$ & $1$ & $a$\\\\\n$b$ & $u$ & $1$ & $1$ & $b$ & $b$\\\\\n$0$ & $u$ & $1$ & $a$ & $b$ & $0$\n\\end{tabular}\n\\caption*{Disjunction}\n\\end{minipage}\n\\centering\n\\begin{minipage}[t]{5cm}\n\\centering\n\\begin{tabular}{l|ccccr}\n$\\wedge$ & $u$ & $1$ & $a$ & $b$ & $0$ \\\\\\hline\n$u$ & $u$ & $1$ & $a$ & $b$ & $0$\\\\\n$1$ & $1$ & $1$ & $a$ & $b$ & $0$\\\\\n$a$ & $a$ & $a$ & $a$ & $0$ & $0$\\\\\n$b$ & $b$ & $b$ & $0$ & $b$ & $0$\\\\\n$0$ & $0$ & $0$ & $0$ & $0$ & $0$\n\\end{tabular}\n\\caption*{Conjunction}\n\\end{minipage}\n\\end{figure}\n\n\n\\begin{figure}[H]\n\\centering\n\\begin{minipage}[t]{5cm}\n\\centering\n\\begin{tabular}{l|ccccr}\n$\\rightarrow$ & $u$ & $1$ & $a$ & $b$ & $0$ \\\\\\hline\n$u$ & $u$ & $u$ & $a$ & $b$ & $0$\\\\\n$1$ & $u$ & $1$ & $a$ & $b$ & $0$\\\\\n$a$ & $u$ & $1$ & $1$ & $b$ & $b$\\\\\n$b$ & $u$ & $1$ & $a$ & $1$ & $a$\\\\\n$0$ & $u$ & $1$ & $1$ & $1$ & $1$\n\\end{tabular}\n\\caption*{Implication}\n\\end{minipage}\n\\begin{minipage}[t]{5cm}\n\\centering\n\\begin{tabular}{l|ccccr}\n$\\uparrow$ & $u$ & $1$ & $a$ & $b$ & $0$ \\\\\\hline\n$u$ & $0$ & $0$ & $0$ & $0$ & $1$\\\\\n$1$ & $0$ & $0$ & $b$ & $a$ & $1$\\\\\n$a$ & $0$ & $b$ & $b$ & $1$ & $1$\\\\\n$b$ & $0$ & $a$ & $1$ & $a$ & $1$\\\\\n$0$ & $1$ & $1$ & $1$ & $1$ & $1$\n\\end{tabular}\n\\caption*{Incompatibility}\n\\end{minipage}\n\\end{figure}\n\n\\begin{figure}[H]\n\\centering\n\\begin{tikzcd}\n & u & \\\\\n & 1 \\arrow[u] & \\\\\n a \\arrow[ru] & & b \\arrow[lu]\\\\\n & 0 \\arrow[ru]\\arrow[lu] & \\\\\n \\end{tikzcd}\n\\caption*{The lattice $(L, \\vee, \\wedge)$}\n\\end{figure}\n\nWe then consider the logical matrix $\\mathfrak{M}=(\\mathfrak{L}, D)$, with $D=\\{u,1\\}$. \n\n\\subsubsection{$\\mathfrak{M}$ is a model of $\\bI$}\n\nNotice that $\\textbf{Ax\\: 1}$, $\\textbf{Ax\\: 3}$, $\\textbf{Ax\\: 4}$, $\\textbf{Ax\\: 5}$, $\\textbf{Ax\\: 6}$, $\\textbf{Ax\\: 7}$, $\\textbf{Ax\\: 8}$ and Modus Ponens of the definition of $\\bI$ ate the beginning of Section \\ref{Defining bI} correspond, respectively, to the axiom schemata and rules $1$, $6$, $4$, $5$, $7$, $8$, $9$ and $10$ of $\\textbf{C}_{1}$ as defined in \\cite{Lewin}; since the operations $\\vee$, $\\wedge$ and $\\rightarrow$ in $\\mathfrak{L}$ are exactly the same as the ones in the algebra of Lewin, Mikenberg and Schwarze's article, and since the logical matrix formed by that algebra and $D$ models $\\textbf{C}_{1}$, we obtain that $\\mathfrak{M}$ models at least these axiom schemata and rules of inference. It remains to be shown it also models $\\textbf{Ax\\: 2}$, $\\textbf{Ax\\: 9}^{*}$, $\\textbf{Ip}$ and $\\textbf{Comm}$.\n\n\\begin{enumerate}\n\\item Concerning $\\textbf{Ax\\: 2}$ of $\\bI$, $(\\alpha\\rightarrow(\\beta\\rightarrow \\gamma)\\big)\\rightarrow\\big((\\alpha\\rightarrow\\beta)\\rightarrow(\\alpha\\rightarrow\\gamma))$, notice that by two applications of the deduction meta-theorem we have that the validity of the axiom schema $2$ of \\cite{Lewin} in $\\bI$ is equivalent to stating that $\\alpha\\rightarrow \\beta, \\alpha\\rightarrow(\\beta\\rightarrow \\gamma)\\vdash_{\\bI}\\alpha\\rightarrow\\gamma$, which by two new applications of the deduction meta-theorem is now equivalent to the validity of $\\textbf{Ax\\: 2}$. Since the matrix of Lewin, Mikenberg and Schwarze's algebra validates the axiom schema $2$, we have $\\mathfrak{M}$ validates $\\textbf{Ax\\: 2}$.\n\n\\item For $\\textbf{Ax\\: 9}^{*}$, $(\\alpha\\rightarrow\\beta)\\vee\\alpha$, we see from the tables below that the image of $\\textbf{Ax\\: 9}^{*}$ under any homomorphism is in $D$, and therefore this axiom schema is validated by $\\mathfrak{M}$.\n\n\\begin{figure}[H]\n\\centering\n\\begin{tabular}{l|ccccc|ccccr}\n& \\multicolumn{5}{c|}{$x\\rightarrow y$} & \\multicolumn{5}{c}{$(x\\rightarrow y)\\vee x$}\\\\ \\hline\n\\diag{.1em}{0.2cm}{$x$}{$y$} & $u$ & $1$ & $a$ & $b$ & $0$ & $u$ & $1$ & $a$ & $b$ & $0$ \\\\\\hline\n$u$ & $u$ & $u$ & $a$ & $b$ & $0$ & $u$ & $u$ & $u$ & $u$ & $u$\\\\\n$1$ & $u$ & $1$ & $a$ & $b$ & $0$ & $u$ & $1$ & $1$ & $1$ & $1$\\\\\n$a$ & $u$ & $1$ & $1$ & $b$ & $b$ & $u$ & $1$ & $1$ & $1$ & $1$\\\\\n$b$ & $u$ & $1$ & $a$ & $1$ & $a$ & $u$ & $1$ & $1$ & $1$ & $1$\\\\\n$0$ & $u$ & $1$ & $1$ & $1$ & $1$ & $u$ & $1$ & $1$ & $1$ & $1$\n\\end{tabular}\n\\caption*{Table for $\\textbf{Ax\\: 9}^{*}$}\n\\end{figure}\n\n\n\n\\item To prove $\\mathfrak{M}$ models $\\textbf{Ip}$, given by $(\\alpha\\uparrow\\beta)\\rightarrow(\\alpha\\rightarrow(\\beta\\rightarrow\\gamma))$, we begin with the two tables below.\n\n\\begin{figure}[H]\n\\centering\n\\begin{minipage}[t]{5cm}\n\\centering\n\\begin{tabular}{l|ccccr}\n\\diag{.1em}{0.2cm}{$y$}{$z$} & $u$ & $1$ & $a$ & $b$ & $0$ \\\\\\hline\n$u$ & $u$ & $u$ & $a$ & $b$ & $0$\\\\\n$1$ & $u$ & $1$ & $a$ & $b$ & $0$\\\\\n$a$ & $u$ & $1$ & $1$ & $b$ & $b$\\\\\n$b$ & $u$ & $1$ & $a$ & $1$ & $a$\\\\\n$0$ & $u$ & $1$ & $1$ & $1$ & $1$\n\\end{tabular}\n\\caption*{Table for $y\\rightarrow z$}\n\\end{minipage}\n\\centering\n\\begin{minipage}[t]{5cm}\n\\centering\n\\begin{tabular}{l|ccccr}\n\\diag{.1em}{0.2cm}{$x$}{$y$} & $u$ & $1$ & $a$ & $b$ & $0$ \\\\\\hline\n$u$ & $0$ & $0$ & $0$ & $0$ & $1$\\\\\n$1$ & $0$ & $0$ & $b$ & $a$ & $1$\\\\\n$a$ & $0$ & $b$ & $b$ & $1$ & $1$\\\\\n$b$ & $0$ & $a$ & $1$ & $a$ & $1$\\\\\n$0$ & $1$ & $1$ & $1$ & $1$ & $1$\n\\end{tabular}\n\\caption*{Table for $x\\uparrow y$}\n\\end{minipage}\n\\end{figure}\n\nBy looking at the table for $\\textbf{Ip}$ below, we see that, since the image of $\\textbf{Ip}$ under any homomorphism is always in $D$, we have $\\mathfrak{M}$ validates this axiom schema.\n\n\\begin{figure}[H]\n\\centering\n\\begin{tabular}{l|c|c|ccccc|C{1.2em}C{1.2em}C{1.2em}C{1.2em}C{1.2em}}\n& & $x\\uparrow y$ & \\multicolumn{5}{c|}{$x\\rightarrow(y\\rightarrow z)$} & \\multicolumn{5}{c}{$(x\\uparrow y)\\rightarrow(x\\rightarrow(y\\rightarrow z))$}\\\\ \\hline\n$x$ & \\diag{.1em}{0.2cm}{$y$}{$z$} & & $u$ & $1$ & $a$ & $b$ & $0$ & $u$ & $1$ & $a$ & $b$ & $0$ \\\\ \\hline\n\\multirow{5}{*}{$u$} & $u$ & $0$ & $u$ & $u$ & $a$ & $b$ & $0$ & $u$ & $u$ & $1$ & $1$ & $1$\\\\\n& $1$ & $0$ & $u$ & $u$ & $a$ & $b$ & $0$ & $u$ & $u$ & $1$ & $1$ & $1$\\\\\n& $a$ & $0$ & $u$ & $u$ & $u$ & $b$ & $b$ & $u$ & $u$ & $u$ & $1$ & $1$\\\\\n& $b$ & $0$ & $u$ & $u$ & $a$ & $u$ & $a$ & $u$ & $u$ & $1$ & $u$ & $1$\\\\\n& $0$ & $1$ & $u$ & $u$ & $u$ & $u$ & $u$ & $u$ & $u$ & $u$ & $u$ & $u$\\\\ \\hline\n\\multirow{5}{*}{$1$} & $u$ & $0$ & $u$ & $u$ & $a$ & $b$ & $0$ & $u$ & $u$ & $1$ & $1$ & $1$\\\\\n& $1$ & $0$ & $u$ & $1$ & $a$ & $b$ & $0$ & $u$ & $1$ & $1$ & $1$ & $1$\\\\\n& $a$ & $b$ & $u$ & $1$ & $1$ & $b$ & $b$ & $u$ & $1$ & $1$ & $1$ & $1$\\\\\n& $b$ & $a$ & $u$ & $1$ & $a$ & $1$ & $a$ & $u$ & $1$ & $1$ & $1$ & $1$\\\\\n& $0$ & $1$ & $u$ & $1$ & $1$ & $1$ & $1$ & $u$ & $1$ & $1$ & $1$ & $1$\\\\ \\hline\n\\multirow{5}{*}{$a$} & $u$ & $0$ & $u$ & $u$ & $1$ & $b$ & $b$ & $u$ & $u$ & $1$ & $1$ & $1$\\\\\n& $1$ & $b$ & $u$ & $1$ & $1$ & $b$ & $b$ & $u$ & $1$ & $1$ & $1$ & $1$\\\\\n& $a$ & $b$ & $u$ & $1$ & $1$ & $b$ & $b$ & $u$ & $1$ & $1$ & $1$ & $1$\\\\\n& $b$ & $1$ & $u$ & $1$ & $1$ & $1$ & $1$ & $u$ & $1$ & $1$ & $1$ & $1$\\\\\n& $0$ & $1$ & $u$ & $1$ & $1$ & $1$ & $1$ & $u$ & $1$ & $1$ & $1$ & $1$\\\\ \\hline\n\\multirow{5}{*}{$b$} & $u$ & $0$ & $u$ & $u$ & $a$ & $1$ & $a$ & $u$ & $u$ & $1$ & $1$ & $1$\\\\\n& $1$ & $a$ & $u$ & $1$ & $a$ & $1$ & $a$ & $u$ & $1$ & $1$ & $1$ & $1$\\\\\n& $a$ & $1$ & $u$ & $1$ & $1$ & $1$ & $1$ & $u$ & $1$ & $1$ & $1$ & $1$\\\\\n& $b$ & $a$ & $u$ & $1$ & $a$ & $1$ & $a$ & $u$ & $1$ & $1$ & $1$ & $1$\\\\\n& $0$ & $1$ & $u$ & $1$ & $1$ & $1$ & $1$ & $u$ & $1$ & $1$ & $1$ & $1$\\\\ \\hline\n\\multirow{5}{*}{$0$} & $u$ & $1$ & $u$ & $u$ & $1$ & $1$ & $1$ & $u$ & $u$ & $1$ & $1$ & $1$\\\\\n& $1$ & $1$ & $u$ & $1$ & $1$ & $1$ & $1$ & $u$ & $1$ & $1$ & $1$ & $1$\\\\\n& $a$ & $1$ & $u$ & $1$ & $1$ & $1$ & $1$ & $u$ & $1$ & $1$ & $1$ & $1$\\\\\n& $b$ & $1$ & $u$ & $1$ & $1$ & $1$ & $1$ & $u$ & $1$ & $1$ & $1$ & $1$\\\\\n& $0$ & $1$ & $u$ & $1$ & $1$ & $1$ & $1$ & $u$ & $1$ & $1$ & $1$ & $1$\\\\\n\\end{tabular}\n\\caption*{Table for $\\textbf{Ip}$}\n\\end{figure}\n\n\\item It only remains to show $\\mathfrak{M}$ validates $\\textbf{Comm}$, an axiom schema given by $(\\alpha\\uparrow\\beta)\\rightarrow(\\beta\\uparrow\\alpha)$, what is done on the following table; notice that the tables for $x\\uparrow y$ and $y\\uparrow x$ are the same since incompatibility is commutative in $\\mathfrak{M}$.\n\n\\begin{figure}[H]\n\\centering\n\\begin{tabular}{l|ccccc|ccccr}\n& \\multicolumn{5}{c|}{$x\\uparrow y=y\\uparrow x$} & \\multicolumn{5}{c}{$(x\\uparrow y)\\rightarrow(y\\uparrow x)$}\\\\\\hline\n\\diag{.1em}{0.2cm}{$x$}{$y$} & $u$ & $1$ & $a$ & $b$ & $0$ & $u$ & $1$ & $a$ & $b$ & $0$ \\\\\\hline\n$u$ & $0$ & $0$ & $0$ & $0$ & $1$ & $1$ & $1$ & $1$ & $1$ & $1$\\\\\n$1$ & $0$ & $0$ & $b$ & $a$ & $1$ & $1$ & $1$ & $1$ & $1$ & $1$\\\\\n$a$ & $0$ & $b$ & $b$ & $1$ & $1$ & $1$ & $1$ & $1$ & $1$ & $1$\\\\\n$b$ & $0$ & $a$ & $1$ & $a$ & $1$ & $1$ & $1$ & $1$ & $1$ & $1$\\\\\n$0$ & $1$ & $1$ & $1$ & $1$ & $1$ & $1$ & $1$ & $1$ & $1$ & $1$\n\\end{tabular}\n\\caption*{Table for $\\textbf{Comm}$}\n\\end{figure}\n\n\\end{enumerate}\n\nWith this, we have proved that $\\mathfrak{M}$ is a model for $\\bI$, meaning that, for any formulas $\\Gamma\\cup\\{\\varphi\\}$ on the signature $\\Sigma_{\\bI}$, $\\Gamma\\vdash_{\\bI}\\varphi$ implies $\\Gamma\\vDash_{\\mathfrak{M}}\\varphi$.\n\n\\subsubsection{There are no non-trivial congruences on $\\mathfrak{L}$}\n\nNow, we wish to prove that the $\\Sigma_{\\bI}$-algebra $\\mathfrak{L}$ has only two congruences, the ones we call trivial: \\label{nabla}$\\nabla$, equal to $L\\times L$, and \\label{Delta}$\\Delta$, given by $\\{(x,x)\\ :\\ x\\in L\\}$. Here, we will perpetrate an abuse of notation: take a signature $\\Sigma$, a $\\Sigma$-algebra $\\mathcal{A}$ with universe $A$, a congruence $\\theta$ on $\\mathcal{A}$, $\\sigma\\in\\Sigma_{n}$ and $\\omega\\in\\Sigma_{m}$, and $a_{1}, \\dotsc , a_{n}, b_{1}, \\dotsc , b_{m}\\in A$: then, if $\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})=a$ and $\\omega_{\\mathcal{A}}(b_{1}, \\dotsc , b_{m})=b$, and $\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})\\theta\\omega_{\\mathcal{A}}(b_{1}, \\dotsc , b_{m})$, we may simply write $a=\\sigma_{\\mathcal{A}}(a_{1}, \\dotsc , a_{n})\\theta\\omega_{\\mathcal{A}}(b_{1}, \\dotsc , b_{m})=b$ to make it clear that $a\\theta b$.\n\n\\begin{enumerate}\n\\item \n\\begin{enumerate}\n\\item If $u\\theta 1$, $0=(u\\uparrow a)\\theta(1\\uparrow a)=b$, $0=(u\\uparrow b)\\theta(1\\uparrow b)=a$ and $1=(a\\rightarrow a)\\theta (b\\rightarrow a)=a$, and therefore $\\theta=\\nabla$.\n\\item If $u\\theta a$, then $u=(u\\vee 1)\\theta (a\\vee 1)=1$ (meaning $a\\theta 1$), $b=(u\\wedge b)\\theta (a\\wedge b)=0$ and $0=(u\\uparrow b)\\theta(1\\uparrow b)=a$, implying $\\theta=\\nabla$.\n\\item If $u\\theta b$, then $u=(u\\vee 1)\\theta(b\\vee 1)=1$, $a=(u\\wedge a)\\theta(b\\wedge a)=0$ and $0=(u\\uparrow a)\\theta(1\\uparrow a)=b$, and again $\\theta=\\nabla$.\n\\item If $u\\theta 0$, then for every $x$ we have $u=(u\\vee x)\\theta (0\\vee x)=x$, meaning $\\theta=\\nabla$.\n\\end{enumerate}\nWith this, we have proved that, if any pair $(u,x)$, with $x\\in L\\setminus\\{u\\}$, is in $\\theta$, then $\\theta=\\nabla$.\n\\item We will ignore the case $1\\theta u$, since it is equivalent to $u\\theta 1$, a case considered before.\n\\begin{enumerate}\n\\item If $1\\theta a$, $b=(1\\wedge b)\\theta (a\\wedge b)=0$, $u=(u\\rightarrow 1)\\theta(u\\rightarrow a)=a$ (and therefore $u\\theta 1$) and $0=(u\\uparrow b)\\theta(1\\uparrow b)=a$, meaning $\\theta=\\nabla$.\n\\item If $1\\theta b$, $a=(1\\wedge a)\\theta(b\\wedge a)=0$, $u=(u\\rightarrow1)\\theta(u\\rightarrow b)=b$ (and therefore $u\\theta 1$) and $0=(u\\uparrow a)\\theta(1\\uparrow a)=b$, and so $\\theta=\\nabla$.\n\\item If $1\\theta 0$, $a=(1\\rightarrow a)\\theta(0\\rightarrow a)=1$, $b=(1\\rightarrow b)\\theta(0\\rightarrow b)=1$ and $u=(u\\rightarrow 1)\\theta (u\\rightarrow 0)=0$, and therefore $\\theta=\\nabla$.\n\\end{enumerate}\nOnce again, we have that if any pair $(1,x)$, with $x\\in L\\setminus\\{1\\}$, is in $\\theta$, then $\\theta=\\nabla$.\n\\item The cases $a\\theta u$ and $a\\theta 1$ correspond to the cases $u\\theta a$ and $1\\theta a$ above.\n\\begin{enumerate}\n\\item If $a\\theta 0$, $1=(a\\vee b)\\theta(0\\vee b)=b$, $u=(u\\rightarrow1)\\theta(u\\rightarrow b)=b$ (and therefore $u\\theta 1$) and $0=(u\\uparrow a)\\theta(1\\uparrow a)=b$, what implies $\\theta=\\nabla$.\n\\item If $a\\theta b$, $b=(a\\rightarrow b)\\theta (b\\rightarrow b)=1$ (implying $a\\theta1$), $0=(a\\wedge b)\\theta(b\\wedge b)=b$ and $u=(u\\rightarrow 1)\\theta(u\\rightarrow a)=a$, what means $\\theta=\\nabla$.\n\\end{enumerate}\nSo, if $(a,x)\\in\\theta$, for $x\\in L\\setminus\\{a\\}$, $\\theta=\\nabla$.\n\\item The cases $b\\theta u$, $b\\theta 1$ and $b\\theta a$ equal the previous cases, respectively, $u\\theta b$, $1\\theta b$ and $a\\theta b$. \n\\begin{enumerate}\n\\item If $b\\theta 0$, $1=(a\\vee b)\\theta(a\\vee 0)=a$, $u=(u\\rightarrow 1)\\theta(u\\rightarrow a)=a$ (and therefore $u\\theta 1$) and $0=(u\\uparrow b)\\theta(1\\uparrow b)=a$, and therefore $\\theta=\\nabla$.\n\\end{enumerate}\nAgain, if $(b,x)\\in \\theta$, for $b\\in L\\setminus\\{b\\}$, then $\\theta=\\nabla$.\n\\item Since the cases $0\\theta u$, $0\\theta1$, $0\\theta a$ and $0\\theta b$ correspond, respectively, to $u\\theta 0$, $1\\theta0$, $a\\theta0$ and $b\\theta0$, we have that, if $(0,x)\\in\\theta$, for $x\\in L\\setminus\\{0\\}$, then $\\theta=\\nabla$.\n\\end{enumerate}\n\nWith all of this, we discover that for any congruence $\\theta$ in $\\mathfrak{L}$, if $(x,y)\\in \\theta$, for $x\\neq y$, then $\\theta=\\nabla$, implying as we had mentioned that the only two congruences in $\\mathfrak{L}$ are $\\nabla$ and $\\Delta$.\n\n\\subsubsection{Two filters whose largest compatible congruence is $\\nabla$}\n\n\\begin{lemma}\\label{Classifying filters}\nGiven a signature $\\Sigma$, a $\\Sigma$-algebra $\\mathcal{A}$ and a logic $\\mathcal{L}$ over $\\Sigma$, $F$ is a $\\mathcal{L}$-filter in $\\mathcal{A}$ if, and only if, the following are satisfied:\n\\begin{enumerate}\n\\item for every $\\Sigma$-homomorphism $\\sigma:\\textbf{F}(\\Sigma, \\mathcal{V})\\rightarrow\\mathcal{A}$ and instance of axiom $\\psi$ of $\\mathcal{L}$, $\\sigma(\\psi)\\in F$;\n\\item for every $\\Sigma$-homomorphism $\\sigma:\\textbf{F}(\\Sigma, \\mathcal{V})\\rightarrow\\mathcal{A}$ and instance of a rule of inference\\\\ $\\psi_{1}, \\dotsc , \\psi_{n}|\\psi$ of $\\mathcal{L}$, if $\\sigma(\\psi_{1}), \\dotsc , \\sigma(\\psi_{n})\\in F$ then $\\sigma(\\psi)\\in F$.\n\\end{enumerate}\n\\end{lemma}\n\n\\begin{proof}\nSuppose $F$ is a $\\mathcal{L}$-filter in $\\mathcal{A}$ and $\\sigma:\\textbf{F}(\\Sigma, \\mathcal{V})\\rightarrow\\mathcal{A}$ is a $\\Sigma$-homomorphism.\n\\begin{enumerate}\n\\item If $\\psi$ is an instance of an axiom, $\\vdash_{\\mathcal{L}}\\psi$, meaning $\\vDash_{(\\mathcal{A}, F)}\\psi$ and, therefore, $\\sigma(\\psi)\\in F$.\n\\item If $\\psi_{1}, \\dotsc , \\psi_{n}|\\psi$ is an instance of an inference rule, $\\psi_{1}, \\dotsc , \\psi_{n}\\vdash_{\\mathcal{L}}\\psi$, meaning $\\psi_{1}, \\dotsc , \\psi_{n}$\\\\$\\vDash_{(\\mathcal{A}, F)}\\psi$ and, therefore, if $\\sigma(\\psi_{1}), \\dotsc , \\sigma(\\psi_{n})\\in F$, then $\\sigma(\\psi)\\in F$.\n\\end{enumerate}\n\nReciprocally, suppose $\\Gamma\\vdash_{\\mathcal{L}}\\varphi$ and that $\\alpha_{1}, \\dotsc , \\alpha_{n}$ is a proof of $\\varphi$ from $\\Gamma$ with $\\alpha_{n}=\\varphi$; we shall prove that, for any $\\Sigma$-homomorphism $\\sigma:\\textbf{F}(\\Sigma, \\mathcal{V})\\rightarrow\\mathcal{A}$ such that $\\sigma(\\Gamma)\\subseteq F$, $\\sigma(\\alpha_{1}), \\dotsc , \\sigma(\\alpha_{n})\\in F$, and therefore $\\sigma(\\varphi)\\in F$ and $\\Gamma\\vDash_{(\\mathcal{A}, F)}\\varphi$.\n\nFor induction hypothesis, being $\\alpha_{1}$ either an element of $\\Gamma$ or an instance of an axiom, assume $\\alpha_{1}$ through $\\alpha_{i-1}$ are mapped, by $\\sigma$, into $F$.\n\\begin{enumerate}\n\\item If $\\alpha_{i}$ is in $\\Gamma$, by hypothesis on $\\sigma$ we have $\\sigma(\\alpha_{i})\\in F$.\n\\item If $\\alpha_{i}$ is an instance of an axiom, by our hypothesis on $F$ we have $\\sigma(\\alpha_{i})\\in F$.\n\\item Finally, if there are $\\alpha_{i_{1}}, \\dotsc , \\alpha_{i_{n}}$, with $i_{1}, \\dotsc , i_{n}\\in\\{1, \\dotsc , i-1\\}$, such that $\\alpha_{i_{1}}, \\dotsc , \\alpha_{i_{n}}|\\alpha_{i}$ is an instance of a rule of inference, by induction hypothesis $\\sigma(\\alpha_{i_{1}}), \\dotsc , \\sigma(\\alpha_{i_{n}})\\in F$, and again by our hypothesis on $F$ we have $\\sigma(\\alpha_{i})\\in F$.\n\\end{enumerate}\n\\end{proof}\n\nSo, we take the subsets $F_{a}=\\{u, 1, a\\}$ and $F_{b}=\\{u, 1, b\\}$ of $L$, and state that both are $\\bI$-filters. First of all, for any $\\Sigma_{\\bI}$-homomorphism $\\sigma:\\textbf{F}(\\Sigma_{\\bI}, \\mathcal{V})\\rightarrow\\mathfrak{L}$ and any instance of an axiom schema $\\psi$ of $\\bI$, since $\\mathfrak{M}=(\\mathfrak{L}, D)$ models $\\bI$, $\\sigma(\\psi)\\in D=\\{u,1\\}\\subseteq F_{a}$ and in much the same way $\\sigma(\\psi)\\in F_{b}$, implying that both $F_{a}$ and $F_{b}$ satisfy the first condition of Lemma \\ref{Classifying filters} for being an $\\bI$-filter.\n\nFurthermore, there is only one rule of inference to analyze, that of Modus Ponens. From the table for implication, we see that if $x\\rightarrow y$ is in $F_{a}$, then either $y\\in F_{a}$ or $x\\in L\\setminus F_{a}$; so, if both $x$ and $x\\rightarrow y$ are in $F_{a}$, then $y\\in F_{a}$. Similarly, if both $x$ and $x\\rightarrow y$ are in $F_{b}$, $y$ is also in $F_{b}$, what implies that both $F_{a}$ and $F_{b}$ are $\\bI$-filters by Lemma \\ref{Classifying filters}.\n\nBut we state that the largest compatible congruence to both $F_{a}$ and $F_{b}$ is $\\Delta$, meaning the Leibniz operator $\\Omega_{\\bI}$ of $\\bI$ is not injective and therefore $\\bI$ is not algebraizable by Blok and Pigozzi. This is easy to prove: $\\nabla$ is not compatible to neither $F_{a}$ nor $F_{b}$, since $u\\Delta 0$ and $u\\in F_{a}\\cap F_{b}$, but $0$ is not in $F_{a}$ nor in $F_{b}$; clearly $\\Delta$ is compatible to both $F_{a}$ and $F_{b}$, and since there are no congruences larger than $\\Delta$ different from $\\nabla$, we obtain the aforementioned result.\n\n\\subsection{$\\nbI$ is not algebraizable by Blok and Pigozzi}\n\nNow, we will follow the reasoning found in Section \\ref{bI is not algebraizable} to show that $\\nbI$ is also not algebraizable by Blok And Pigozzi, showing that its Leibniz operator is not bijective, the only difference being that we must add a negation to $\\mathfrak{L}$. So, consider the $\\Sigma_{\\nbI}$-algebra $\\mathfrak{L}_{\\nbI}$ with universe $L=\\{u,1,a,b,0\\}$, $\\sigma_{\\mathfrak{L}_{\\nbI}}$ equal to $\\sigma_{\\mathfrak{L}}$ for any $\\sigma\\in\\{\\vee, \\wedge, \\rightarrow, \\uparrow\\}$ and negation defined by the table below; we drop $\\mathfrak{L}_{\\nbI}$ from indexing the operations for simplicity.\n\n\\begin{figure}[H]\n\\centering\n\\begin{tabular}{l|ccccr}\n$x$ & $u$ & $1$ & $a$ & $b$ & $0$\\\\ \\hline\n$\\neg x$ & $1$ & $0$ & $b$ & $a$ & $1$\n\\end{tabular}\n\\caption*{Table for Negation}\n\\end{figure}\n\nWe then define the logical matrix $\\mathfrak{M}_{\\nbI}=(\\mathfrak{L}_{\\nbI}, D)$, with $D=\\{u,1\\}$. Since the operations associated to the symbols in $\\{\\vee, \\wedge, \\rightarrow, \\uparrow\\}$ are the same as those of $\\mathfrak{L}$, we see that $\\mathfrak{M}_{\\nbI}$ models the axiom schemata $\\textbf{Ax\\: 1}$ through $\\textbf{Ax\\: 8}$ of $\\bI$, plus $\\textbf{Ax\\: 9}^{*}$, $\\textbf{Ip}$ and $\\textbf{Comm}$, since those axiom schemata involve only the connectives on $\\Sigma_{\\bI}$.\n\nThe proof that $\\mathfrak{M}_{\\nbI}$ models $\\textbf{Ax\\: 11}^{*}$ is not necessary since it corresponds to the axiom scheme number $3$ for $\\textbf{C}_{1}$ in the axiomatization found in \\cite{Lewin}, and we use the same negation.\n\nThe proof that there are only two congruences on $\\mathfrak{L}_{\\nbI}$ goes as expected, beginning with the supposition that a pair $(x,y)$, with $x\\neq y$, is in $\\theta$, what then implies $\\theta=\\nabla$.\n\nFinally, we trivially find both $F_{a}=\\{u,1,a\\}$ and $F_{b}=\\{u,1,b\\}$ are $\\nbI$-filters whose largest compatible congruence is $\\Delta$, what proves $\\nbI$ is not-algebraizable according to Blok and Pigozzi.\n\n\\subsection{$\\nbIciw$, $\\nbIci$ and $\\nbIcl$ are not algebraizable by Blok and Pigozzi}\n\nWe state that $\\mathfrak{M}_{\\nbI}$ also models $\\nbIciw$, $\\nbIci$ and $\\nbIcl$, and since $F_{a}$ and $F_{b}$ are still, respectively, $\\mathcal{L}$-filters, for $\\mathcal{L}\\in\\{\\nbIciw, \\nbIci, \\nbIcl\\}$, we prove that none of these three is algebraizable by Blok and Pigozzi. The proof that $\\mathfrak{M}$ models $\\textbf{ciw}^{*}$, $\\textbf{ci}^{*}$ and $\\textbf{cl}^{*}$ is in the following tables.\n\n\\begin{figure}[H]\n\\centering\n\\begin{tabular}{l|cccc}\n$x$ & $\\neg x$ & $x\\wedge\\neg x$ & $x\\uparrow\\neg x$ & $(x\\uparrow\\neg x)\\vee(x\\wedge\\neg x)$ \\\\\\hline\n$u$ & $1$ & $1$ & $0$ & $1$ \\\\\n$1$ & $0$ & $0$ & $1$ & $1$ \\\\\n$a$ & $b$ & $0$ & $1$ & $1$ \\\\\n$b$ & $a$ & $0$ & $1$ & $1$ \\\\\n$0$ & $1$ & $0$ & $1$ & $1$ \n\\end{tabular}\n\\caption*{Table for $\\textbf{ciw}^{*}$}\n\\end{figure}\n\n\n\\begin{figure}[H]\n\\centering\n\\begin{tabular}{l|ccccc}\n$x$ & $\\neg x$ & $x\\wedge\\neg x$ & $x\\uparrow\\neg x$ & $\\neg(x\\uparrow\\neg x)$ & $\\neg(x\\uparrow\\neg x)\\rightarrow(x\\wedge\\neg x)$ \\\\\\hline\n$u$ & $1$ & $1$ & $0$ & $1$ & $1$ \\\\\n$1$ & $0$ & $0$ & $1$ & $0$ & $1$ \\\\\n$a$ & $b$ & $0$ & $1$ & $0$ & $1$\\\\\n$b$ & $a$ & $0$ & $1$ & $0$ & $1$\\\\\n$0$ & $1$ & $0$ & $1$ & $0$ & $1$\n\\end{tabular}\n\\caption*{Table for $\\textbf{ci}^{*}$}\n\\end{figure}\n\n\\begin{figure}[H]\n\\centering\n\\begin{tabular}{l|ccccc}\n$x$ & $\\neg x$ & $x\\wedge\\neg x$ & $x\\uparrow\\neg x$ & $\\neg(x\\wedge\\neg x)$ & $\\neg(x\\wedge\\neg x)\\rightarrow(x\\uparrow\\neg x)$ \\\\\\hline\n$u$ & $1$ & $1$ & $0$ & $0$ & $1$ \\\\\n$1$ & $0$ & $0$ & $1$ & $1$ & $1$ \\\\\n$a$ & $b$ & $0$ & $1$ & $1$ & $1$\\\\\n$b$ & $a$ & $0$ & $1$ & $1$ & $1$\\\\\n$0$ & $1$ & $0$ & $1$ & $1$ & $1$\n\\end{tabular}\n\\caption*{Table for $\\textbf{cl}^{*}$}\n\\end{figure}\n\n\n\n\n\n\\section{$\\bI$ and $\\nbI$ are not characterizable by finite Nmatrices}\\label{bI is not characterized by Nmatrices section}\n\nIt is well known that da Costa's hierarchy is both not algebraizable according to Blok and Pigozzi (\\cite{Lewin}, \\cite{Mortensen80}) and not characterizable by a finite Nmatrix (\\cite{Avron}). We have already shown that $\\bI$, and the other systems of incompatibility we have here defined, are not algebraizable according to Blok and Pigozzi, giving them a difficulty to be approached close to that of the logics $C_{n}$; but here, we show furthermore that $\\bI$ and $\\nbI$ are not characterizable by finite Nmatrices, and are, therefore, specially hard systems from a semantical standpoint. Interestingly, most proofs found so far in this chapter, such as those for the facts that bivaluations and Fidel structures both characterize our logics, or that these systems are not algebraizable according to Blok and Pigozzi, are incredibly similar to the proofs of these results for logics of formal inconsistency, the only difference being that some amount of care must be taking while dealing with incompatibility; however, the proofs in this section require formulas $\\phi_{ij}$ specially composed to prove the non-characterizability (by either finite Nmatrices or finite Rmatrices) of systems of incompatibility, and are therefore intrinsically different from any demonstrations produced on the field of $\\textbf{LFI}$'s.\n\n\nWe start by supposing there exists an Nmatrix $\\mathcal{M}=(\\mathcal{A}, D)$ that characterizes $\\bI$, with $A$ the universe of $\\mathcal{A}$ and $U=A\\setminus D$ the set of undesignated elements; we also remember we have defined $\\bot_{\\alpha\\beta}$ as $\\alpha\\wedge(\\beta\\wedge(\\alpha\\uparrow\\beta))$ and ${\\sim}\\alpha$ as $\\alpha\\rightarrow\\bot_{\\alpha\\alpha}$, since both of these will play an important role in the proof to come. For simplicity, we will drop the indexes from the operations on $\\mathcal{A}$ and use the infix notation.\n\n\\begin{lemma}\\label{basic facts about Nmatrix semantics}\nSuppose $d_{1}, d_{2}\\in D$ and $u_{1}, u_{2}\\in U$:\n\\begin{enumerate}\n\\item $d_{1}\\wedge d_{2}\\subseteq D$, while $d_{1}\\wedge u_{1}$, $u_{1}\\wedge d_{1}$ and $u_{1}\\wedge u_{2}$ are subsets of $U$;\n\\item$d_{1}\\rightarrow d_{2}$, $u_{1}\\rightarrow d_{1}$ and $u_{1}\\rightarrow u_{2}$ are subsets of $D$, while $d_{1}\\rightarrow u_{1}\\subseteq U$;\n\\item $d_{1}\\vee d_{2}$, $d_{1}\\vee u_{1}$ and $u_{1}\\vee d_{1}$ are subsets of $D$, while $u_{1}\\vee u_{2}\\subseteq U$;\n\\item for any formulas $\\alpha$ and $\\beta$ and valuation $\\nu:\\textbf{F}(\\Sigma_{\\bI}, \\mathcal{V})\\rightarrow \\mathcal{A}$ for $\\mathcal{A}$, $\\nu(\\bot_{\\alpha\\beta})\\in U$;\n\\item for any formula $\\alpha$ and valuation $\\nu$, if $\\nu(\\alpha)\\in D$, then $\\nu({\\sim}\\alpha)\\in U$; and if $\\nu(\\alpha)\\in U$, then $\\nu({\\sim}\\alpha)\\in D$.\n\\end{enumerate}\n\\end{lemma}\n\n\\begin{proof}\n\\begin{enumerate}\n\\item Given variables $p$ and $q$, from $\\textbf{Ax\\: 4}$ and $\\textbf{Ax\\: 5}$ one gets $p\\wedge q\\vdash_{\\bI}p$ and $p\\wedge q\\vdash_{\\bI}q$; given a valuation $\\nu$ for $\\mathcal{A}$, if $\\nu(p)\\in U$ or $\\nu(q)\\in U$, then $\\nu(p\\wedge q)\\in U$, and therefore $\\nu(p)\\wedge\\nu(q)\\subseteq U$: this is the case since, otherwise, one could define a valuation $\\nu^{*}: \\textbf{F}(\\Sigma_{\\bI}, \\mathcal{V})\\rightarrow \\mathcal{A}$ such that $\\nu^{*}(p)=\\nu(p)$, $\\nu^{*}(q)=\\nu(q)$ but $\\nu^{*}(p\\wedge q)\\in D$, meaning that either $p\\wedge q\\vdash_{\\bI}p$ or $p\\wedge q\\vdash_{\\bI}q$ is not validated by $\\mathcal{M}$.\n\nFrom $\\textbf{Ax\\: 3}$ and the deduction meta-theorem, we obtain $p, q\\vdash_{\\bI}p\\wedge q$, meaning that if $\\nu(p), \\nu(q)\\in D$, then $\\nu(p\\wedge q)\\in D$, and so $\\nu(p)\\wedge\\nu(q)\\subseteq D$.\n\n\\item We have that, for variables $p$ and $q$, $q\\rightarrow (p\\rightarrow q)$ is an instance of an axiom schema of $\\bI$, and therefore $q\\vdash_{\\bI}p\\rightarrow q$; taking a homomorphism $\\nu:\\textbf{F}(\\Sigma_{\\bI}, \\mathcal{V})\\rightarrow \\mathcal{A}$ such that $\\nu(q)\\in D$, we must have $\\nu(p\\rightarrow q)\\in D$ and therefore $\\nu(p)\\rightarrow \\nu(q)\\subseteq D$. This, of course, corresponds to both $d_{1}\\rightarrow d_{2}$ and $u_{1}\\rightarrow d_{1}$ being subsets of $D$.\n\nNow, $p, p\\rightarrow q\\vdash_{\\bI} q$, and therefore if $\\nu(q)\\in U$, we must have either $\\nu(p)\\in U$ or $\\nu(p\\rightarrow q)\\in U$; so, if $\\nu(q)\\in U$ and $\\nu(p)\\in D$, one must necessarily have $\\nu(p\\rightarrow q)\\in U$, meaning $\\nu(p)\\rightarrow\\nu(q)\\subseteq U$, corresponding to $d_{1}\\rightarrow u_{1}\\subseteq U$. \n\nFinally, suppose $\\nu(p), \\nu(q)\\in U$ and, without loss of generality, $\\nu(r)\\in D$; from $\\textbf{Ax\\: 2}$ and the deduction meta-theorem, $p\\rightarrow (q\\rightarrow r)\\vdash_{\\bI}(p\\rightarrow q)\\rightarrow(p\\rightarrow r)$ and, from the previous investigations, one finds that $\\nu(q\\rightarrow r)\\in D$ and $\\nu(p\\rightarrow(q\\rightarrow r))\\in D$, meaning $\\nu((p\\rightarrow q)\\rightarrow(p\\rightarrow r))\\in D$. Since $\\nu(p\\rightarrow r)\\in D$, one necessarily obtains that $\\nu(p\\rightarrow q)\\in D$, forcibly implying that $\\nu(p)\\rightarrow\\nu(q)\\subseteq D$, what corresponds to $u_{1}\\rightarrow u_{2}\\subseteq D$.\n\n\\item Given variables $p$ and $q$, we have that $p\\vdash_{\\bI}p\\vee q$ and $q\\vdash_{\\bI}p\\vee q$ from axiom schemata $\\textbf{Ax\\: 6}$ and $\\textbf{Ax\\: 7}$ and the deduction meta-theorem. So, if $\\nu(p)\\in D$ or $\\nu(q)\\in D$, one necessarily finds $\\nu(p\\vee q)\\in D$, and therefore $\\nu(p)\\rightarrow\\nu(q)\\subseteq D$.\n\nNow, from axiom schema $\\textbf{Ax\\: 8}$ and the deduction meta-theorem, $p\\rightarrow r, q\\rightarrow r\\vdash_{\\bI}(p\\vee q)\\rightarrow r$; if $\\nu(p)\\in U$ and $\\nu(q)\\in U$, suppose, without loss of generality, that $\\nu(r)\\in U$; this means $\\nu(p\\rightarrow r)$ and $\\nu(q\\rightarrow r)$ are both in $D$, and so must be $\\nu((p\\vee q)\\rightarrow r)$, what only happens if $\\nu(p\\vee q)\\in U$. This, of course, implies that $\\nu(p)\\vee\\nu(q)\\subseteq U$.\n\n\\item Let $p$, $q$ and $r$ be propositional variables: from $\\textbf{Ip}$ and the deduction meta-theorem one finds that $p, q, p\\uparrow q\\vdash_{\\bI}r$, yet $p, q\\not\\vdash_{\\bI}r$.\\footnote{To see that, take a bivaluation $\\nu$ for $\\bI$ with $\\nu(p)=\\nu(q)=1$ and $\\nu(r)=0$} If we suppose $\\nu$ is a valuation for $\\mathcal{A}$ such that $\\nu(p), \\nu(q)\\in D$ and $\\nu(r)\\in U$, it becomes clear that one must have $\\nu(p\\uparrow q)\\in U$, since otherwise one would forcibly have $\\nu(r)\\in D$. This means we can never have all three $\\nu(\\alpha)$, $\\nu(\\beta)$ and $\\nu(\\alpha\\uparrow\\beta)$ in $D$, and therefore \n\\[\\nu(\\bot_{pq})=(\\nu(p)\\wedge\\nu(q))\\wedge\\nu(p\\uparrow q)\\in U.\\]\nOf course, for arbitrary formulas $\\alpha$ and $\\beta$, $\\bot_{\\alpha\\beta}\\vdash_{\\bI}\\bot_{pq}$ and $\\bot_{pq}\\vdash_{\\bI}\\bot_{\\alpha\\beta}$, meaning that, for any $\\nu$, $\\nu(\\bot_{\\alpha\\beta})\\in U$.\n\n\\item If $\\nu(\\alpha)\\in D$, since $\\nu(\\bot_{\\alpha\\alpha})\\in U$ we obtain $\\nu({\\sim}\\alpha)=\\nu(\\alpha\\rightarrow\\bot_{\\alpha\\alpha})\\in U$. Reciprocally, if $\\nu(\\alpha)\\in U$, $\\nu({\\sim}\\alpha)=\\nu(\\alpha\\rightarrow\\bot_{\\alpha\\alpha})\\in D$.\n\\end{enumerate}\n\\end{proof}\n\n\n\n\\begin{lemma}\n\\begin{enumerate}\n\\item For any two elements $a, b\\in A$, either $a\\uparrow b\\subseteq D$ or $a\\uparrow b\\subseteq U$;\n\\item for any two elements $a, b\\in A$, either both $a\\uparrow b$ and $b\\uparrow a$ are subsets of $D$, or both are subsets of $U$.\n\\end{enumerate}\n\\end{lemma}\n\n\\begin{proof}\nSuppose that there are values $d, u\\in a\\uparrow b$ such that $d\\in D$ and $u\\in U$, and let $p$ and $q$ be propositional variables throughout the proof. Since $b\\uparrow a$ is necessarily not empty, it must contain either an element $d^{*}\\in D$, or an $u^{*}\\in U$.\n\\begin{enumerate}\n\\item In the first case, take a valuation $\\nu:\\textbf{F}(\\Sigma_{\\bI}, \\mathcal{V})\\rightarrow\\mathcal{A}$ satisfying $\\nu(p)=a$, $\\nu(q)=b$, $\\nu(p\\uparrow q)=u$ and $\\nu(q\\uparrow p)=d^{*}$, and then we have that $\\nu((q\\uparrow p)\\rightarrow(p\\uparrow q))\\in d^{*}\\rightarrow u$, which by Lemma \\ref{basic facts about Nmatrix semantics} is contained in $U$. \n\nThis shows $\\nu$ does not validate $\\textbf{Comm}$, what is absurd given that $\\mathcal{M}$ characterizes $\\bI$.\n\\item In the second case, take now a valuation $\\nu$ with $\\nu(p)=a$, $\\nu(q)=b$, $\\nu(p\\uparrow q)=d$ and $\\nu(q\\uparrow p)=u^{*}$. Then $\\nu((p\\uparrow q)\\rightarrow(q\\uparrow p))\\in d\\rightarrow u^{*}$, again contained in $U$ according to Lemma \\ref{basic facts about Nmatrix semantics}.\n\\end{enumerate}\n\nEither way we reach a contradiction, an the conclusion must be that either $a\\uparrow b$ is contained in $D$, or it is contained in $U$.\n\nFinally, suppose that there are values $a, b\\in A$ with $a\\uparrow b\\subseteq D$ and $b\\uparrow a\\subseteq U$, and then, for any valuation $\\nu$ with $\\nu(p)=a$ and $\\nu(q)=b$ (and there are many of them), one necessarily finds that $\\nu(p\\uparrow q)\\in D$ and $\\nu(q\\uparrow p)\\in U$, and from Lemma \\ref{basic facts about Nmatrix semantics} we have $\\nu((p\\uparrow q)\\rightarrow(q\\uparrow p))\\in U$; this again contradicts the fact that $\\mathcal{M}$ characterizes $\\bI$, since it implies that $\\nu$ does not model $\\textbf{Comm}$.\n\\end{proof}\n\nThe following fact is trivial, but important in the following discussion so we make a point of proving it: suppose $\\Gamma$ is a set of formulas of $\\bI$ such that there exists $\\varphi$ with the property that $\\Gamma\\not\\vdash_{\\bI}\\varphi$; then we state that there exists a valuation $\\nu$ for $\\mathcal{A}$ with $\\nu(\\Gamma)\\subseteq D$ but $\\nu(\\varphi)\\in U$. This is true since, otherwise, if we had that for every valuation $\\nu$ for $\\mathcal{A}$ satisfying $\\nu(\\Gamma)\\subseteq D$ one also had $\\nu(\\varphi)\\in D$, this would imply $\\Gamma\\vDash_{\\mathcal{M}}\\varphi$. Since $\\mathcal{M}$ characterizes $\\bI$, this would mean $\\Gamma\\vdash_{\\bI}\\varphi$, against our suppositions.\n\nConsider then two disjoint sets of distinct variables $\\{p_{n}\\ :\\ n\\in\\mathbb{N}\\}$ and $\\{q_{n}\\ :\\ n\\in\\mathbb{N}\\}$ and the formulas, for $i, j\\in\\mathbb{N}$,\n\\[\\phi_{ij}=\\begin{cases*}\n \\quad p_{i}\\uparrow q_{j} & if $i, line width=1pt, line cap=round, dash pattern=on 0pt off 2\\pgflinewidth] (1.25,3.75) -- (3.75, 1.25) node[midway,above right] {Ideal};\n\\end{axis}\n\n\\end{tikzpicture}\n\\end{figure}\n\nUsing the previous diagram, we can see that, in $\\bI$'s case, we are able to start at some point in the far top-left, where $\\textbf{2}(\\bI)$ lies, and navigate to the bottom-right without ever leaving the, at least on the diagram, loosely-defined environment of restricted non-deterministic matrices, arriving at $\\mathbb{2}_{\\bI}$; notice this is very much in line with what we previously observed, that going from the first of these RNmatrices to the second involves restricting the underlying operations while expanding the accepted homomorphisms. Worthy of notice is that, while working in $\\bI$, we are never able to leave RNmatrices since $\\bI$ can not be characterized by either (finite) logical matrices, Nmatrices or Rmatrices.\n\nIn a different case, consider $\\bI^{-}$: we know that it admits a finite RNmatrix, namely $\\textbf{2}(\\bI^{-})$, located somewhere on the top-left region of the schematic diagram; but, in this case, we also have a finite Nmatrix which characterizes $\\bI^{-}$, although no (finite) logical matrices are available as far as we know. So, at least in this logic's case, navigating ever bottom-right indeed leads outside the environment of RNmatrices and into the methodology of Nmatrices.\n\nNow, as we have assigned numbers to the other generalizations of a finite logical matrix, let us take $(4)$ as exchanging the underlying algebra of a matrix for a multialgebra (making a logical matrix into an Nmatrix, an Rmatrix into an RNmatrix and so on), and $(5)$ as restricting the valuations to be taken into consideration (turning a logical matrix into an Rmatrix and so on).\n\n\\begin{problem}\nWhich combinations of generalizations of a (finite) logical matrix have the expressive power to characterize every tarskian logic?\n\\end{problem}\n\nW\\'ojciki proved, in \\cite{Woj} and \\cite{Woj2}, that the combinations of $(1)$ and $(2)$, and $(1)$ and $(3)$, corresponding to classes of potentially infinite logical matrices and potentially infinite matrices with multiple sets of distinguished elements, are enough to characterize all tarskian logics; Piochi, in \\cite{Piochi}, proved that $(1)$ plus $(5)$ is also enough, and as we have proved that every tarskian logic is characterizable by a two-elements RNmatrix (again Theorem \\ref{2-valued RNmatrix}), $(4)$ plus $(5)$ is also enough. From this, it is clear that combining any four of these conditions characterizes all such logics, although it seems the same can not yet be said about combinations of three of them, as, for example, we were not able to find a reference concerning $(2)$ plus $(3)$ and $(5)$ or any subset thereof. So, we have the following table to summarize these observations.\n\n\\begin{figure}[H]\n\\centering\n\\begin{tabular}{cc|ccccc}\n& & \\multicolumn{5}{c}{Generalizations} \\\\[5pt]\n & & (1) & (2) & (3) & (4) & (5) \\\\[5pt] \\cline{2-7}\n\\multirow{5}{*}{\\rotatebox[origin=c]{90}{Generalizations}}\n& (1) & & \\checkmark & \\checkmark & \\checkmark & ? \\\\[5pt]\n & (2) & \\checkmark & & ? & ? & ? \\\\[5pt]\n & (3) & \\checkmark & ? & & ? & ? \\\\[5pt]\n& (4) & \\checkmark & ? & ? & & \\checkmark \\\\[5pt]\n & (5) & ? & ? & ? & \\checkmark & \\\\[5pt]\n\\end{tabular}\n\\end{figure}\n\nWe have left the diagonal of the table empty, as it does not deal with combinations of generalizations \\textit{per se}, but rather the generalizations themselves. Of course, there is nothing stopping us from considering which, if any, of the generalizations can characterize all tarskian logics, leading to our second problem.\n\n\\begin{problem}\nAre there any generalizations which, alone, can characterize all tarskian logics?\n\\end{problem}\n\nAs we have shown, $\\bI$ is not characterizable by either finite Nmatrices or finite Rmatrices, so we can already place some limitatitive results in our table.\n\n\\begin{figure}[H]\n\\centering\n\\begin{tabular}{cc|ccccc}\n& & \\multicolumn{5}{c}{Generalizations} \\\\[5pt]\n & & (1) & (2) & (3) & (4) & (5) \\\\[5pt] \\cline{2-7}\n\\multirow{5}{*}{\\rotatebox[origin=c]{90}{Generalizations}}\n& (1) & ? & \\checkmark & \\checkmark & \\checkmark & ? \\\\[5pt]\n & (2) & \\checkmark & ? & ? & ? & ? \\\\[5pt]\n & (3) & \\checkmark & ? & ? & ? & ? \\\\[5pt]\n& (4) & \\checkmark & ? & ? & \\ding{55} & \\checkmark \\\\[5pt]\n & (5) & ? & ? & ? & \\checkmark & \\ding{55} \\\\[5pt]\n\\end{tabular}\n\\end{figure}\n\nOf course, the first course of action one probably thinks of, when finding the problems mentioned above and looking at the tables, is attempting to fill in the missing results; this does not seem impossible, but it does not seem trivial either. Proving that certain semantics can not characterize all tarskian logics involves most probably presenting counter-examples, preferably examples among already known and studied systems.\n\nBut one can always increase the rows and columns in our little illustrative tables, by considering other generalizations of logical matrices. Here, for one, we have not included making the operations of a matrix partial, instead of non-deterministic: when applied to Nmatrices, this procedure returns the semantical objects known as PNmatrices. These play a unique role indeed, as they are weaker than RNmatrices, yet we do not know if strictly so or if both semantics characterize the same logics. We are then tempted to consider a hierarchy of strength among combinations of generalizations of logical matrices, instead of the simply binary ``do they characterize all tarskian logics?'', and things start to become involved... \n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\\newpage\n\\printbibliography[segment=\\therefsegment,heading=subbibliography]\n\\end{refsegment}\n\n\n\\begin{refsegment}\n\\defbibfilter{notother}{not segment=\\therefsegment}\n\\setcounter{chapter}{8}\n\\chapter{Translating paraconsistent logics}\\label{Chapter9}\\label{Chapter 9}\n\n\n\n\nTake the signature \\label{SigmaLFICPL}$\\Sigma_{\\textbf{LFI}}^{\\textbf{CPL}}$ such that $(\\Sigma_{\\textbf{LFI}}^{\\textbf{CPL}})_{0}=\\{\\bot, \\top\\}$, $(\\Sigma_{\\textbf{LFI}}^{\\textbf{CPL}})_{1}=\\{\\neg, {\\sim}, \\circ\\}$, $(\\Sigma_{\\textbf{LFI}}^{\\textbf{CPL}})_{2}=\\{\\vee, \\wedge, \\rightarrow\\}$ and $(\\Sigma_{\\textbf{LFI}}^{\\textbf{CPL}})_{n}=\\emptyset$ for $n>2$.\n\nWe classically define a Fidel structure to be a $\\Sigma_{\\textbf{LFI}}^{\\textbf{CPL}}-$multialgebra $\\mathcal{E}=(A, \\{\\sigma_{\\mathcal{E}}\\}_{\\sigma \\in\\Sigma_{\\textbf{LFI}}^{\\textbf{CPL}}})$ such that:\n\\begin{enumerate}\n\\item $(A, \\{\\sigma_{\\mathcal{E}}\\}_{\\sigma\\in\\Sigma^{\\textbf{CPL}}})$ is a Boolean algebra;\n\\item for every $a\\in A$ and $b\\in \\neg a$, $a\\vee b=\\top$;\n\\item for every $a\\in A$ and $b\\in \\neg a$, there exists a non-empty subset $O_{ab}$ of $\\circ a$, defined case by case, designed to capture the logical structure intended to be emulated by the Fidel structure.\n\\end{enumerate}\n\nIntuitively, one looks at $\\neg a$ as all possible negations of $a$, and at $O_{ab}$ as all possible consistencies for $a$ given that $b$ is its negation. To give one example, in $\\textbf{mbC}$, where we usually denote $O_{ab}$ by $BC_{ab}$, we require that\n\\[a\\wedge (b\\wedge c)=\\bot,\\quad \\forall b\\in \\neg a,\\quad \\forall c\\in BC_{ab}.\\]\n\nWhen looking at previous instances in this text of a \"Fidel structure\", maybe the most important distinction was the intuitive replacement of consistency for incompatibility: so, for example, instead of the axiom schema $\\textbf{bc1}$ of $\\textbf{mbC}$ given by $\\circ\\alpha\\rightarrow(\\alpha\\rightarrow(\\neg\\alpha\\rightarrow\\beta))$, we used a similar, but distinct, schema $\\textbf{Ip}$, given by $(\\alpha\\uparrow\\beta)\\rightarrow(\\alpha\\rightarrow(\\beta\\rightarrow\\gamma))$.\n\nOne clearly notices how our binary incompatibility operator, the generalized Sheffer's stroke, as studied above seems inherently distinct from the unary consistency operator \"$\\circ$\". In fact, one must translate accordingly the axiomatization of $\\textbf{LFI}$'s to that of our $\\textbf{LIp}$'s to show that the latter generalize the former, and logics such as $\\textbf{mbC}$, $\\textbf{mbCciw}$, $\\textbf{mbCci}$ and $\\textbf{mbCcl}$ become sublogics of, respectively, $\\nbI$, $\\nbIciw$, $\\nbIci$ and $\\nbIcl$. \n\nBut, and this is the important finding of this chapter, our intuition can still be validated: inconsistency, at least as found in the simpler paraconsistent logics here exhibited, can be obtained from incompatibility, well-behaved formulas being precisely those formulas incompatible with their negations. Since there are systems on which incompatibility clearly does not reduce back to inconsistency, such as $\\textbf{bI}$ (which does not even have a paraconsistent negation), we must reach the conclusion that incompatibility appears to non-trivially generalize the notion of inconsistency, giving some extra validation to the work we performed so far. \n\nThe developments found here have been submitted, as a preprint, in \\cite{Frominconsistency}.\n\n\\section{Preliminaries}\n\nSo, let us define a translation from the usual signature $\\Sigma_{\\textbf{LFI}}$ for $\\textbf{LFI}'s$, with $(\\Sigma_{\\textbf{LFI}})_{1}=\\{\\neg , \\circ\\}$, $(\\Sigma_{\\textbf{LFI}})_{2}=\\{\\vee, \\wedge, \\rightarrow\\}$ and $(\\Sigma_{\\textbf{LFI}})_{n}=\\emptyset$ for $n\\notin\\{1,2\\}$, to the signature $\\Sigma_{\\nbI}$. For $\\mathcal{V}$ a countable set of propositional variables, consider the function\\label{T} \n\\[T:F(\\Sigma_{\\textbf{LFI}}, \\mathcal{V})\\rightarrow F(\\Sigma_{\\nbI}, \\mathcal{V})\\]\nsuch that:\n\\begin{enumerate}\n\\item $T(p)=p$ for every $p\\in \\mathcal{V}$;\n\\item $T(\\neg\\alpha)=\\neg T(\\alpha)$;\n\\item $T(\\alpha\\#\\beta)=T(\\alpha)\\# T(\\beta)$ for every $\\#\\in\\{\\vee, \\wedge, \\rightarrow\\}$;\n\\item $T(\\circ\\alpha)=T(\\alpha)\\uparrow\\neg T(\\alpha)$.\n\\end{enumerate}\nEssentially, $T$ changes all occurrences of the form $\\circ\\alpha$ to $\\alpha\\uparrow\\neg\\alpha$. What one does is then to consider for a set of axiom schemata $\\Gamma$ of an $\\textbf{LFI}$ its translation $T(\\Gamma)=\\{T(\\psi)\\ :\\ \\psi\\in \\Gamma\\}$: to give one example, many instances of $\\textbf{Ip}$ are translations of instances of $\\textbf{bc1}$.\n\n\\begin{proposition}\n$T$ is an injective function.\n\\end{proposition}\n\n\\begin{proof}\nSuppose $T(\\alpha)=T(\\beta)$: we proceed by double induction on the orders of $\\alpha$ and $\\beta$, to show that in this case $\\alpha=\\beta$. \n\nIf $\\alpha$ is of order $0$, then $\\alpha=p$ for some $p\\in \\mathcal{V}$ and therefore $T(\\alpha)=\\alpha$: if $\\beta$ is of order $0$, then it is a propositional variable $q$, and then again $T(\\beta)=\\beta$, implying we have $\\alpha=T(\\alpha)=T(\\beta)=\\beta$.\n\nSo, assume $\\beta$ is of order at least $1$, and we show that we can not actually have, in this case, $T(\\alpha)=T(\\beta)$:\n\\begin{enumerate}\n\\item if $\\beta=\\neg\\beta_{0}$, $T(\\beta)=\\neg T(\\beta_{0})=T(\\alpha)$, which is absurd since $T(\\alpha)$ is a propositional variable and can not contain an unary connective;\n\\item if $\\beta=\\beta_{0}\\#\\beta_{1}$, for $\\#\\in\\{\\vee, \\wedge, \\rightarrow\\}$, $T(\\beta)=T(\\beta_{0})\\# T(\\beta_{1})=T(\\alpha)$, which is again absurd since $T(\\alpha)$ is a propositional variable;\n\\item if $\\beta=\\circ\\beta_{0}$, $T(\\beta)=T(\\beta_{0})\\uparrow\\neg T(\\beta_{0})$, which clearly contains even more than one connective and therefore can not equal $T(\\alpha)$.\n\\end{enumerate}\n\nNow, suppose that for every $\\alpha$ of order at most $m$ we have that, if $T(\\alpha)=T(\\beta)$, then $\\alpha=\\beta$, and take a formula $\\alpha$ of order $m+1$: we have that either $\\alpha=\\alpha_{0}\\#\\alpha_{1}$, for $\\#\\in\\{\\vee, \\wedge, \\rightarrow\\}$, $\\alpha=\\neg\\alpha_{0}$ or $\\alpha=\\circ\\alpha_{0}$, with the orders of $\\alpha_{0}$ and $\\alpha_{1}$ being at most $m$.\n\nIf $\\beta$ is of order $0$, then $\\beta=q$ for some $q\\in \\mathcal{V}$, and in that case $T(\\beta)=\\beta$ is a formula of order $0$: we then can not really have $T(\\alpha)=T(\\beta)$, since $T(\\alpha)$ is either $\\neg T(\\alpha_{0})$, $T(\\alpha_{0})\\# T(\\alpha_{1})$ or $T(\\alpha_{0})\\uparrow\\neg T(\\alpha_{0})$, none of which is of order $0$; by vacuity, the lemma holds. \n\nInductively, suppose that for all $\\beta$ of order at most $n$, if $T(\\alpha)=T(\\beta)$ then $\\alpha=\\beta$, and take $\\beta$ of order $n+1$: we have that either $\\beta=\\beta_{0}\\ast\\beta_{1}$, for $\\ast\\in\\{\\vee, \\wedge, \\rightarrow\\}$, $\\beta=\\neg\\beta_{0}$ or $\\beta=\\circ\\beta_{0}$, with the orders of $\\beta_{0}$ and $\\beta_{1}$ being at most $n$.\n\nIf $\\alpha=\\alpha_{0}\\#\\alpha_{1}$, $T(\\alpha)=T(\\alpha_{0})\\# T(\\alpha_{1})$: in this case, $\\beta$ can not equal $\\neg\\beta_{0}$, for in this case we would have $T(\\beta)=\\neg T(\\beta_{0})$ which has a leading connective of arity different from that of $T(\\alpha)$; for similar reasons we can not have $\\beta=\\circ\\beta_{0}$ or $\\beta=\\beta_{0}\\ast\\beta_{1}$ for $\\ast$ different of $\\#$. The same can be done when $\\alpha=\\neg\\alpha_{0}$, in which case $\\beta$ must also be of the form $\\neg\\beta_{0}$, and when $\\alpha=\\circ\\alpha_{0}$, when we must have $\\beta=\\circ\\beta_{0}$.\n\nSo there remains three cases to check:\n\\begin{enumerate}\n\\item if $\\alpha=\\alpha_{0}\\#\\alpha_{1}$ and $\\beta=\\beta_{0}\\#\\beta_{1}$, $T(\\alpha)=T(\\beta)$ implies that $T(\\alpha_{0})\\# T(\\alpha_{1})=T(\\beta_{0})\\# T(\\beta_{1})$, and so $T(\\alpha_{0})=T(\\beta_{0})$ and $T(\\alpha_{1})=T(\\beta_{1})$; by our induction hypothesis, $\\alpha_{0}=\\beta_{0}$ and $\\alpha_{1}=\\beta_{1}$, so that $\\alpha=\\beta$;\n\\item if $\\alpha=\\neg\\alpha_{0}$ and $\\beta=\\neg\\beta_{0}$, $T(\\alpha)=T(\\beta)$ implies that $\\neg T(\\alpha_{0})=\\neg T(\\beta_{0})$ and so $T(\\alpha_{0})=T(\\beta_{0})$; by our induction hypothesis, $\\alpha_{0}=\\beta_{0}$ and therefore $\\alpha=\\beta$;\n\\item if $\\alpha=\\circ\\alpha_{0}$ and $\\beta=\\circ\\beta_{0}$, $T(\\alpha)=T(\\beta)$ implies that \n\\[T(\\alpha_{0})\\uparrow\\neg T(\\alpha_{0})=T(\\beta_{0})\\uparrow\\neg T(\\beta_{0})\\]\nand so $T(\\alpha_{0})=T(\\beta_{0})$; by our induction hypothesis, $\\alpha_{0}=\\beta_{0}$ and therefore $\\alpha=\\beta$.\n\\end{enumerate}\nThis, of course, finishes the proof.\n\\end{proof}\n\nOne important thing to notice is that a formula $\\alpha$ in $F(\\Sigma_{\\nbI},\\mathcal{V})$ is not in $T(F(\\Sigma_{\\textbf{LFI}},\\mathcal{V}))$ if and only if it contains a subformula $\\beta_{1}\\uparrow\\beta_{2}$ such that $\\beta_{2}\\neq\\neg\\beta_{1}$.\n\nOne direction is clear: if $\\alpha$ contains a subformula $\\beta_{1}\\uparrow\\beta_{2}$ with $\\beta_{2}\\neq\\neg \\beta_{1}$ then it is not in $T(F(\\Sigma_{\\textbf{LFI}},\\mathcal{V}))$, since such a formula is not a translation of anything over the signature $\\Sigma_{\\textbf{LFI}}$.\n\nReciprocally, we proceed by induction on the order of $\\alpha$: if it is $0$, $\\alpha$ is a propositional variable, and it is always the case that $\\alpha$ is a translation; so assume that $\\alpha$ is of order $n>1$ and that the result holds for formulas of order smaller than $n$. Then we have three cases to consider:\n\\begin{enumerate}\n\\item if $\\alpha=\\alpha_{0}\\#\\alpha_{1}$, for $\\#\\in\\{\\vee, \\wedge, \\rightarrow\\}$, and $\\alpha_{0}$ and $\\alpha_{1}$ are translations, so is $\\alpha$; therefore, if $\\alpha\\notin T(F(\\Sigma_{\\textbf{LFI}},\\mathcal{V}))$, then one of $\\alpha_{0}$ or $\\alpha_{1}$ has, by induction hypothesis, a subformula $\\beta_{1}\\uparrow\\beta_{2}$ with $\\beta_{2}\\neq\\neg \\beta_{1}$, and the result holds;\n\\item if $\\alpha=\\neg\\alpha_{0}$ and $\\alpha_{0}$ is a translation, so is $\\alpha$; therefore, if $\\alpha\\notin T(F(\\Sigma_{\\textbf{LFI}},\\mathcal{V}))$, then $\\alpha_{0}\\notin T(F(\\Sigma_{\\textbf{LFI}},\\mathcal{V}))$ and thus has, by induction hypothesis, a subformula of the desired form, making the result hold once again;\n\\item finally, if $\\alpha=\\alpha_{0}\\uparrow\\alpha_{1}$ but $\\alpha\\notin T(F(\\Sigma_{\\textbf{LFI}},\\mathcal{V}))$, either $\\alpha_{0}$ and $\\alpha_{1}$ are translations but $\\alpha_{1}\\neq\\neg\\alpha_{0}$, and the result holds; or one of $\\alpha_{0}$ and $\\alpha_{1}$ is not a translation, and therefore contains a subformula $\\beta_{1}\\uparrow\\beta_{2}$ with $\\beta_{2}\\neq\\neg\\beta_{1}$, what ends the proof.\n\\end{enumerate}\n\n\n\\section{$T$ is a conservative translation}\n\nGiven logics $\\mathcal{L}_{1}$ and $\\mathcal{L}_{2}$ over the signatures $\\Sigma_{1}$ and $\\Sigma_{2}$, a function $\\mathcal{T}:F(\\Sigma_{1}, \\mathcal{V})\\rightarrow F(\\Sigma_{2},\\mathcal{V})$ is said to be a translation\\index{Translation}, originally defined in \\cite{Itala}, when, for every set of formulas $\\Gamma\\cup\\{\\varphi\\}$ over the signature $\\Sigma_{1}$,\n\\[\\Gamma\\vdash_{\\mathcal{L}_{1}}\\varphi\\quad\\text{implies}\\quad\\mathcal{T}(\\Gamma)\\vdash_{\\mathcal{L}_{2}}\\mathcal{T}(\\varphi);\\]\na translation $\\mathcal{T}$ is said to be a conservative translation\\index{Translation, Conservative} whenever, for every set of formulas $\\Gamma\\cup\\{\\varphi\\}$ over the signature $\\Sigma_{1}$,\n\\[\\mathcal{T}(\\Gamma)\\vdash_{\\mathcal{L}_{2}}\\mathcal{T}(\\varphi)\\quad\\text{implies}\\quad\\Gamma\\vdash_{\\mathcal{L}_{1}}\\varphi.\\]\n One may look at definition $2.4.1$ of \\cite{ParLog} for a reference for our definition, or the original work concerning conservative translations, \\cite{ConservativeTranslationsReference}: such a notion is recurring in a contemporary approach to logic, where systems may have been formulated in apparently distinct ways that prove to be, under translation, equivalent; coincidentally, we have already seen translations very briefly in Example \\ref{PTS}.\n\nWe shall prove that the function $T$ we previously defined is a translation, and furthermore, a conservative one in many cases. To prove the following lemma, and consequently the following theorem, remember a $\\Sigma$-homomorphism $\\sigma:F(\\Sigma, \\mathcal{V})\\rightarrow F(\\Sigma, \\mathcal{V})$ is determined by its action on $\\mathcal{V}$, given $F(\\Sigma, \\mathcal{V})$ is deterministic.\n\n\\begin{lemma}\\label{Existence of translated substitution}\nGiven a $\\Sigma_{\\LFI}$-homomorphism $\\sigma:F(\\Sigma_{\\LFI}, \\mathcal{V})\\rightarrow F(\\Sigma_{\\LFI}, \\mathcal{V})$, the $\\Sigma_{\\nbI}$-homomor\\-phism $\\overline{\\sigma}:F(\\Sigma_{\\nbI}, \\mathcal{V})\\rightarrow F(\\Sigma_{\\nbI}, \\mathcal{V})$ given by, for a propositional variable $p\\in\\mathcal{V}$, \n\\[\\overline{\\sigma}(p)=T(\\sigma(p))\\]\nsatisfies that, for any formula $\\alpha$ on $\\Sigma_{\\LFI}$, $T(\\sigma(\\alpha))=\\overline{\\sigma}(T(\\alpha))$.\n\\end{lemma}\n\n\\begin{proof}\nThe result is trivially true for formulas of order $0$, since they are invariant under $T$. So, proceeding inductively, assume the result holds for the formulas $\\alpha$ and $\\beta$:\n\\begin{enumerate}\n\\item for $\\#\\in\\{\\vee, \\wedge, \\rightarrow\\}$, \n\\[T(\\sigma(\\alpha\\#\\beta))=T(\\sigma(\\alpha)\\#\\sigma(\\beta))=T(\\sigma(\\alpha))\\# T(\\sigma(\\beta))=\\overline{\\sigma}(T(\\alpha))\\#\\overline{\\sigma}(T(\\beta))=\\]\n\\[\\overline{\\sigma}(T(\\alpha)\\# T(\\beta))=\\overline{\\sigma}(T(\\alpha\\#\\beta));\\]\n\\item $T(\\sigma(\\neg\\alpha))=T(\\neg\\sigma(\\alpha))=\\neg T(\\sigma(\\alpha))=\\neg \\overline{\\sigma}(T(\\alpha))=\\overline{\\sigma}(\\neg T(\\alpha))=\\overline{\\sigma}(T(\\neg\\alpha))$;\n\\item finally, remembering $\\overline{\\sigma}$ is a homomorphism,\n\\[T(\\sigma(\\circ\\alpha))=T(\\circ\\sigma(\\alpha))=T(\\sigma(\\alpha))\\uparrow\\neg T(\\sigma(\\alpha))=\\overline{\\sigma}(T(\\alpha))\\uparrow\\neg\\overline{\\sigma}(T(\\alpha))=\\]\n\\[\\overline{\\sigma}(T(\\alpha))\\uparrow\\overline{\\sigma}(\\neg T(\\alpha))=\\overline{\\sigma}(T(\\alpha)\\uparrow\\neg T(\\alpha))=\\overline{\\sigma}(T(\\circ\\alpha)),\\]\nwhat ends the proof.\n\\end{enumerate}\n\\end{proof}\n\n\\begin{theorem}\\label{Soundness of translations}\nIf $\\mathcal{L}$ is a logic over the signature $\\Sigma_{\\textbf{LFI}}$ with axiom schemata $\\Psi$ and $\\mathcal{L}^{*}$\\label{L*} is the logic over the signature $\\Sigma_{\\nbI}$ with axiom schemata $T(\\Psi)$, then $\\Gamma\\vdash_{\\mathcal{L}} \\varphi$ implies that $T(\\Gamma)\\vdash_{\\mathcal{L}^{*}} T(\\varphi)$.\n\\end{theorem}\n\n\\begin{proof}\nLet $\\alpha_{1}, \\dotsc , \\alpha_{n}$ be a demonstration of $\\varphi$ from $\\Gamma$, with $\\alpha_{n}=\\varphi$: we want to show that in this case $T(\\alpha_{1}), \\dotsc , T(\\alpha_{n})$ is a demonstration of $T(\\varphi)$; first of all, obviously $T(\\alpha_{n})=T(\\varphi)$. Then:\n\\begin{enumerate}\n\\item if $\\alpha_{i}$ is an instance of an axiom schema $\\psi$, $T(\\alpha_{i})$ is an instance of the axiom schema $T(\\psi)$, by an immediate application of Lemma \\ref{Existence of translated substitution};\n\\item if $\\alpha_{j}$ is in $\\Gamma$, $T(\\alpha_{j})\\in T(\\Gamma)$;\n\\item finally, given that our only rule of deduction is Modus Ponens, if $\\alpha_{k}$ is such that there exist $\\alpha_{i}$ and $\\alpha_{j}$ with $i, jt$. The Volterra series\ncan therefore also model memory effects (which are assumed to be\nof finite length) and it is not restricted to instantaneous effects.\n\n\n\nThe discretized form of the Volterra series truncated to second order (i.e.\\ $P=2$)\nis given by (\\cite[Eq.~2.25]{Mathews2000})\n\\begin{equation}\ny_n =h^{(0)}+\\sum_{j=0}^{R-1}h^{(1)}_j\\, x_{n{-}j}\n+\\sum_{k,\\ell=0}^{R-1}\nh^{(2)}_{k\\ell}\\, x_{n{-}k} \\, x_{n{-}\\ell},\\label{eqn:3.2}\n\\end{equation}\nThe discrete output entries $y_n$ have $N$ time steps with $n \\in \\{0,\\ldots,N{-}1\\}$\nwhich are obtained {from $L$ discrete input entries $x_q$ where $x_q=0$\nfor $q<0$. Note that\n$N = L +R -1 \\geq L$, where\n$R\\geq 1$} denotes the\nassumed memory length of the distortion.\nThe memory length $R$ quantifies how the response\nat the current time step depends\non the input of previous time steps, i.e., $R$ bounds the number\nof previous time steps that can affect the current one.\nVolterra kernel coefficients of\nthe zeroth, first, and second order are represented by\n$h^{(0)}$, $h^{(1)}_j$, and $h^{(2)}_{k\\ell}$.\nThe matrix given by $h^{(2)}_{k\\ell}$ is symmetric.\nWe are characterizing the transfer function by estimating\nthe kernel coefficients in Eq.~\\eqref{eqn:3.2}.\nThe number $M$ of the to-be-estimated coefficients scales quadratically with\nthe memory length $R$ (in general, the number of coefficients\nscales with $R^P$). Although the Volterra estimation can be\nextended to any higher order $P>2$, we will focus in this work on\nthe quadratic case.\n\nFor the estimation process, we assume that we are provided with\na training data set consisting of input-output pulse pairs $(x(t),y(t))$ from an experimental device\n(or a sequence of devices) which causes the distortion. Next, we discuss how given the\ntraining data, we can estimate the kernel coefficients in Eq.~\\eqref{eqn:3.2}\nby minimizing some error measures (such as the mean square error)\nbetween the modeled output and the measured output.\n\n\\subsection{Truncated Volterra series via least squares\\label{sec:level3b}}\nWe can choose from different methods to estimate the Volterra series.\nThe most widely used ones are the crosscorrelation method of\nLee and Schetzen \\cite{doi:10.1080\/00207176508905543} and\nthe exact orthogonal method of Korenberg \\cite{Korenberg:1988uv}.\nWe choose the latter due to its simplicity and\nas it does not require an infinite-length input.\nWe can write Eq.~\\eqref{eqn:3.2} as\n\\begin{align}\\label{eqn:3.3}\ny_n&=\\sum_{m=0}^{M-1}u_{nm}\\,k_m\n\\intertext{or equivalently as the matrix equation $Y = U K$ or}\n\\label{eqn:3.5}\n\\begin{bmatrix} y_0\\\\ y_1 \\\\[-1mm] \\vdots \\\\ y_{N-1} \\end{bmatrix}\n&=\n\\begin{bmatrix}\nu_{00} & u_{01} & \\cdots & u_{0,M-1} \\\\\nu_{10} & u_{11} & \\cdots & u_{1,M-1} \\\\[-1mm]\n\\vdots & \\vdots & & \\vdots \\\\\nu_{N-1,0} & u_{N-1,1} & \\cdots & u_{N-1,M-1} \\\\\n\\end{bmatrix} \\begin{bmatrix} k_0 \\\\ k_1 \\\\[-1mm] \\vdots \\\\k_{M-1}\n\\end{bmatrix},\n\\end{align}\nwhere $K$ is defined in Eq.~\\eqref{eqn:K} below.\nWe follow the convention that the entries\nof a given matrix (or vector) $D$ are\nrepresented by $d_{ij}$ (or $ d_i$).\nHere, $n\\in\\{0,\\ldots,N{-}1\\}$ and $m\\in \\{0,\\ldots,M{-}1\\}$ where\n\\begin{equation}\nM= 1{+} R {+} R(R{+}1)\/2\n\\label{eqn:3.11}\n\\end{equation}\ndenotes the number\nof coefficients that need to be estimated to describe\nthe quadratic Volterra series.\nIn particular, $u_{nm}$ are obtained from the input pulses\nvia (recall again $x_q=0$ for $q<0$)\n\\[\nu_{nm} =\n\\begin{cases}\n 1& \\text{for }\\; m=0,\\\\\n x_{n{-}m{+}1} & \\text{for }\\; m\\in \\{1,\\ldots,R\\},\\\\\n {x_{n{-}a}\\, x_{n{-}b}} & \\text{for }\\; m\\in \\{R{+}1,\\ldots,M{-}1\\},\n\\end{cases}\n\\]\n{where $(a,b)$ with $0\\leq a \\leq b \\leq R{-}1$ is the $(m{-}R{-}1)$th element\nin the lexicographically ordered sequence from $(0,0)$ to $(R{-}1,R{-}1)$.}\nAs the quadratic distortion coefficients $h^{(2)}_{k\\ell}$ are symmetric, only the upper (or lower)\ntriangular entries need to be considered. {The column vector\n\\begin{equation}\nK=[h^{(0)},\nh^{(1)}_0,\\ldots,h^{(1)}_{R-1}, h^{(2)}_{00},\\ldots,h^{(2)}_{R-1,R-1}]^T\n\\label{eqn:K}\n\\end{equation}\nconsists of all the Volterra kernels, where $k_m = h^{(2)}_{ab}$ for\n$R{+}1 \\leq m \\leq M{-}1$ and $(a,b)$\nis chosen as above.\n\nThe example of $R=2$, $L=3$, $N=L+R-1=4$, and $M=6$\nresults in (with $x_q=0$ for $q<0$)}\n\\begin{align}\n\\begin{bmatrix} y_0\\\\ y_1 \\\\y_2 \\\\ y_{3} \\end{bmatrix}\n&=\n\\begin{bmatrix}\n1 & x_0 & x_{-1} & x_{0}x_{0}&x_{0}x_{-1}&x_{-1}x_{-1} \\\\\n1 & x_1 & x_{0} & x_{1}x_{1}&x_{1}x_{0}&x_{0}x_{0}\\\\\n1 & x_2 & x_{1} & x_{2}x_{2}&x_{2}x_{1}&x_{1}x_{1} \\\\\n1 & x_3 & x_{2} & x_{3}x_{3}&x_{3}x_{2}&x_{2}x_{2} \\\\\n\\end{bmatrix} \\begin{bmatrix} h^{(0)} \\\\ h^{(1)}_0 \\\\ h^{(1)}_1 \\\\h^{(2)}_{00}\\\\ h^{(2)}_{01}\\\\h^{(2)}_{11}\n\\end{bmatrix}.\n\\end{align}\n\n\n\nFor the estimation of the distortions, we need to determine the values of $K$\nby solving the matrix equation \\eqref{eqn:3.5} with the method of least squares.\nWe assume now that the output data vector $Y$ has been measured in an experimental setup.\nWe can also concatenate multiple output pulses into a single vector to form $Y$,\nwhich allows us to perform the estimation using multiple short pulses with different\ncharacteristics as compared to a single long pulse. This provides the freedom of choosing\nthe format for our training data while observing experimental constraints.\nIn addition to taking a single long pulse or a set of short pulses,\nwe can also repeatedly use the same set of pulses to reduce the measurement error.\n\n\nAs the matrix $U$ contains higher-order terms of the input $x_n$, different\ncolumns of $U$ are highly correlated with each other. This leads to the problem\nof solving a linear regression model with a correlated basis set, i.e., the input variables are dependent\non each other. The precision of the estimation is adversely affected and less robust\nwhen naively applying the method of least squares to solve the matrix equation \\eqref{eqn:3.5}.\nWe resolve this problem by first orthogonalizing\nthe columns of the matrix $U$. The orthogonalization transforms the input variables (stacked in columns of $U$)\nsuch that they are independent of each other. After orthogonalizing $U$ to $V$,\nEq.~\\eqref{eqn:3.5} is transformed to\n\\begin{equation}\\label{eqn:3.6}\nY=VW.\n\\end{equation}\nNow we can solve the modified matrix equation \\eqref{eqn:3.6}\nusing the method of least squares to robustly obtain the values of the vector $W$.\nFinally, if the Gram-Schmidt method is\nused for orthogonalization, then one can convert $W$ to $K$ by recursive methods (as explained in \\cite{Korenberg:1988uv})\nto extract the Volterra kernels $h^{(0)}$, $h^{(1)}_{j}$, and $h^{(2)}_{k\\ell}$. In this work, we use the\nQR factorization method which directly provides the values for $K$ \\cite{HornJohnson:1985,GolubVanLoan:1996}.\n\n\\subsection{Gradient of the input response function\\label{sec:gradientdistortion}}\n\nAssuming that we have successfully estimated the transfer function,\nwe want to include this information in our gradient-based optimization.\nThis would allow us to also go beyond\nthe piecewise-constant control basis of GRAPE by\nincluding arbitrarily deformed controls, generalizing further along the\nlines of Ref.~\\cite{PhysRevA.84.022307}.\nWe provide now an analytic expression for the corresponding gradient\n(i.e.\\ Jacobian) to build upon the earlier work discussed in Appendix~\\ref{app:optim}.\n\nWe apply the commutativity of the convolution (i.e.\\ $f*g=g*f$), e.g., by\nchanging the integration variable from $\\tau$ to $z=t-\\tau$ in Eq.~\\eqref{eqn:2.5}.\nUsing a slight generalization, Eq.~\\eqref{eqn:3.2} can be rewritten\nas\\footnote{Note that using Eq.~\\eqref{eqn:3.7} for the estimation in\nSecs.~\\ref{sec:level3a}-\\ref{sec:level3b} would require\na number of coefficients given by $N\\times M$ instead of only $M$ and is therefore not recommended.}\n\n\\begin{equation}\ny_n = h^{(0)} + \\sum_{j=0}^{L-1} h^{(1)}_{n{-}j}\\, x_j\n+\\sum_{k,\\ell=0}^{L-1} h^{(2)}_{n{-}k,n{-}\\ell}\\, x_{k}\\, x_{\\ell},\n\\label{eqn:3.7}\n\\end{equation}\nwhere the upper summation bound $L{-}1$ differs from $R{-}1$ in Eq.~\\eqref{eqn:3.2},\ni.e.~integrating over the length of the input instead of the length of the kernel.\nFrom Eq.~\\eqref{eqn:3.7}, we specify for each time step (indexed by $n$) a\nscalar $K^{(0)}=h^{(0)}$, a column vector $K^{(1)}_n$ with\nentries $[K^{(1)}_n]_j=h^{(1)}_{n{-}j}$, and a\nmatrix $K^{(2)}_n$ with entries $[K^{(2)}_n]_{k\\ell}=h^{(2)}_{n{-}k,n{-}\\ell}$ for\n$j, k,\\ell \\in \\{0,\\ldots,L{-}1\\}$. With this notation, we can write Eq.~\\eqref{eqn:3.7}\nas a matrix equation\n\\begin{equation}\\label{eqn:3.8}\ny_n=K^{(0)} + {X}^T K^{(1)}_n+{X}^T K^{(2)}_n {X},\n\\end{equation}\nwhere the column vector $X=(x_0,\\ldots,x_{L-1})^T$ has length $L$.\nThe corresponding partial derivatives are\ngiven by\n\\begin{equation}\\label{eqn:3.9}\n\\frac{\\partial {y_n}}{\\partial {X}}= K^{(1)}_n+ {X}^T [K^{(2)}_n+(K^{(2)}_n)^T],\n\\end{equation}\nwhich simplifies for a symmetric quadratic kernel to\n\\begin{equation}\\label{eqn:3.10}\n\\frac{\\partial {y_n}}{\\partial {X}}= K^{(1)}_n+ 2{X}^T K^{(2)}_n\n\\end{equation}\nWe can calculate\n$\\partial {y_n}\/\\partial{X}$ for all $n$ and then determine the Jacobian.\nEventually, the gradient of the cost function \\eqref{eqn:2.3} is obtained\nusing the chain rule as, e.g., in \\cite{PhysRevA.84.022307} and as discussed in Appendix~\\ref{app:optim}.\n\n\\section{\\label{sec:level4}Non-linear distortions during Rydberg excitations}\n\n\n\nWe illustrate our scenario of non-linear distortions during controlled quantum dynamics with robust state-to-state\ntransfers in a single Rydberg atom experiment. In recent years, Rydberg\natoms have been proven to be a promising platform for quantum\nsimulation \\cite{Weimer:2010vq} and quantum\ncomputation \\cite{RevModPhys.82.2313}. One of the most distinctive features of\nthese atoms in quantum experiments is their strong and tunable\ndipole-dipole interactions \\cite{Jau:2016tg,Browaeys_2016}.\nFor larger Rydberg atom arrays as for quantum simulators, excitation\nprotocols (and more general operations)\nfrom the ground state to the Rydberg state are crucial.\nWe consider a gradient-based optimization of control pulses\n(without feedback) for tailored\nexcitation pulses as outlined in Sec.~\\ref{sec:level2}, (see also\nSec.~\\ref{sec:level6}\nand Appendix~\\ref{app:optim}).\n\n\\begin{figure}[t]\n\\includegraphics{energy}\n\\caption{(a) Path of the input control pulse\nfrom the computer code via an arbitrary waveform generator (AWG)\nand an acousto-optic modulator (AOM) to the atom. Before the atom,\nthe output control pulse can be measured using a photodiode (PD).\n(b) Three-level excitation scheme for a single Rydberg atom (see text).\\label{fig:4.2}}\n\\end{figure}\n\n\nThe Lindblad master equation for the time evolution of the system is given by Eq.~\\eqref{eqn:2.1}.\nFollowing \\cite{PhysRevA.97.053803}, the model Hamiltonian for a single\nRydberg atom is equal to\n\\begin{align}\nH(t) &= \\Omega_b(t) \\frac{\\ket{g}\\bra{p}+\\ket{p}\\bra{g}}{2}+\n\\Omega_r(t) \\frac{\\ket{p}\\bra{r}+\\ket{r}\\bra{p}}{2} \\nonumber \\\\\n&- \\Delta \\ket{p}\\bra{p} - \\delta \\ket{r}\\bra{r}.\n\\label{eqn:4.2}\n\\end{align}\nThe Rabi frequency $\\Omega_b(t)$ of the blue laser excites the atom from the\nground state $\\ket{g}$ to the intermediate state $\\ket{p}$ and the Rabi frequency $\\Omega_r(t)$\nexcites the atom from $\\ket{p}$ to the desired Rydberg state $\\ket{r}$ (see Fig.~\\ref{fig:4.2}(b)).\nIn terms of Eq.~\\eqref{eqn:2.2}, $\\Omega_b(t)$ and $\\Omega_r(t)$ constitute the\ntime-dependent control pulses. Moreover, $\\Delta$ and $\\delta$ are the single-photon and the\ntwo-photon resonance detuning, which will be for simplicity assumed to be zero\n($\\Delta=0$ MHz and $\\delta=0$ MHz).\nThe Lindblad operator \\cite{Lindblad:1976} reads as \\cite{PhysRevA.97.053803}\n\\begin{equation}\\label{eqn:4.3}\n{\\mathcal{L}(\\rho) =\\sum\\limits_{j\\in\\{d,g,g'\\}}\n(V_j \\rho V_j^{\\dagger}) -\\tfrac{1}{2}(V_j^{\\dagger}V_j \\rho + \\rho V_j^{\\dagger} V_j)}\n\\end{equation}\nwhere $V_g = \\sqrt{\\Gamma_{g}} \\ket{g}\\bra{p}$,\n$V_{g'}= \\sqrt{\\Gamma_{g'}} \\ket{g'}\\bra{p}$,\nand $V_d = \\sqrt{\\Gamma_{d}} \\ket{r}\\bra{r}$ are the Kraus operators.\nHere,\n$\\Gamma_{g}={\\Gamma}\/{3}$ and $\\Gamma_{g'}={2\\Gamma}\/{3}$\ndenote the probability for spontaneous emission from $\\ket{p}$ to the\nground state $\\ket{g}$ or to $\\ket{g'}$ which represents all other ground-state sublevels.\nRealistic experimental parameters $\\Gamma=2\\pi\\times 1.41$ MHz and\n$\\Gamma_d=2\\pi\\times 0.043$ MHz have been provided by the Browaeys group,\nwhere $\\Gamma_d$ is the Doppler effect.\nIn a real experiment, the gradients of the controls are restricted\ndue to bandwidth limitations.\nIn particular, the controls cannot have derivatives larger than a certain\nrise speed given by the experimental setup.\nIn our simulations, we take realistic values for the rise times of $0.1\\mu s$ and $0.15 \\mu s$ for\nthe red and blue laser pulses respectively (which translate into rise speeds).\n\n\n\nLet us now discuss how systematic distortions can be introduced in this experimental\nplatform during the processing and forwarding of the control signals which finally act\non the atom(s). The path of the control signals is sketched in Fig.~\\ref{fig:4.2}(a).\nStarting from some computer program, the input pulse (modulated with a fixed carrier\nfrequency) is passed through an arbitrary waveform generator (AWG) to produce\nthe radio-frequency pulse. This pulse is then used as an input for an\nacousto-optic modulator (AOM) which modulates the intensity of a laser beam.\nThe final laser pulse is then applied to the atom(s) to perform the excitation.\nThe AOM can shape pulses using optical effects such as\ndispersion \\cite{doi:10.1063\/1.2409868,doi:10.1063\/1.5020796}.\nIn this experimental setup, one can measure the laser signal before\nit acts on the atom(s) using a photo diode. In summary, one can choose the input\npulse and measure the output pulse; multiple measured input-output pulse\npairs serve as training data, which is used to determine systematic distortions.\n\nIn our simulation, we excite the Rydberg atom using the system\nHamiltonian from Eq.~\\eqref{eqn:4.2} by applying our optimized\ninput control pulses. After\nthat, we introduce quadratic distortions to the control pulses and\nrepeat the simulation. The discrete linear and quadratic distortions are\nprepared from Gaussian distributions described by\n\\begin{align}\\label{eqn:4.4}\nh_{1}(t) &= \\frac{1}{\\sigma\\sqrt{2\\pi}}\\exp[{-\\tfrac{(t{-}\\mu)^2}{2{\\sigma_{1}}^2}}], \\\\\n\\label{eqn:4.5}\nh_{2}(t_{1},t_{2}) & =\nJ\\exp[{-\\frac{(t_{1}{-}\\mu_{1})^2+(t_{2}{-}\\mu_{2})^2}{2{\\sigma_{2}}^2}}].\n\\end{align}\nThe memory length of the discretized dimensionless distortion is $R$.\nFor the distortions A, B, and C, we have chosen $R=50$,\nstandard deviations $\\sigma_1$\nof $1$, $6$, and $11$, and $\\sigma_2$ of $4.25$, $6.37$, and $8.50$.\nSimilarly, for the distortions D, E, and F, we have varied $R$ between\n$20$, $40$, and $60$ while fixing\n$\\sigma_{1}=1$ and $\\sigma_{2}=4.25$. The amplitude term $J$ has been\nkept constant at $5 \\times 10^{-6}$ in all cases.\nThe example distortion C is shown in Figs.~\\ref{fig:5.2}(a1) and \\ref{fig:5.2}(b1).\nThroughout this work, the zeroth order kernel is set to $h^{0}=0.1$.\n\n\n\nWe observe optimized controls with a simulated\nRydberg excitation error in the range from $0.06$ to $0.008$ for\ndifferent pulse lengths (see Fig.~\\ref{fig:4.1}). As expected, longer total durations\nfor the excitation lead to smaller simulated errors. But longer pulse durations\nmight lead to further decoherence effects in the experimental implementation\n(particularly when combined with additional experimental steps).\nWe, therefore, aim at reducing the length of the pulses\n(e.g.\\ to a pulse duration around $0.3\\mu s$) with reduced excitation errors.\nIn Fig.~\\ref{fig:4.1}(a), we notice a uniform increase in the error\nmagnitude when we increase the standard deviation of the Gaussian kernels of\nEqs.~\\eqref{eqn:4.4}-\\eqref{eqn:4.5} for the distortions A to C.\nThe standard deviation is kept constant in Fig.~\\ref{fig:4.1}(b), but we increase\nthe memory length for the distortions D to F which also results\nin bigger excitation error.\nThe increased excitation errors suggest that optimized control pulses\nwould be susceptible to distortions when applied in the Rydberg atom\nexperiments (and particularly for short pulse lengths). In Sec.~\\ref{sec:level5},\nwe present estimation results building on Sec.~\\ref{sec:level3}\nfor the considered types of distortions.\n\n\\begin{figure}\n\\includegraphics{excitation_error}\n\\caption{Reduced excitation efficiencies of optimized\ncontrol pulses\ndue to non-linear distortions in a simulated\nsingle Rydberg atom for distortions with\n(a) an increasing variance but constant memory length (A-C) and\n(b) a constant\nvariance but increasing memory length (D-F); refer to Sec.~\\ref{sec:level4}.\\label{fig:4.1}}\n\\end{figure}\n\n\n\\begin{figure*}\n\\includegraphics{kernels}\n\\caption{Estimation of both the linear and quadratic components for a non-linear distortion: (a) The linear component (a1) of the distortion C\nis compared with its estimated value (a2). The amplitude and\ntime steps are dimensionless. (a3) The mean absolute scaled error\n[as defined in Eq.~\\eqref{eqn:5.4}]\nbetween the actual and the estimated values\nis calculated for various types of distortions A-F [see Eqs.~\\eqref{eqn:4.4}--\\eqref{eqn:4.5}].\n(b) Quadratic component similar as in (a).\n\\label{fig:5.2}}\n\\end{figure*}\n\n\n\\section{\\label{sec:level5}Numerical estimation results}\n\n\n\nWe report in this section on different simulated estimation results\nwhich describe the characteristics and precision of applying\nthe Truncated Volterra series method while also comparing\nmultiple types of input control pulses used in the estimation.\nWe also perform the optimization for a single Rydberg\nexcitation again by including the distortions in the algorithm.\nIn each analysis, the estimated results are compared with the actual ones\nusing the mean absolute scaled error (MASE) measure\n\\begin{equation}\\label{eqn:5.4}\n\\text{MASE}=\\frac{1}{N}\\sum_{i=1}^{N}\n\\abs{\\frac{z^{\\text{true}}_i}{\\norm{z^{\\text{true}}}}-\\frac{z^{\\text{est}}_i}{\\norm{z^{\\text{est}}}}}\n\\end{equation}\nwhere $z^{\\text{true}}$ is the actual value, $z^{\\text{est}}$ is the estimated value, and $\\norm{z}$ is the\nFrobenius norm of the observable $z$ of length $N$. The\nMASE is numerically more stable compared to the mean relative error, which can be very large\nwhen the measured and the actual values are very small.\n\n\n\n\\subsection{\\label{subsec:level1}Estimation of distortions}\n\nWe start with the results presented in Fig.~\\ref{fig:5.2}\nwhere numerical distortions are estimated by relying on\na single randomly generated control pulse with $4000$ time steps. We apply\ndifferent distortions to the pulse and employ the resulting\ninput-output pulse pairs in the estimation. In order to provide\na more realistic analysis, we add an additional noise term to\nthe output pulse\n\\begin{equation}\\label{eqn:5.3}\ny_{\\text{noise}}=y_{\\text{output}}+ \\tfrac{1}{\\sigma\\sqrt{2\\pi}}\n\\exp[{-\\tfrac{(t{-}\\mu)^2}{2{\\sigma_{}}^2}}],\n\\end{equation}\nwhere the noise is drawn from a normal distribution\nwith mean $\\mu=0$ and standard deviation $\\sigma=10^{-4}$.\nFigures~\\ref{fig:5.2}(a1)-(b1) display the linear and quadratic\ncontribution of the distortion C. The corresponding\nestimated contributions are shown in Figs.~\\ref{fig:5.2}(a2)-(b2)\nwhich match closely with values in Figs.~\\ref{fig:5.2}(a1)-(b1).\nThe results also emphasize that precisely knowing\nthe memory length of the distortion (which is here $R=50$)\nis not required as redundant coefficients are automatically set to\nzero during the estimation for\na sufficiently large $R$ (here set to $60$).\nThe estimation process has been repeated for\nmultiple distortions of type A to F and we observe in Figs.~\\ref{fig:5.2}(a3)-(b3)\nlow estimation errors of approximately $10^{-7}$ to $10^{-8}$.\n\n\n\nWe now also compare the estimation method of Sec.~\\ref{sec:level3}\nwith a linear estimation method in the time domain which relies on\na linear impulse response [cf.\\ Eq.~\\eqref{eqn:2.5}]. We omit here\nthe very similar linear estimation in the frequency domain. We again\nuse the distortion types A to F from Sec.~\\ref{sec:level4} for this\ncomparison and apply them again to a random-noise pulse of $4000$ steps to obtain\ninput-output pulse pairs for the estimation.\nFigure~\\ref{fig:5.3}(a) shows the effect of the\ntrue and estimated distortion C when applied to\nan example pulse of $0.4$$\\mu s$ duration. The example pulse is\nstretched under the distortion to a final duration of $0.65$$\\mu s$.\nThe linear estimation is considerably less\nprecise when compared to the quadratic estimation. This effect\nis confirmed in Fig.~\\ref{fig:5.3}(b) which plots the estimation errors\nfor the different distortion types A-F. Naturally, this also validates that\nthe chosen distortion types contain some non-linearity which is not\naccounted for by a linear estimation.\n\n\\subsection{\\label{subsec:level2}Orthogonalization}\n\nOne important step of the estimation method is orthogonalization\nand we have discussed its significance in Sec.~\\ref{sec:level3}.\nTo further highlight the benefits of orthogonalizing the basis functionals,\nwe test the estimation by directly\nsolving the matrix equation\n\\begin{equation}\\label{eqn:5.1}\n{U^TY=U^TUK},\n\\end{equation}\nwhere $U$ is the matrix of the non-orthogonalized and correlated\nbasis functionals, $K$ is the to-be-estimated\nvector of linear\nor nonlinear kernel coefficients and $Y$ is the\nmeasured output vector.\nWe compare the results with coefficients we get from solving\nthe matrix equation \\eqref{eqn:3.6} with the orthogonalized basis set.\n\nIn this analysis, along with the benefit of orthogonalization, we also demonstrate how the estimation depends on\nthe number $M$ of the to-be-estimated coefficients for the distortion,\nthe amount of training data, and the presence of noise in the output pulse.\nFigures~\\ref{fig:5.4}(a)-(b) discuss the case without added noise.\nThe nonlinear distortion with $\\sigma_{1}=0.1, \\sigma_{2}= 0.42$ and $R=5$ is estimated\nusing spline input pulses as the training data.\nEach test and training pulse has 500 time steps and a unique frequency.\nFor a fixed number of spline pulses, we observe\nin Fig.~\\ref{fig:5.4}(a) an increasing estimation error for an\nincreasing number of coefficients $M$ (or memory length $R$\nas $M\\propto{R^2}$).\nFor each $M$, we apply the estimation results on\n$50$ different spline pulses which serve as test data.\nThe corresponding mean error is plotted as a line and\nthe $95\\%$ confidence interval is shown\nas a shaded region around the mean.\n\\begin{figure}\n\\includegraphics{with_without_ortho}\n\\caption{Comparison of the simulated estimation of\nnon-linear distortions without and with orthogonalization\nsolving respectively Eq.~\\eqref{eqn:5.1} and \\eqref{eqn:3.6}:\n(a) Using a fixed number of noiseless training data for spline\ninput pulses, the relative error rises with an increasing number of coefficients\n$M$. The plotted\nline shows the mean error and the shaded area\nindicates the spread between the $95\\%$ confidence interval found from applying\nthe estimation results to $50$ different test pulses.\nOrthogonalization is advantageous for\na larger number of coefficients. (b) Average estimation errors for different\ntraining data sets (see text) highlight the importance of increasing the frequency\ncontent of the available data. The averaging is performed over the full\nrange of all number of coefficients $M$ in (a).\n(c)-(d) As in (a)-(b), but the added noise in the output pulses of the data\nrequires a higher frequency content for comparable error rates.\n\\label{fig:5.4}}\n\\end{figure}\nIn Fig.~\\ref{fig:5.4}(b), we gradually increase the number of training\npulses used in the estimation.\nFor each fixed number of pulses, we perform the estimation on all the values of $M$\nas shown in Fig.~\\ref{fig:5.4}(a).\nHence each point in Fig.~\\ref{fig:5.4}(b) is averaged over $500$ results.\nIn all cases, the estimation\nbenefits from being performed with orthogonalization. Also, extending\nthe amount of training data points by adding more spline pulses with different\nfrequencies improves the estimation precision as seen in Fig.~\\ref{fig:5.4}(b).\nFor Figs.~\\ref{fig:5.4}(c)-(d) in the presence\nof a noise term in the output pulse with a standard deviation of $10^{-9}$, we\nobserve higher estimation errors which need to be compensated with\nadditional training data points. One can also reduce correlations\npresent in the training data by considering a random input pulse as its\nautocorrelation is zero. However, even a completely random input pulse\nresults in correlations in $U$ from Eqs.~\\eqref{eqn:3.5} and \\eqref{eqn:5.1}\nwhich contains various non-linear terms of the same input vector\n\\cite[p.~165]{Mathews2000}. In summary, Fig.~\\ref{fig:5.4} illustrates the positive\neffect of orthogonalization on the error rates in the estimation of the distortion.\n\n\n\\begin{figure}[t]\n\\includegraphics{bandwidth_comparison}\n\\caption{Simulated estimation errors of distortions with multiple types\nof data:\n(a) training data with more frequency content (such as random-noise pulses)\nperform better, even as the number of to-be-estimated coefficients $M$ increases.\n(b) A lower error can be achieved by increasing the frequency content of\nthe training data. For cosine and Gaussian pulses, the frequency content\nis increased by adding more pulses with different frequencies, whereas\nthe number of knots is increased within a single pulse for splines.\nSpectrally rich random-noise pulses are highly effective while keeping the data requirements low.\n(c)-(d) Similar to (a) and (b), but noisy training data\nincreases the overall error, while random-noise pulses are the most robust.\nThe estimation setup is similar to Fig.~\\ref{fig:5.4}.\n\\label{fig:5.5}}\n\\end{figure}\n\n\n\\subsection{\\label{subsec:level3}Frequency requirements}\n\n\n\n\n\n\nWe investigate different types of training data and their performance\nin the estimation following the setup of Fig.~\\ref{fig:5.5}. We can order\ndifferent training data types according to their increasing frequency content,\nwith Gaussian pulses having the minimum frequency and random-noise pulses\nhaving the maximum. Here, the frequency content describes the spectral content\nof the training data while its value depends on the type of pulses used\n(see Fig.~\\ref{fig:5.4}(b) and (d).\nThere are different errors for spline and cosine pulses depending on the amount of data.\nFor a fixed number of pulses, the estimation error grows with an\nincreasing number of coefficients $M$ [see Fig.~\\ref{fig:5.5}(a)]. Gaussian input pulses\nare most strongly affected by this, while this effect is essentially negligible\nin the case of random-noise pulses.\nThis illustrates the importance of spectrally rich input training pulses, which is further\nemphasized in Fig.~\\ref{fig:5.5}(b) where the estimation error is plotted, relative\nto the frequency content. For different types of input pulses, the frequency\ncontent is increased differently: we add more pulses with different standard\ndeviations for Gaussian pulses, we add more pulses with different frequencies\nfor cosine pulses, we add more\nrandom knots to a single spline. Since a random-noise pulse has a very large bandwidth,\nwe aim at increasing the frequency content by increasing the number of random-noise pulses which\nonly slightly reduces the estimation error.\nFigure~\\ref{fig:5.5}(b) highlights that the frequency content is crucial for the\nestimation and even a single random-noise pulse is highly effective due to its\nhigh-frequency content.\nSplines start to outperform the cosine pulses as soon as they\nattain higher frequency content than the latter.\nSimilar conclusions hold under noise as shown\nin Fig.~\\ref{fig:5.5}(c)-(d) while the overall estimation error increases\nfor the different input pulse types when compared to the noiseless case.\nThe data suggests that a high-frequency content in the training pulses\nmight prevent overfitting noise, which is important when\nworking with real experimental data. Also under noise,\nrandom-noise input pulses are most effective in the estimation due to their high frequency content.\n\n\n\n\\section{\\label{sec:level6}Application in optimal control}\nStarting from early developments in the field, various theoretical\nand experimental aspects of quantum control have been discussed\nin the recent review \\cite{Koch2022}.\nThe overall aim of quantum control is to shape a set of\nexternal field pulses that drive a quantum system and perform a given quantum process efficiently.\nWhile the analytical way of finding the\ncontrol parameters works for special cases, one can use\nhighly developed numerical tools in the context of\noptimal control theory. One solves the Schr\u00f6dinger or master equation iteratively and\nproduces pulse shapes that perform the desired time evolution.\nQuantum optimal control is broadly divided into at least the two categories of\nopen-loop and closed-loop. Open-loop\nmethods can be gradient-based or not.\nOpen-loop control is based on the available information about the\nHamiltonian of the system and hence it suffers when the system parameters are\nnot completely known such as in the case of an engineered quantum system (such as solid state systems) or when\nthe model cannot be solved precisely as in the case of many-body dynamics \\cite{PhysRevLett.109.153603}.\nThese limitations might be overcome by means of closed-loop optimal control where the control parameters are updated\nbased on the earlier measurements results \\cite{PhysRevA.84.022326,PhysRevLett.106.190501}.\nClosed-loop quantum optimal control can be implemented via\nboth gradient-based and gradient-free algorithms \\cite{PhysRevLett.118.150503, PhysRevLett.112.240504, PhysRevApplied.7.041001}.\nIn some cases, hybrid approaches have also been suggested \\cite{PhysRevLett.112.240503}.\nBut in the case where the system Hamiltonian is well known, open-loop control\nprovides more freedom to precisely tune the controls depending on experimental constraints and\ngenerally explore a wider range of control solutions.\nMoreover, it also gives a better understanding of the system and works well with systems where\nfast measurements are not feasible or very noisy, in contrast to closed-loop methods which may require many\nmeasurements to converge.\n\n\\begin{figure}\n\\includegraphics{excitation_error_tr}\n\\caption{Reduced excitation errors after correcting for the distortion with a\ngradient-based optimization relying on the trust-region method.\n(a) distortion C: significant reduced errors, (d) distortion F: mostly recovers\nthe ideal case.\n\\label{fig:5.6}}\n\\end{figure}\n\n\n\n\nTo take full advantage of the open-loop control method and to provide more robust pulses,\none can also characterize the experimental system completely or at least partially. Here, we highlight how\nthe estimation method from Sec.~\\ref{sec:level3} can be employed in an open-loop control setting\nto minimize the cost function $C$ in Eq.~\\eqref{eqn:A.2}\nby relying on the corresponding gradients as computed via\nEqs.~\\eqref{eqn:3.9} and \\eqref{eqn:A.9}.\nWe refer to Appendix~\\ref{app:optim} for details.\nThis compensates for distortions\nand decreases the error. Figure~\\ref{fig:5.6} shows test minimizations of the cost function\nusing the trust-region constrained algorithm \\cite{doi:10.1137\/1.9780898719857}, which\ncan perform constraint minimization with linear or non-linear constraints on the control pulses.\nTrust-region methods allow us to explicitly observe bandwidth limitations of the control hardware such\nas limited rise speeds as discussed in Sec.~\\ref{sec:level4} by enforcing the corresponding\npulse constraints.\nSince distortions C and F defined by Eqs.~\\eqref{eqn:4.4} and \\eqref{eqn:4.5}, have the strongest effects on the Rydberg excitation error (see Fig.~\\ref{fig:4.1}),\nwe correct the control pulses affected by them in the simulations.\nWe limit our test to pulses with shorter durations ranging from 0.1$\\mu s$ to 0.4$\\mu s$\nas they are less susceptible to decoherence and hence might be more suitable for the excitation process.\nWe compare the excitation error\nproduced from the corrected pulse with the ideal and the distorted pulse excitation error.\nIn particular, Figure~\\ref{fig:5.6} shows that the effect of the distortion C can be significantly\nreduced, but it cannot be completely corrected\ndue to a large standard deviation and long memory length in the distortion.\nThe distortion F has a small standard deviation combined with a\nlong memory length which still produces strong effects on the control pulse but\nwith a generally weaker distortion. In this case, the effect of the distortion can\nbe almost completely corrected.\n\n\n\n\nThe estimation of transfer functions in order to correct for distortions has one\nadditional benefit. The experimental hardware given by, e.g., AWGs and AOMs usually has\nbandwidth limitations which translate into limited rise speeds as discussed in\nSec.~\\ref{sec:level4}. In the process of characterizing the experimental devices\nvia their transfer function, we also estimate the effects of these bandwidth limitations.\nThe estimated transfer function is then applied during the optimization,\nwhich mirrors the effects in the experimental platform and implicitly enforces limitations\non the bandwidth or rise time.\nAssuming that the bandwidth-limiting effect of the estimated transfer function\nis pronounced enough,\nthis allows us to use the limited memory\nBroyden\u2013Fletcher\u2013Goldfarb\u2013Shanno (L-BFGS) algorithm to perform the\nminimization of the cost function \\cite{de2011second}. L-BFGS usually offers\na more efficient optimization but it cannot explicitly\naccount for general linear or non-linear constraints.\nIn the corresponding\noptimizations, we only enforce simple box constraints\nto limit the amplitude of the controls while using the extended L-BFGS or L-BFGS-B algorithm\n\\cite{53712fe04a3448cfb8598b14afab59b3}.\nThe results are shown in Fig.~\\ref{fig:5.7} where L-BFGS-B improves the excitation efficiency\nmore effectively than the trust-region method (which needs to also explicitly enforce\nthe constraints on the rise speeds).\nIn summary, combining the estimation of distortions with gradient-based optimizations\ncan often effectively compensate for these non-linear distortions during an open-loop\noptimization.\n\n\n\n\n\\section{\\label{sec:level7}Conclusion}\nWe have proposed a method for estimating\nnon-linear pulse distortions originating from\nexperimental hardware. Hardware limitations affect the performance\nof optimal control pulses as highlighted using numerical data for\nsingle Rydberg atom excitations. In this case, the errors are increased\nfor distorted control pulses beyond purely linear effects.\nWe provide a general model for\ndescribing the complex characteristics of these non-linear effects.\nTo incorporate estimated distortions into\nopen-loop optimizations, we have detailed\na formula to determine\nthe Jacobian of the transfer function.\n\n\\begin{figure}\n\\includegraphics{excitation_error_bfgs}\n\\caption{Reduced excitation error for L-BFGS-B when\ncompared to the trust-region method, even though\nL-BFGS-B does not explicitly enforce constraints on\nthe control pulses (such as limited rise speeds).\nBut the correctly estimated transfer function\nwill implicitly account for pulse constraints.\nAlso, L-BFGS-B is more effective\nin the optimization.\n\\label{fig:5.7}}\n\\end{figure}\n\n\nWe tested and validated our\nproposed method by\nefficiently estimating different numerical quadratic distortions with varying strength\nand duration. We have also shown that linear estimation methods\ncannot effectively handle non-linear transfer functions. From our detailed analysis and tests,\nwe deduce that the orthogonalization (as described\nin Sec.~\\ref{subsec:level2}) is key for a robust estimation.\nA robust least-squares estimation is effective only after\nthe orthogonalization is applied to the matrix containing the training data\nas its correlated columns would otherwise interfere with the estimation.\nAnother critical requirement for effectively performing\nthe estimation is training data with enough frequency content.\nLarge frequency content such as in random-noise pulses better\ncaptures the non-linear features of transfer functions, particularly in the\npresence of measurement noise.\n\nSince the estimation method is independent\nof any particular type of device characteristics, it can easily be adapted to a\nwide range of experimental platforms.\nCombining our estimation method with\nexisting numerical optimization techniques can improve the quality and\nrobustness of quantum operations. Our work thereby addresses a key challenge\nof enhancing the accuracy and robustness\nof experimental quantum technology platforms.\n\n\\begin{acknowledgments}\nThe authors acknowledge funding from the EU H2020-FETFLAG-03-2018 under\ngrant agreement No 817482 (PASQuanS). We also appreciate support from the German Federal Ministry of Education\nand Research through the funding program quantum technologies---from basic research to market\nunder the project FermiQP, 13N15891. We would like to thank\nAntoine Browaeys, Daniel Barredo, Thierry Lahaye, Pascal Scholl, and Hannah Williams for\nthe illuminating discussions about the Rydberg system as well as providing detailed\nexperimental parameters. R.Z. would like to thank\nJian Cui for initial discussions about the Rydberg setup.\n\\end{acknowledgments}\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\\section{Introduction}\n\\lblsec{intro}\n\n\n\n\n\n\n\\begin{figure}\n \\centering\n \\vspace{-0.5cm}%\n \\includegraphics[scale=0.58, trim=230 106 150 20, clip]{figures\/figure_2_dataset_distillation.pdf}\\vspace{-0.85cm}\n \\caption{Dataset distillation aims to generate a small synthetic dataset for which a model trained on it can achieve a similar test performance as a model trained on the whole real train set.}%\n \\lblfig{dataset_distillation}\\vspace{-5pt}\n\\end{figure}\n\nIn the seminal 2015 paper, Hinton et al.~\\cite{hinton2015distilling} proposed {\\em model distillation}, which aims to distill the knowledge of a complex model into a simpler one. \n{\\em Dataset distillation}, proposed by Wang et al.~\\cite{dd}, is a related but orthogonal task: rather\nthan distilling the model, the idea is to distill the dataset. \nAs shown in Figure~\\ref{fig:dataset_distillation}, the goal is to distill the knowledge from a large training dataset into a very small set of synthetic training images (as low as one image per class) such that training a model on the distilled data would give a similar test performance as training one on the original dataset. Dataset distillation has become a lively research topic in machine learning~\\cite{bohdal2020flexible,sucholutsky2021soft,dc,dsa,nguyen2020dataset,nguyen2021dataset,dm} with various applications, such as continual learning, neural architecture search, and privacy-preserving ML. Still, the problem has so far been of mainly theoretical interest, since most prior methods focus on toy datasets, like MNIST and \\mbox{CIFAR}, while struggling on real, higher-resolution images. In this work, we present a new approach to dataset distillation that not only outperforms previous work in performance, but is also applicable to large-scale datasets, as shown in \\reffig{teaser}.\n\n\nUnlike classical data compression, dataset distillation aims for a small synthetic dataset that still retains adequate task-related information so that models trained on it can generalize to unseen test data, as shown in \\reffig{dataset_distillation}. Thus, the distilling algorithm must strike a delicate balance by heavily compressing information without completely obliterating the discriminative features. \nTo do this, dataset distillation methods attempt to discover exactly which aspects of the real data are critical for learning said discrimination. Several methods consider end-to-end training~\\cite{dd,nguyen2020dataset,nguyen2021dataset} but often require huge compute and memory and suffer from inexact relaxation~\\cite{nguyen2020dataset,nguyen2021dataset} or training instability of unrolling many iterations~\\cite{dd,maclaurin2015gradient}. To reduce the optimization difficulty, other methods~\\cite{dc,dsa} focus on short-range behavior, enforcing a single training step on distilled data to match that on real data. However, error may accumulate in evaluation, where the distilled data is applied over many steps. We confirm this hypothesis experimentally in \\refsec{shortlongrange}.\n\n\nTo address the above challenges, we sought to directly imitate the long-range training dynamics of networks trained on real datasets. In particular, we match segments of parameter trajectories trained on synthetic data with segments of pre-recorded trajectories from models trained on real data and thus avoid being short-sighted (\\ie focusing on single steps) or difficult to optimize (\\ie modeling the full trajectories). \nTreating the real dataset as the gold standard for guiding the network's training dynamics, we can consider the induced sequence of network parameters to be an \\emph{expert trajectory}. If our distilled dataset were to induce a network's training dynamics to follow these expert trajectories, then the synthetically trained network would land at a place close to the model trained on real data (in the parameter space) and achieve similar test performance. %\n\n\\input{figText\/overview}\n\n\nIn our method, our loss function \\textit{directly} encourages the distilled dataset to guide the network optimization along a similar trajectory (\\Cref{fig:method_traj_match}). We first train a set of models from scratch on the real dataset and record their expert training trajectories. We then initialize a new model with a random time step from a randomly chosen expert trajectory and train for several iterations on the \\textit{synthetic} dataset. Finally, we penalize the distilled data based on how far this synthetically trained network deviated from the expert trajectory and back-propagate through the training iterations. Essentially, we transfer the knowledge from many expert training trajectories to the distilled images. \n\nExtensive experiments show that our method handily outperforms existing dataset distillation methods as well as coreset selection methods on standard datasets, including CIFAR-10, CIFAR-100, and Tiny ImageNet. For example, we achieve 46.3\\% with a \\emph{single} image per class and 71.5\\% with 50 images per class on CIFAR-10, compared to the previous state of the art (28.8\\% \/ 63.0\\% from \\cite{dsa, dm} and 36.1\\% \/ 46.5\\% from \\cite{nguyen2021dataset}).\nFurthermore, our method also generalizes well to larger data, allowing us to see high $128\\times128$-resolution images distilled from ImageNet \\cite{deng2009imagenet} for the first time. Finally, we analyze our method through additional ablation studies and visualizations. %\n\\href{https:\/\/github.com\/GeorgeCazenavette\/mtt-distillation}{Code} and models are also available on our \\href{https:\/\/georgecazenavette.github.io\/mtt-distillation\/}{webpage}.\n\\lblfig{intro}\n\\section{Related Work}\n\n\n\n\\myparagraph{Dataset Distillation.}\nDataset distillation was first introduced by Wang et al.~\\cite{dd}, who proposed expressing the model weights as a function of distilled images and optimized them using gradient-based hyperparameter optimization~\\cite{maclaurin2015gradient}, which is also widely used in meta-learning research \\citep{finn2017model,nichol2018first}. %\nSubsequently, several works significantly improved the results by learning soft labels \\cite{bohdal2020flexible,sucholutsky2021soft}, amplifying learning signal via gradient matching \\cite{dc}, adopting augmentations~\\cite{dsa}, and optimizing with respect to the infinite-width kernel limit \\cite{nguyen2020dataset,nguyen2021dataset}. Dataset distillation has enabled various applications including continual learning~\\cite{dd,dc,dsa}, efficient neural architecture search~\\cite{dc,dsa}, federated learning~\\cite{goetz2020federated,zhou2020distilled,sucholutsky2020secdd}, and privacy-preserving ML~\\cite{sucholutsky2020secdd,li2020soft} for images, text, and medical imaging data. As mentioned in the introduction, our method does not rely on single-step behavior matching~\\cite{dc,dsa}, costly unrolling of full optimization trajectories~\\cite{dd,sucholutsky2021soft}, or large-scale Neural Tangent Kernel computation~\\cite{nguyen2020dataset,nguyen2021dataset}. Instead, our method achieves long-range trajectory matching by transferring the knowledge from pre-trained experts. %\n\n\nConcurrent with our work, the method of Zhao and Bilen~\\cite{dm} completely disregards optimization steps, instead focusing on distribution matching between synthetic and real data. While this method is applicable to higher-resolution datasets (\\eg Tiny ImageNet) due to reduced memory requirements, it attains inferior performance in most cases (\\eg when compared to previous works~\\cite{dc,dsa}). In contrast, our method simultaneously reduces memory costs while outperforming existing works~\\cite{dc,dsa} and the concurrent method~\\cite{dm} on both standard benchmarks and higher-resolution datasets. \n\nA related line of research learns a generative model to synthesize training data \\cite{such2020generative,masarczyk2020reducing}. However, such methods do not generate a small-size dataset, and are thus not directly comparable with dataset distillation methods.\n\n\n\n\\myparagraph{Imitation Learning.} \nImitation learning attempts to learn a good policy by observing a collection of expert demonstrations \\cite{osa2018algorithmic,peng2018deepmimic,peng2021amp}. Behavior cloning trains the learning policy to act the same as expert demonstrations. Some more sophisticated formulations involve on-policy learning with labeling from the expert \\cite{ross2011reduction}, while other approaches avoid any label at all, \\eg via distribution matching\n\\cite{ho2016generative}. Such methods (behavior cloning in particular) have been shown to work well in offline settings \\cite{fu2020d4rl,gulcehre2020rl}. \nOur method can be viewed as imitating a collection of expert network training trajectories, which are obtained via training on real datasets. Therefore, it can be considered as doing imitation learning over optimization trajectories.\n\n\\myparagraph{Coreset and Instance Selection.}\nSimilar to dataset distillation, coreset~\\cite{tsang2005core,har2007smaller,bachem2017practical,sener2018active,chen2010super} and instance selection~\\cite{olvera2010review} aim to select a subset of the entire training dataset, where training on this small subset achieves good performance. Most of such methods do not apply to modern deep learning, but new formulations based on bi-level optimization have shown promising results on applications like continual learning \\cite{borsos2020coresets}. Related to coreset, other lines of research aim to understand which training samples are ``valuable'' for modern machine learning, including measuring single-example accuracy \\citep{lapedriza2013all} and counting misclassification rates \\citep{toneva2018empirical}. In fact, dataset distillation is a generalization of such ideas, as the distilled data do not need to be realistic or come from the training set.\n\n\n\n\n\n\n\n\\section{Method}\n\\lblsec{method}\n\\emph{Dataset Distillation} refers to the curation of a small, synthetic training set $\\mathcal{D}_\\mathsf{syn}$ such that a model trained on this synthetic data will have similar performance on the real test set as a model trained on the large, real training set $\\mathcal{D}_\\mathsf{real}$. In this section, we describe our method that directly mimics the long-range behavior of real-data training, matching multiple training steps on distilled data to \\textit{many} more steps on the real data.\n\n\n\nIn \\refsec{expert}, we discuss how we obtain expert trajectories of networks trained on real datasets. In \\refsec{matching}, we describe a new dataset distillation method that explicitly encourages the distilled dataset to induce similar long-range network parameter trajectories as the real dataset, resulting in a synthetically-trained network that performs similarly to a network trained on real data. Finally, \\refsec{memory} describes our techniques to reduce memory consumption.\n\n\n\\subsection{Expert Trajectories}\n\\lblsec{expert}\nThe core of our method involves using \\emph{expert trajectories} $\\tau^*$ to guide the distillation of our synthetic dataset. By expert trajectories, we mean the time sequence of parameters $\\{\\theta^*_t\\}_0^T$ obtained during the training of a neural network on the \\emph{full, real dataset}. To generate these expert trajectories, we simply train a large number of networks on the real dataset and save their snapshot parameters at every epoch. We call these sequences of parameters ``expert trajectories'' because they represent the theoretical upper bound for the dataset distillation task: the performance of a network trained on the full, real dataset. Similarly, we define student parameters $\\hat{\\theta}_t$ as the network parameters trained on synthetic images at the training step $t$. Our goal is to distill a dataset that will induce a similar trajectory (given the same starting point) as that induced by the real training set such that we end up with a similar model.\n\nSince these expert trajectories are computed using only real data, we can pre-compute them before distillation. All of our experiments for a given dataset were performed using the same pre-computed set of expert trajectories, allowing for rapid distillation and experimentation.\n\n\\input{sections\/algorithm}\n\\subsection{Long-Range Parameter Matching}\n\\lblsec{matching}\nOur distillation process learns from the generated sequences of parameters making up our expert trajectories $\\{\\theta^*_t\\}_0^T$. Unlike previous work, our method directly encourages the long-range training dynamics induced by our synthetic dataset to match those of networks trained on the real data.\n\nAt each distillation step, we first sample parameters from one of our expert trajectories at a random timestep $\\theta^*_t$ and use these to initialize our student parameters $\\hat{\\theta}_t \\coloneqq \\theta^*_t$. Placing an upper bound $T^+$ on $t$ lets us ignore the less informative later parts of the expert trajectories where the parameters do not change much.\n\nWith our student network initialized, we then perform $N$ gradient descent updates on the student parameters with respect to the classification loss of the \\emph{synthetic} data:\n\\begin{equation}\n \\hat{\\theta}_{t+n+1} = \\hat{\\theta}_{t+n} - \\alpha \\nabla \\ell(\\mathcal{A}(\\mathcal{D}_\\mathsf{syn}); \\hat{\\theta}_{t+n}),\n\\end{equation}\nwhere $\\mathcal{A}$ is the differentiable augmentation technique~\\cite{stylegan2ada,zhao2020image,tran2020towards,diffaug} used in previous work~\\cite{dsa}, and $\\alpha$ is the (trainable) learning rate used to update the student network. Any data augmentation used during distillation must be differentiable so that we can back-propagate through the augmentation layer to our synthetic data. Our method does not use differentiable \\textit{Siamese} augmentation since there is no real data used during the distillation process; we are only applying the augmentations to synthetic data at this time. However, we do use the same types of differentiable augmentations on real data during the generation of the expert trajectories.\n\nFrom this point, we return to our expert trajectory and retrieve the expert parameters from $M$ training updates after those used to initialize the student network $\\theta^*_{t+M}$. Finally, we update our distilled images according to the weight matching loss: i.e., the normalized squared $L_2$ error between the updated student parameters $\\hat{\\theta}_{t+N}$ and the known future expert parameters $\\theta^*_{t+M}$:\n\\begin{equation}\n \\mathcal{L} = \\frac{\\|\\hat{\\theta}_{t+N} - \\theta^*_{t+M}\\|^2_2}{\\|\\theta^*_{t} - \\theta^*_{t+M}\\|_2^2}, %\n \\lbleq{weight_matching}\n\\end{equation}%\nwhere we normalize the $L_2$ error by the expert distance traveled so that we still get a strong signal from later training epochs where the expert does not move as much. This normalization also helps self-calibrate the magnitude difference across neurons and layers. We have also experimented with other choices of loss functions such as a cosine distance, but find our simple $L_2$ loss works better empirically. We also tried to match the network's output logits between expert trajectory and student network but did not see a clear improvement. We speculate that backpropagating from the network output to the weights introduces additional optimization difficulty. %\n\n\nWe then minimize this objective to update the pixels of our distilled dataset, along with our trainable learning rate $\\alpha$, by back-propagating through all $N$ updates to the student network. The optimization of trainable learning rate $\\alpha$ serves as automatic adjusting for the number of student and expert updates (hyperparameters $M$ and $N$). We use SGD with momentum to optimize $\\mathcal{D}_\\mathsf{syn}$ and $\\alpha$ with respect to the above objective. \\refalg{main} illustrates our main algorithm. \n\\input{figText\/table_sota_new}\n\n\\subsection{Memory Constraints}\n\\lblsec{memory}\nGiven that we are back-propagating through many gradient descent updates, memory consumption quickly becomes an issue when our distilled dataset is sufficiently large, as we have to jointly optimize all the images of all the classes at each optimization step. To reduce memory consumption and ease the learning problem, previous methods distill one class at a time~\\cite{dc, dsa, dm}, but this may not be an ideal strategy for our method since the expert trajectories are trained on all classes simultaneously.\n\nWe could potentially circumvent this memory constraint by sampling a new mini-batch at every distillation step (the outer loop in Line \\ref{step:outer} of \\refalg{main}). \nUnfortunately, this comes with its own issues, as redundant information could be distilled into multiple images across the synthetic dataset, degrading to catastrophic mode collapse in the worst case.\n\nInstead, we can sample a new mini-batch $b$ for every update of the \\emph{student network} (i.e., the inner loop in Line \\ref{step:inner} of \\refalg{main}) such that all distilled images will have been seen by the time the final weight matching loss (\\refeq{weight_matching}) is calculated. The mini-batch $b$ still contains images from different classes but has much fewer images per class. In this case, our student network update then becomes%\n\\vspace{-5pt}\\begin{equation}\n\\begin{split}\n & b_{t+n} \\sim \\mathcal{D}_\\mathsf{syn}\\\\\n & \\hat{\\theta}_{t+n+1} =\\hat{\\theta}_{t+n} - \\alpha \\nabla \\ell(\\mathcal{A}(b_{t+n}); \\hat{\\theta}_{t+n}).\n\\end{split}\\vspace{-10pt}\n\\end{equation}%\nThis method of batching allows us to distill a much larger synthetic dataset while ensuring some amount of heterogeneity among the distilled images of the same class. \n\n\n\n\n\n\n\\section{Experiments}\n\\lblsec{expr}\nWe evaluate our method on various datasets, including \\begin{itemize}[topsep=1pt, partopsep=9pt, itemsep=-1pt, parsep=0.5ex]\n \\item $32\\times32$ CIFAR-10 and CIFAR-100 (\\refsec{low-res}), two commonly used datasets in dataset distillation literature,\n \\item $64\\times 64$ Tiny ImageNet (\\refsec{tiny}), a recent benchmark by the concurrent work~\\cite{dm}, and \n \\item our new $128 \\times 128$ ImageNet subsets (\\refsec{imagenet}).\n\\end{itemize}\nWe provide additional visualizations and ablation studies in the supplementary material. \n\n\n\\myparagraph{Evaluation and Baselines.} %\nWe evaluate various methods according to the standard protocol: training a randomly initialized neural network from scratch on \\emph{distilled} data and evaluating on the validation set. %\n\nTo generate the distilled images for our method, we employ the distillation process detailed in the previous section and Algorithm \\ref{alg:alg}, using the same suite of differentiable augmentations as done in previous work~\\cite{dsa,dm}. The hyperparameters used for each setting (real epochs per iteration, synthetic updates per iteration, image learning rate, etc.) can be found in the supplemental material. \n\n\n\n\n\nWe compare to several recent methods including Dataset Distillation~\\cite{dd} (\\texttt{DD}), Flexible Dataset Distillation ~\\cite{bohdal2020flexible} (\\texttt{LD}), Dataset Condensation~\\cite{dc} (\\texttt{DC}), and Differentiable Siamese Augmentation~\\cite{dsa} (\\texttt{DSA}), along with a method based on the infnite-width kernel limit~\\citep{nguyen2020dataset,nguyen2021dataset} (\\texttt{KIP}) and concurrent works Distribution Matching~\\cite{dm} (\\texttt{DM}) and Aligning Features~\\cite{wang2022cafe} (\\texttt{CAFE}). We also compare our methods with instance selection algorithms including random selection (\\texttt{random}), herding methods~\\cite{chen2010super} (\\texttt{herding}), and example forgetting~\\cite{toneva2018empirical} (\\texttt{forgetting}). \n\n\\myparagraph{Network Architectures.}\nStaying with precedent~\\cite{dc, dsa, dm, nguyen2021dataset}, we mainly employ a simple ConvNet architecture designed by Gidaris and Komodakis~\\cite{gidaris2018dynamic} for our distillation tasks. The architecture consists of several convolutional blocks, each containing a\n$3\\times 3$ convolution layer with 128 filters, Instance normalization~\\cite{ulyanov2016instance}, RELU, and $2\\times2$ average pooling with stride 2. After the convoluation blocks, a single linear layer produces the logits. The exact number of such blocks is decided by the dataset resolution and is specified below for each dataset.\nStaying with this simple architecture allows us to directly analyze the effectiveness of our core method and remain comparable with previous works.\n\n\\subsection{Low-Resolution Data (32$\\times$32)}\n\\lblsec{low-res}\n\nFor low-resolution tasks, we begin with the 32$\\times$32 CIFAR-10 and CIFAR-100 datasets~\\cite{CIFAR10}. For these datasets, we employ ZCA whitening as done in previous work~\\cite{nguyen2020dataset, nguyen2021dataset}, using the Kornia \\cite{kornia} implementation with default parameters.\nStaying with precedent, we use a depth-3 ConvNet taken directly from the open-source code~\\cite{dc, dsa}.\n\nAs seen in \\reftbl{sota}, our method significantly outperforms all baselines in every setting. In fact, on the one image per class setting, we improve the next best method (\\texttt{DSA} \\cite{dsa}) to almost twice test accuracy, on both datasets. For CIFAR-10, these distilled images can be seen in \\reffig{CIFAR-10}. CIFAR-100 images are visualized in the supplementary material\n\nIn \\reftbl{transfer_nn}, we also compare with a recent method \\texttt{KIP}~\\cite{nguyen2020dataset,nguyen2021dataset}, where the distilled data is learned with respect to the Neural Tangent Kernel. Because \\texttt{KIP} training is agnostic to actual network width, we test their result on both a ConvNet of the same width as us and other methods (128) and a ConvNet of larger width (1024) (which is shown in \\texttt{KIP} paper \\citep{nguyen2021dataset}). Based on the the infinite-width network limit, \\texttt{KIP} may exhibit a gap with practical finite-width networks. Our method does not suffer from this limitation and generally achieves better performance. In all settings, our method, trained on the 128-width network, outperforms \\texttt{KIP} results evaluated on both widths, except for just one setting where \\texttt{KIP} is applied on the much wider 1024-width network. \n\nAs noted in the previous methods \\citep{dd}, we also see significant diminishing returns when allowing more images in our synthetic datasets. For instance, on CIFAR-10, we see an increase from 46.3\\% to 65.3\\% classification accuracy when increasing from 1 to 10 images per class, but only an increase from 65.3\\% to 71.5\\% when increasing the number of distilled images per class from 10 to 50.\n\nIf we look at the one image per class visualizations in \\reffig{CIFAR-10} (top), we see very abstract, yet still recognizable, representations of each class. When we limit the task to just one synthetic image per class, the optimization is forced to squeeze as much of the class's distinguishing information as possible into just one sample. When we allow more images in which to disperse the class's information, the optimization has the freedom to spread the class's discriminative features among the multiple samples, resulting in a diverse set of structured images we see in \\reffig{CIFAR-10} (bottom) (e.g., different types of cars and horses with different poses).\n\\input{figText\/kip_to_nn}\n\\input{figText\/crossarch}\n\n\\myparagraph{Cross-Architecture Generalization.}\nWe also evaluate how well our synthetic data performs on architectures different from the one used to distill it on the CIFAR-10, 1 image per class task. In \\reftbl{cross}, we show our baseline \\linebreak ConvNet performance and evaluate on ResNet~\\cite{resnet}, VGG~\\cite{vgg}, and AlexNet~\\cite{alexnet}. \n \nFor \\texttt{KIP}, instead of the Kornia \\cite{kornia} ZCA implementation, we use the authors' custom ZCA implementation for evaluation of their method. Our method is solidly the top performer on all the transfer models except for AlexNet where we lie within one standard deviation of \\texttt{DSA}. This could be attributed to our higher baseline performance, but it still shows that our method is robust to changes in architecture.\n\n\\begin{figure}\n \\centering\n \\vspace{-6pt}\n \\includegraphics[width=0.88\\linewidth]{figures\/CIFAR10_1_ipc.pdf}\\\\[-0.6ex]\n {\\smaller 1 image per class} \\vspace{0.5cm}\\\\\n \\vspace{-0.34cm}\n\n \\includegraphics[width=0.88\\linewidth, trim=0 0 0 20, clip]{figures\/CIFAR10.pdf}\\\\[-0.6ex]\n {\\smaller 10 images per class}\n \\vspace{-7pt}\n \\caption{CIFAR-10: The 1 image per class images are more abstract but also more information-dense while the 10 images per class images are more expressive and contain more structure.}\n \\lblfig{CIFAR-10}\n \\vspace{-7pt}\n\\end{figure}\n\n\n\\subsection{Short-Range vs.~Long-Range Matching}\n\\lblsec{shortlongrange}\n\nUnlike some prior works (\\texttt{DC} and \\texttt{DSA}), our method performs long-range parameter matching, where $N$ training steps on distilled data match a much larger $M$ steps on real data. Methods that optimize over entire training processes (\\eg \\texttt{DD} and \\texttt{KIP}) can be viewed as even longer range matching. However, their performances fall short of our method (\\eg in \\reftbl{sota}), likely due to related instabilities or inexact approximations. Here, we experimentally confirm our hypothesis that long-range matching achieved by larger $M$ and $N$ in our method is superior to the short-range counterparts (such as small $M$ and $N$ and \\texttt{DSA}).\n\n\nIn \\reffig{mn-alpha} (left), we evaluate our method on different settings of $M$ and $N$. Really short-range matching (with $N=1$ and small $M$) generally exhibits worse performance than long-range matching, with the best performance attained when both $N$ and $M$ are relatively large. Furthermore, as we increase $N$, the power of $N$ combined steps (on distilled data) becomes stronger and can approximate longer-range behavior, leading to the optimal $M$ values shifting to greater values correspondingly.\n\n\\input{figText\/ablation_plots2}\n\nIn \\reffig{mn-alpha} (right), we evaluate our method and a short-range matching work (\\texttt{DSA}) on their abilities to approximate real training behavior over short and long ranges. Starting from a set of initial parameters, we set the target parameters to be the result of $\\Delta_t$ training steps on real data (\\ie the long-range behavior that distilled data should mimic). A small (or large) $\\Delta_t$ means evaluating matching over a short (or long) range. For both methods, we test how close they can train the network (using distilled data) from the same initial parameters to the target parameters. \\texttt{DSA} is only optimized to match short-range behavior, and thus errors may accumulate during longer training. Indeed, as $\\Delta_t$ grows larger, \\texttt{DSA} fails to mimic the real data behavior over longer ranges. In comparison, our method is optimized for long-range matching and thus performs much better.\n\n\n\n\n\n\\input{figText\/imagenet_table}\n\n\n\\subsection{Tiny ImageNet (64$\\times$64)}\n\\lblsec{tiny}\n\nIntroduced to the dataset distillation task by the concurrent work, Distribution Matching (\\texttt{DM})~\\cite{dm}, we also show the effectiveness of our algorithm on the 200 class, 64$\\times$64 Tiny ImageNet~\\cite{tiny} dataset (a downscaled subset of ImageNet \\cite{deng2009imagenet}). %\nTo account for the higher image resolution, we move up to a depth-4 ConvNet, similar to \\texttt{DM}~\\cite{dm}.\n\nMost dataset distillation methods (other than \\texttt{DM}) are unable to handle this larger resolution due to extensive memory or time requirement, as the \\texttt{DM} authors also observed \\citep{dm}. In \\reftbl{sota}, our method consistently outperforms the only viable such baseline, \\texttt{DM}. Notably, on the 10 images per class task, our method improves the concurrent work \\texttt{DM} from 12.9\\% and 23.2\\%. A subset of our results is shown in \\reffig{tinyimagenet}. The supplementary material contains the rest of the images. %\n\n\nAt 200 classes and 64$\\times$64 resolution, Tiny ImageNet certainly poses a much harder task than previous datasets. Despite this, many of our distilled images are still recognizable, with a clear color, texture, or shape pattern.\n\\begin{figure}[t]\n \\centering\n \\includegraphics[width=\\linewidth]{figures\/Tiny.pdf}\n \\vspace{-7pt}\n \\caption{Selected samples distilled from Tiny ImageNet, one image per class. Despite the higher resolution, our method still produces high-fidelity images. (Can you guess which classes these images represent? Check your answers in the footnote!\\protect\\footnotemark)}\n \\lblfig{tinyimagenet}%\n \\vspace{-7pt}\n\\end{figure}\n\n\\footnotetext{\\smaller{Answers for \\reffig{tinyimagenet}: \\textbf{First Row:} African Elephant, Jellyfish, Kimono, Lampshade, Monarch} \\textbf{Second Row:} Organ, Pizza, Pretzel, Teapot, Teddy}\n\\subsection{ImageNet Subsets (128$\\times$128)}\n\\lblsec{imagenet}\n\\input{figText\/imagenet_fig}\nNext, we push the boundaries of dataset distillation even further by running our method on yet higher resolution images in the form of 128$\\times$128 subsets of ImageNet~\\cite{deng2009imagenet}. Again, due to the higher resolution, we increase the depth of our architecture and use a depth-5 ConvNet for the 128$\\times$128 ImageNet subsets.\n\n\nImageNette (assorted objects) and ImageWoof (dog breeds) are existing subsets~\\cite{imagenette} designed to be easy and hard to learn respectively. We also introduce ImageFruit (fruits), ImageMeow (cats), ImageSquawk (birds), and ImageYellow (yellow-ish things) to further illustrate our algorithm. %\n\nSimilar to Tiny ImageNet, most dataset distillation baselines do not scale up to our ImageNet subset settings.\nAs the code of \\texttt{DM}~\\cite{dm} is not publicly available now, we choose to only compare to the networks trained on the full dataset. We wish to show that our method transfers well to large images and still produces meaningful results at a higher resolution. Validation set accuracies are presented in \\reftbl{imagenet}.\n\nWhile all of the generated images are devoid of high-frequency noise, the tasks still differ in the type of distilled image they induce. For tasks where all the classes have a similar structure but unique textures like ImageSquawk (\\reffig{imagenet}), the distilled images may not have much structure but instead store discriminating information in the textures.\n\nConversely, for tasks where all classes have similar color or textures like ImageYellow (\\reffig{imagenet}), the distilled images seem to diverge from their common trait and accentuate the structure or secondary color that makes them unique. Specifically, note the differences between the distilled ``Banana'' images for the ImageFruit and ImageYellow (bottom row, \\reffig{imagenet}). Although the expert trajectory-generating networks saw the same ``Banana'' training images, the distilled images differ between the two tasks. The distilled ``Banana'' for the ImageYellow task is clearly much ``greener'' than the equivalent image for the ImageFruit task. This implies that the expert networks identify different features by which to identify classes based on the other classes in the task.\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\n\\section{Discussion and Limitations}\n\\lblsec{discussion}\nIn this work, we introduced a dataset distillation algorithm by means of directly optimizing the synthetic data to induce similar network training dynamics as the real data. The main difference between ours and prior approaches is that we are neither limited to the short-range single-step matching nor subject to instability and compute intensity of optimizing over the full training process. Our method balances these two regimes and shows improvement over both. \n\nUnlike prior methods, ours is the first to scale to $128\\times 128$ ImageNet images, which not only allows us to gain interesting insights of the dataset (\\eg in \\refsec{imagenet}) but also may serve as an important step towards practical applications of dataset distillation on real-world datasets. \n\n\\myparagraph{Limitations.} Our use of pre-computed trajectories allows for significant memory saving, at the cost of additional disk storage and computational cost for expert model training. The computational overhead of training and storing expert trajectories is\nquite high. For example, CIFAR experts took $\\sim$3 seconds per epoch (8 GPU hours total for all 200 CIFAR experts) while each ImageNet (subset) expert took $\\sim$11 seconds per epoch (15 GPU hours total for all 100 ImageNet experts). Storage-wise, each CIFAR expert took up $\\sim$60MB of storage while each ImageNet expert took up $\\sim$120MB.\n\n\\section{Appendix}\\subsection{Additional Visualizations}\nWe first include some additional visualizations here. CIFAR-100 (1 image per class) can be seen in \\reffig{cifar-100}. \nAll of Tiny ImageNet (1 image per class) is broken up into Figures \\ref{fig:tiny1} and \\ref{fig:tiny2}. We specifically show the 10 best and worst-performing distilled classes in Figures \\ref{fig:tinygood} and \\ref{fig:tinybad} respectively.\nWe include 10 image per class visualizations of all our 128$\\times$128 ImageNet subsets in Figures \\ref{fig:nette_10}-\\ref{fig:yellow_10}.\n\n\n\n\\subsection{Additional Quantitative Results}\n\n\n\\myparagraph{Analysis of learned learning rates \\boldmath{$\\alpha$}.}\nIn \\reffig{lr}, we explore the effect of our learnable synthetic step size $\\alpha$. The left plot confirms that we learn different values of $\\alpha$ for different combinations of $M$ and $N$. The logic here is that different numbers of synthetic steps $N$ require a different step size $\\alpha$ to cover the same distance as $M$ real steps. The right plot illustrates the practical benefits of our adaptive learning rate; instead of yet another hyper-parameter to tune, our adaptive learning rate works from a wide range of initializations.\n\n\n\n\\myparagraph{Effects of ZCA Whitening}\nNote that \\texttt{DC}, \\texttt{DSA}, and \\texttt{DM} do not use ZCA normalization, while \\texttt{KIP} started using ZCA as it was a ``crucial ingredient for [their] strong results.'' We report our results w\/o ZCA in \\reffig{zca} (Left).\nWe find that ZCA normalization is \\textit{not} critical to our performance. \nHowever, the expert models trained without ZCA normalization take significantly longer to converge. Thus, when distilling using these models as experts, we must use a larger value of $T^+$ (and therefore save more model snapshots). When we use a larger value of $T^+$ for non-ZCA distillations, we get results comparable to or even better than those of the ZCA distillations. \nIn short, ZCA helps expert convergence but does not notably improve distillation performance. \n\n\\input{figText\/zca}\n\n\\subsubsection{Additional Ablation Studies}\n\n\n\\myparagraph{Initialization, normalization, and augmentation.}\nIn the main paper, we show ablations over several hyper-parameters. Here, we study the role of initialization, data normalization, and data augmentation for CIFAR-100 (1 image per class) in \\reftbl{ablation}. For initialization in particular, recall that we typically initialize our synthetic images with real samples. Here, we evaluate initializing with Gaussian noise instead. Visualizations of these distilled sets can be seen in Figures \\ref{fig:noise}-\\ref{fig:noaug}. We also include a visualization of a set distilled with only one expert trajectory in \\reffig{1exp}.\n\\input{figText\/ablation_tab}\n\n\\input{figText\/ablation_plots}\n\\input{figText\/ablation_plots3}\n\n\\myparagraph{Performance w.r.t. the number of expert trajectories.}\nSince they effectively make up our method's ``training set,'' it is reasonable to assume that having more expert trajectories would lead to better performance. We see that this is indeed the case for the CIFAR-100, 1 image per class setting in \\reffig{experts} (left). However, what's most interesting is the sharp, logarithmic increase in validation accuracy w.r.t. the number of experts. We note the most amount of improvement when increasing from 1 to 20 experts but see almost complete saturation by the time we reach 200. Given how high-dimensional the parameter space of a neural network is, it is remarkable that we can achieve such high performance with so few expert trajectories. \\vspace{10pt}\n\n\\myparagraph{Performance w.r.t. expert time-step range.}\nWhen we initialize our student networks, we do so at a randomly selected time-step from an expert trajectory. We find that it is important to put an upper bound on this starting time-step (\\reffig{experts}, right). If the upper bound is too high, the synthetic data receives gradients from points where the experts movements are small and uninformative. If it is too low, the synthetic data is never exposed to mid and later points in the trajectories, missing out on a significant portion of the training dynamics.\n\n\n\\begin{figure}\n \\centering\n \\includegraphics[width=\\linewidth]{figures\/CIFAR100.pdf}\n \\caption{CIFAR-100, 1 Image Per Class}\n \\lblfig{cifar-100}\n \\vspace{-9pt}\n\\end{figure}\n\n\\subsection{Experiment Details}\n\\myparagraph{Hyper-Parameters.}\nIn \\reftbl{hparams}, we enumerate the hyper-parameters used for our results reported in the main text. Limited compute forced us to batch our synthetic data for some of the larger sets. The ``ConvNet'' architectures are as explained in the main text.\n\\input{figText\/table_hparams}\n\n\\myparagraph{Compute resources.} We had a relatively limited compute budget for our experiments, using any GPU we could access. As such, our experiments were run on a mixture of RTX2080ti, RTX3090, and RTX6000 GPUs. The largest amount of VRAM we used for a single experiment was 144GB over 6xRTX6000 GPUs.\n\n\\myparagraph{Training Time.} Distillation time varied based on dataset and type and number of GPUs used. Regardless of dataset or compute resources, time per distillation iteration scaled linearly with the number of synthetic steps $N$. For CIFAR-100, 1 image per class with $N=20$, we had an average time of 0.6 seconds per distillation step when using a single RTX3090. We ran our experiments for 10000 distillation steps but saw the most improvement within the first 1000.\n\nOur distillation time, in general, is comparable to \\texttt{DC\/DSA}, as they also utilize a bi-level optimization. In the 10 img\/class setting (for example), \\texttt{DC\/DSA} trains on the synthetic data for 50 epochs on the between each update. We include a sample distillation curve in \\reffig{zca} (Right). Both experiments were run on RTX3090. Note that \\texttt{KIP} requires over 1,000 GPU hours.\n\nRegarding the distillation time for learning different sets on CIFAR10\/100 and TinyImageNet, we report them in \\reftbl{time}. Note that most improvement occurs within the first 1k iterations, but we continue training for 10k.\n\\input{figText\/time}\n\n\\myparagraph{KIP to NN}\nIn the \\texttt{KIP} paper, results are presented for images distilled using the neural tangent kernel method and then evaluated by training a modified width-1024 ConvNetD3. Aside from the increased width of the finite model, the ConvNet architecture used in the \\texttt{KIP} paper also has an additional 1-layer convolutional stem.\n\nUsing the training notebook provided with the \\texttt{KIP} paper, we perform an exhaustive search over a reasonable set of hyper-parameters for the KIP to width-128 NN problem: \\texttt{checkpoint} $\\in$ \\{112, 335, 1000\\}, \\texttt{weight\\_decay} $\\in $ \\{0, 0.0001, 0.001, 0.01\\}, \\texttt{aug} $\\in$ \\{\\texttt{True}, \\texttt{False}\\}, \\texttt{zca} $\\in$ \\{\\texttt{True}, \\texttt{False}\\}, \\texttt{label\\_learn} $\\in$ \\{\\texttt{True}, \\texttt{False}\\}, and \\texttt{norm} $\\in$ \\{\\texttt{none}, \\texttt{instance}\\}. The architecture originally used for KIP to NN in the \\texttt{KIP} paper contained no normalization layers. However, we found that with the smaller width, this model could not even converge on the synthetic training data for CIFAR-100, so we added instance normalization layers as found in the ConvNets we and \\texttt{DC}, \\texttt{DSA}, and \\texttt{DM} use.\n\nIn \\reftbl{kip_hyperparams}, we include the optimal hyper-parameters from this search that were used to obtain the KIP to NN (128\\nobreakdash-width) values reported in the main text.\n\n\\input{figText\/kip_to_nn_params}\n\n\n\n\n\n\n\n\\begin{figure}\n \\centering\n \\includegraphics[width=\\linewidth]{figures\/Tiny_Good.pdf}\n \\vspace{-20pt}\n \\caption{Most-correct distilled images for Tiny ImageNet \\smaller{($\\geq 30\\%$)}}\n \\label{fig:tinygood}\n \\vspace{-8pt}\n\\end{figure}\n\\begin{figure}\n \\centering\n \\includegraphics[width=\\linewidth]{figures\/Tiny_Bad.pdf}\n \\vspace{-8pt}\n \\caption{Least-correct distilled images for Tiny ImageNet \\smaller{($\\leq 4\\%$)}}\n \\label{fig:tinybad}\n \\vspace{-8pt}\n\\end{figure}\n\n\n\\clearpage\n\n\\begin{figure}\n \\centering\n \\includegraphics[width=\\linewidth]{figures\/cifar100-noise.pdf}\n \\caption{CIFAR-100, Initialized as Random Noise}\n \\label{fig:noise}\n\\end{figure}\n\\begin{figure}\n \\centering\n \\includegraphics[width=\\linewidth]{figures\/cifar100-nozca.pdf}\n \\caption{CIFAR-100, No ZCA Whitening}\n \\label{fig:nozca}\n\\end{figure}\n\\begin{figure}\n \\centering\n \\includegraphics[width=\\linewidth]{figures\/cifar100-noaug.pdf}\n \\caption{CIFAR-100, No Differentiable Augmentation}\n \\label{fig:noaug}\n\\end{figure}\n\\begin{figure}\n \\centering\n \\includegraphics[width=\\linewidth]{figures\/cifar100-single.pdf}\n \\caption{CIFAR-100, Only 1 Expert Trajectory}\n \\label{fig:1exp}\n\\end{figure}\n\n\\begin{figure*}\n \\centering\n \\includegraphics[width=\\linewidth]{figures\/tiny_all_1.pdf}\n \\caption{Tiny ImageNet, 1 Image Per Class (Classes 1-100)}\n \\label{fig:tiny1}\n\\end{figure*}\n\n\\begin{figure*}\n \\centering\n \\includegraphics[width=\\linewidth]{figures\/tiny_all_2.pdf}\n \\caption{Tiny ImageNet, 1 Image Per Class (Classes 101-200)}\n \\label{fig:tiny2}\n\\end{figure*}\n\n\\begin{figure*}\n \\centering\n \\includegraphics[width=\\linewidth]{figures\/imagenet_10\/imagenette_all.pdf}\n \\caption{ImageNette, 10 Images Per Class}\n \\label{fig:nette_10}\n\\end{figure*}\n\\begin{figure*}\n \\centering\n \\includegraphics[width=\\linewidth]{figures\/imagenet_10\/imagewoof_all.pdf}\n \\caption{ImageWoof, 10 Images Per Class}\n \\label{fig:woof_10}\n\\end{figure*}\n\\begin{figure*}\n \\centering\n \\includegraphics[width=\\linewidth]{figures\/imagenet_10\/imagesquawk_all.pdf}\n \\caption{ImageSquawk, 10 Images Per Class}\n \\label{fig:squawk_10}\n\\end{figure*}\n\\begin{figure*}\n \\centering\n \\includegraphics[width=\\linewidth]{figures\/imagenet_10\/imagemeow_all.pdf}\n \\caption{ImageMeow, 10 Images Per Class}\n \\label{fig:meow_10}\n\\end{figure*}\n\\begin{figure*}\n \\centering\n \\includegraphics[width=\\linewidth]{figures\/imagenet_10\/imagefruit_all.pdf}\n \\caption{ImageFruit, 10 Images Per Class}\n \\label{fig:fruit_10}\n\\end{figure*}\n\\begin{figure*}\n \\centering\n \\includegraphics[width=\\linewidth]{figures\/imagenet_10\/imageyellow_all.pdf}\n \\caption{ImageYellow, 10 Images Per Class}\n \\label{fig:yellow_10}\n\\end{figure*}\n\\vfill\\break\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\n\nAs one of the most intriguing features of quantum mechanics, quantum entanglement has been regarded as a key resource to detect and understand properties of complex many-body systems. For instance, recently a great deal of effort has been devoted to establishing a deep understanding about quantum phase transitions (QPTs) via the examination of their entanglement behaviors~\\cite{osterloh2002scaling,vidal2003entanglement,lambert2004entanglement,gu2004entanglement,le2008entanglement,pereira2016effects,wu2004quantum,wu2006linking,yang2005reexamination,braun2017quantum,vidal2006concurrence}. Quantum phase transition is defined as a transition between distinct ground states of quantum many-body systems when a controlled parameter in the Hamiltonian crosses a critical point~\\cite{sachdev2011quantum}. Compared with the great development of studying QPTs in equilibrium systems, the understanding of the non-equilibrium dynamical phase transitions (DPTs) is still inadequate~\\cite{polkovnikov2011colloquium}. Although some investigations have been done to study the properties of DPTs~\\cite{lo1990ising,jung1990scaling,ryu1996dynamical,sides1998kinetic,yuzbashyan2006relaxation,sciolla2010quantum,diehl2010dynamical,calabrese2011quantum,mitra2012time,heyl2013dynamical,heyl2014dynamical,klinder2015dynamical}, only few works linked it with entanglement~\\cite{lin2016non,jurcevic2017direct,Zhang2017observation}. Recently, the observation of many-body DPTs with up to 10 trapped ion qubits has shown that DPTs in the simulated Ising models can control entanglement production~\\cite{jurcevic2017direct}. Another elegant experiment with a quantum simulator composed of up to 53 ion qubits with long-range Ising interactions has also uncovered the connection between DPTs and many-body correlations~\\cite{Zhang2017observation}. Inspired by those achievements, we investigate theoretically an extension of this connection in quantum systems other than Ising model. For instance, a rich variety of quantum phases in spin-orbit-coupled (SOC) Bose-Einstein condensates (BECs) has been investigated theoretically and experimentally~\\cite{dalibard2011colloquium, galitski2013spin, zhou2013unconventional, goldman2014light, zhai2015degenerate, zhang2016properties}. A proposal of simulating spin DPT by ultra-cold atoms has been reported in this system~\\cite{poon2016quantum}. We are wondering whether this kind of DPT can be characterized by the behavior of two-mode entanglement in the synthetic spin space. \n\nIn this paper, by introducing an external perturbation (switching on an additional lattice potential) in the Hamiltonian of a spin-1\/2 BEC of $^{87}$Rb atoms with 1D synthetic spin-orbit coupling, we study the behavior of dynamical two-mode entanglement in spin space and use it to characterize spin DPT. A previous analysis~\\cite{poon2016quantum} shows that the additional lattice potential can drive a periodic evolution of the system in spin space, and there exists a DPT between magnetized and unmagnetized states at a critical lattice depth. By examining the entropic entanglement measure~\\cite{hines2003entanglement, xie2006quantum, byrnes2012quantum}, we show that the periodic motion in spin space leads to the periodic evolution of two-mode entanglement, and the time-averaged entropic entanglement over an oscillation period reaches a maximal value at the DPT. \n\nAlthough the entropy of entanglement provides an excellent indicator of DPT in such a system, it works only for pure states at zero temperature and requires reconstruction of the quantum states via tomography in experiments. To overcome these difficulties, we also use a correlation-based entanglement criterion that is suitable for measuring to detect DPT and study the influence of thermal excitations. Specifically, a criterion was originally introduced by Hillery and Zubairy (HZ)~\\cite{hillery2006entanglement} and developed for double-well BEC systems~\\cite{he2011einstein, he2012einstein}. We show that the HZ criterion is an excellent proxy for entropic entanglement measure, hence can be used to characterize the DPT. Moreover, we find that the thermal effects will change the critical point of the DPT, which can also be confirmed by the shifts of the maximum of time-averaged HZ entanglement parameter. \n\nThe remainder of this paper is organized as follows. In Sec.~\\ref{Sec:Model}, we introduce the model of DPT induced by external perturbation in BEC with SOC. We then characterize the two-mode entanglement of the system across the DPT via von Neumann entropy and HZ criterion in Sec.~\\ref{Sec:entropy} and Sec.~\\ref{Sec:HZ}, respectively. The results show that the behavior of entanglement can infer the existence of DPT. In Sec.~\\ref{Sec:thermal}, we discuss finite temperature effect and show that the entanglement is robust to thermal excitations and can also be used to signature DPT. Finally we summarize our results and discuss the experimental feasibility in Sec.~\\ref{Sec:conclusion}.\n\n\\section{Model}\\label{Sec:Model}\n\nWe consider a two-component BEC of $^{87}$Rb atoms with one-dimensional (1D) synthetic spin-orbit coupling. As discussed in Ref.~\\cite{poon2016quantum}, such a system features a quantum spin DPT in the presence of an additional lattice potential as external perturbation. By labeling the two atomic components as (pseudo-)spin up and down, the Hamiltonian reads (with natural units $\\hbar=m=1$)\n\\begin{eqnarray}\n\tH&=&H_{0}+H_{\\rm int} ,\\nonumber\\\\\n\tH_{0}&=&\\sum_{s,s'=\\uparrow,\\downarrow}\\int d^3r \\psi_s^\\dagger \\left( -\\frac{\\nabla_r^2}{2}+ik_0\\partial_x\\sigma_z+\\frac{\\Omega}{2}\\sigma_x \\right)_{ss'}\\psi_{s'} , \\nonumber\\\\\n\tH_{\\rm int}&=&\\int d^3r \\frac{g_s}{2}\\left(\\psi_{\\downarrow}^{\\dagger}\\psi_{\\downarrow}^{\\dagger}\\psi_{\\downarrow}\\psi_{\\downarrow}+\\psi_{\\uparrow}^{\\dagger}\\psi_{\\uparrow}^{\\dagger}\\psi_{\\uparrow}\\psi_{\\uparrow}\\right) \\nonumber\\\\\n\t&&+\\int d^3r g_a \\psi_{\\downarrow}^{\\dagger}\\psi_{\\uparrow}^{\\dagger}\\psi_{\\uparrow}\\psi_{\\downarrow}.\n\\end{eqnarray}\nHere $H_0$ is the single-particle Hamiltonian including 1D SOC along the $x$ direction, and $H_{\\rm int}$ is the interaction term. The field operators $\\psi_{s}$ ($\\psi_{s}^{\\dagger}$) annihilates (creates) an atom with spin $s=\\uparrow, \\downarrow$, $\\Omega$ is the Raman coupling strength, and $k_0$ is determined by the Raman laser's wave vector. Note that we assume a spin-symmetric interaction with $g_{\\uparrow\\uparrow}=g_{\\downarrow\\downarrow}=g_{s}$ and $g_{\\uparrow\\downarrow}=g_{a}$, and the two-photon detuning $\\delta=0$ for simplicity. \n\nBy diagonalizing $H_{0}$, the eigenenergies of two subbands are given by $E_{k}^{\\pm}={k^2}\/{2}\\pm \\sqrt{k_0^2k_x^2+{\\Omega^2}\/{4}}$. For $\\Omega<2k_0^2$, the lower subband has a double-minimum structure at $k_x=\\pm k_{\\rm min}$ with $k_{\\rm min}=k_0\\sqrt{1-\\Omega^2\/4k_0^4}$. For $\\Omega>2k_0^2$, it has only a single minimum at $k_{\\rm min}=0$. This structure is the origin of phase transitions among magnetized plane-wave phase, unpolarized stripe phase, and zero-momentum normal phase.\n\nWhen the interaction is concerned, we can assume that the ground-state wave function of the condensate takes the form\n\\begin{equation}\n\t\\Psi=\\sqrt{n}\\left[ \\alpha\\left( \n\t\t\t\\begin{array}{c}\n\t\t\t\t\\cos\\theta \\\\ -\\sin\\theta\n\t\t\t\\end{array}\n\t\\right)e^{ik_mx}+\\beta\\left( \n\t\t\\begin{array}{c}\n\t\t\t\\sin\\theta \\\\ -\\cos\\theta\n\t\t\\end{array}\n\t\\right)e^{-ik_mx} \\right] ,\n\t\\label{ground-state}\n\\end{equation}\nwhere $\\alpha$ and $\\beta$ are arbitrary complex numbers satisfying $|\\alpha|^2+|\\beta|^2=1$, and $\\tan2\\theta=\\Omega\/(2k_0k_m)$. In general, the momenta of many-body eigenstates $\\pm k_m$ are different from the minimal positions of single-particle dispersion $\\pm k_{\\rm min}$. But in our parameter region where $\\Omega\/k_0^2<0.6$ and $g_s n \\approx 1.0k_0^2$ ($n=n_{\\uparrow}+n_\\downarrow$ represents the condensate density), we can safely set $k_m\\simeq k_{\\rm min}$. The remarkable property is that there exist two critical Raman couplings~\\cite{li2012quantum}, given by\n\\begin{eqnarray}\n\t\\Omega_{c1}&=&2(k_0^2-2G_2) ,\\\\\n\t\\Omega_{c2}&=&2\\sqrt{(k_0^2+G_1)(k_0^2-2G_2)\\frac{2G_2}{G_1+2G_2}} ,\n\\end{eqnarray}\nwhere $\\Omega_{c1}>\\Omega_{c2}$ and $G_{1}={n}(g_{s}+g_{a})\/4$, $G_{2}={n}(g_{s}-g_{a})\/4$. When $\\Omega<\\Omega_{c2}$, the ground state is a superposition of states with $k_x=\\pm k_{m}$ ($|\\alpha|^2=|\\beta|^2=1\/2$), which is referred as the stripe phase. For $\\Omega_{c2}<\\Omega<\\Omega_{c1}$, there exist two degenerate states with $k_x=k_m$ ($|\\alpha|=1$, $|\\beta|=0$) or $k_x=-k_m$ ($|\\alpha|=0$, $|\\beta|=1$), called the magnetized phase. If the Raman coupling is large enough that $\\Omega>\\Omega_{c1}$, the ground state is at $k_x=0$, giving the normal phase~\\cite{dalibard2011colloquium, galitski2013spin, zhou2013unconventional, goldman2014light, zhai2015degenerate, zhang2016properties, li2012quantum, martone2012anisotropic}.\n\nFor a magnetized phase with, e.g., $k_x=k_m$ ($|\\alpha|=1, |\\beta|=0$), an additional external potential is switched on at $t=0$~\\cite{poon2016quantum},\n\\begin{equation}\n\tV_{\\rm ex}(r,t)=\\left\\{\n\t\t\\begin{array}{lc}\n\t\t\t0, & t<0,\\\\\n\t\t\tV_{0}\\int d^3r \\cos^2k_m x(\\psi_{\\uparrow}^{\\dagger}\\psi_{\\uparrow}+\\psi_{\\downarrow}^{\\dagger}\\psi_{\\downarrow}), & t>0,\n\t\t\\end{array}\n\t\\right. \n\t\\label{Vex}\n\\end{equation}\nwhere $V_0$ is the strength of the perturbation. $V_{\\rm ex}$ can drive resonant couplings between the two degenerate magnetized phases $\\psi_R$ and $\\psi_L$ at $k_{m}$ and $-k_m$, respectively, as indicated schematically in Fig.~\\ref{fig:BEC_ground_state}. The normalized time-dependent condensate wave function takes the form \n\\begin{equation}\n\t|\\Psi_{\\rm BEC}(r,t)\\rangle = \\frac{1}{\\sqrt{N!}} \\left[ \\alpha^{\\ast}(t)\\psi_{R}^{\\dagger}+\\beta^{\\ast}(t)\\psi_{L}^{\\dagger} \\right]^{N}|vac\\rangle ,\n\t\\label{wave-function}\n\\end{equation}\nwhere $N$ is the total number of atoms, $|vac\\rangle$ denotes the vacuum state, and $\\psi_{R}=(\\cos\\theta\\psi_{\\uparrow}-\\sin\\theta\\psi_{\\downarrow})e^{ik_mx}$, $\\psi_{L}=(\\sin\\theta\\psi_{\\uparrow}-\\cos\\theta\\psi_{\\downarrow})e^{-ik_mx}$ are the field operators.\n\\begin{figure}[t]\n\t\\centering\n\t\\includegraphics[width=0.35\\textwidth]{figure1}\n\t\\caption{The external perturbation $V_{\\rm ex}$ described by Eq.~(\\ref{Vex}) can induce resonant couplings between the two magnetized phases, marked as $\\psi_{R}$ and $\\psi_{L}$ at $k_m$ and $-k_m$, respectively. }\n\t\\label{fig:BEC_ground_state}\n\\end{figure}\n\nThe quantum spin dynamics of the system under the perturbation given in Eq.~(\\ref{Vex}) has been studied in Ref.~\\cite{poon2016quantum}. It is found that a critical external perturbation strength\n\\begin{equation}\n\tV_{0,{\\rm crit}}=\\frac{2(E_{s}-E_{m})}{\\sin2\\theta} \n\t\\label{V0c}\n\\end{equation}\nis required to have a full transition from one magnetized phase to another. Specifically, when the perturbation strength is below $V_{0,{\\rm crit}}$, the maximum number of atoms that can transit from the initial state with $k_x=k_m$ to the other degenerate state at $k_x=-k_m$ is less than half of the total atom number. On the other hand, when the perturbation strength $V_{0}>V_{0,{\\rm crit}}$, all atoms can fully transit during time evolution. Thus, the critical perturbation strength $V_{0,{\\rm crit}}$ reveals a quantum DPT. We also stress that the critical point $V_{0,{\\rm crit}}$ depends on the condensate interaction energy $g_sn$ via $E_{s}=2G_{1}\\cos^{2}\\theta\\sin^{2}\\theta$ and $E_{m}=G_{2}\\cos^{2}2\\theta$, which characterize the interaction energies for the stripe phase and magnetized phase, respectively. \n\nThe transition mentioned above can be characterized by an order parameter defined as the time average of spin polarization $\\langle s_z(t)\\rangle= |\\alpha(t)|^2-|\\beta(t)|^2$ ($s_z$ is the Pauli matrix acting on the pseudospin space spanned by the two magnetized states) over an oscillation period $T_R$~\\cite{poon2016quantum}\n\\begin{equation}\n\t\\bar{M}=\\frac{1}{T_{R}}\\int_{0}^{T_{R}}\\langle s_{z}(t)\\rangle dt .\n\t\\label{order-parameter}\n\\end{equation}\nIt is straightforward to check that $\\bar{M}> 0$ when the perturbation strength is lower than the critical point and the system is in the dynamical magnetized phase, and $\\bar{M}\\equiv 0$ when the system is dynamically non-magnetized with perturbation strength exceeding the critical point.\n\n\\section{Entropy of entanglement}\\label{Sec:entropy}\n\nPrevious works have studied entanglement of pure states of bipartite systems using entropy of entanglement, which is the von Neumann entropy of the reduced density operator of either of the subsystems. In this section we first study whether the quantum spin DPT at zero temperature can be characterized by the entropic entanglement measure. As the two degenerate ground states of the magnetized phase (labeled by $\\psi_{R}$ and $\\psi_{L}$) can be considered as an analogue of a double-well BEC in momentum space, we can adopt the same method which is commonly used in spatially separated double-well BEC~\\cite{spekkens1999spatial,he2012einstein,he2011einstein,vidal2007entanglement}, and treat $\\psi_{R}$ and $\\psi_{L}$ as two modes within the two-mode approximation~\\cite{spekkens1999spatial}. Our results show that the behavior of the entropic entanglement can act as a indicator of the quantum spin DPT in this system.\n\nBy applying the two-mode approximation, the time-dependent wave function in Eq.~(\\ref{wave-function}) can be expanded in term of Fock states as\n\\begin{equation}\n\t|\\Psi_{\\rm BEC}(t)\\rangle = \\sum_{n=0}^{N} \\sqrt{C_{N}^{n}} \\alpha^{\\ast n}\\beta^{\\ast N-n}|n,N-n\\rangle_{R,L} ,\n\t\\label{two-mode_wave_function}\n\\end{equation} \nwhere $C_{N}^{n}$ is the binomial coefficient, and the Fock basis are defined as\n\\begin{equation}\n\t|n,N-n\\rangle_{R,L}=\\frac{(\\psi_{R}^{\\dagger})^{n}}{\\sqrt{n!}}\\frac{(\\psi_{L}^{\\dagger})^{N-n}}{\\sqrt{(N-n)!}}|vac\\rangle .\n\t\\label{number_state}\n\\end{equation}\n\nThe dynamics can be described by defining the field operator $\\psi(t)=\\alpha(t)\\psi_{R}+\\beta(t)\\psi_{L}$, and considering Heisenberg equation\n\\begin{equation}\n\ti\\frac{d\\psi(t)}{dt}=\\left[ \\psi(t), H_{0}+H_{\\rm int}+V_{\\rm ex} \\right],\n\t\\label{Heisenberg}\n\\end{equation}\nleading to the following equation of motion\n\\begin{equation}\n\ti\\frac{d}{dt}\\left( \n\t\t\\begin{array}{c}\n\t\t\t \\alpha(t) \\\\ \\beta(t)\n\t\t\\end{array}\n\\right)=H_{\\rm eff}\\left( \n\t\\begin{array}{c}\n\t\t\\alpha(t) \\\\ \\beta(t)\n\t\\end{array}\n\\right).\n\t\\label{motion}\n\\end{equation}\nHere, the effective two-mode Hamiltonian is given by\n\\begin{eqnarray}\n\tH_{\\rm eff}&=&E_{k}^{-}+\\frac{V_{0}}{2}+G_{1}+V_{p}s_{x}+E_{m}(|\\alpha|^{2}-|\\beta|^{2})s_{z} \\nonumber\\\\\n\t&&+2E_{s}\\left[ {\\rm Re}(\\alpha\\beta^{\\ast})s_{x}-{\\rm Im}(\\alpha\\beta^{\\ast})s_{y} \\right],\n\t\\label{Heff}\n\\end{eqnarray}\nwhere $s_{x,y,z}$ are spin matrices spanned by the two magnetized states, and $V_{p}=V_{0}\\cos\\theta\\sin\\theta\/2$ represents the coupling strength induced by external perturbation.\n\nTheoretically, the von Neumann entropy $E_{\\rm vn}=-{\\rm Tr}\\left[ \\rho_{R}\\log\\rho_{R} \\right]$ can be used to evaluate the entanglement between two subsystems, where $\\rho_{R}={\\rm Tr}_{L}[\\rho_{R,L}]$ is the reduced density operator for $\\psi_R$ and $\\rho_{R,L}=|\\Psi_{\\rm BEC}\\rangle\\langle\\Psi_{\\rm BEC}|$. Considering the specific form of the time-dependent wave function $|\\Psi_{\\rm BEC}\\rangle$ described in Eq.~(\\ref{two-mode_wave_function}), the entropy of entanglement between $\\psi_R$ and $\\psi_L$ thus reads\n\\begin{equation}\n\tE_{\\rm vn}=-\\sum_{n=0}^{N}C_{N}^{n}|\\alpha|^{2n}|\\beta|^{2(N-n)}\\log_2\n\t\\left[C_{N}^{n}|\\alpha|^{2n}|\\beta|^{2(N-n)}\\right].\n\t\\label{entanglement_entropy}\n\\end{equation}\nHere, $E_{\\rm vn}=0$ for separable product states, and $E_{\\rm max}=\\log_2(N+1)$ for maximally entangled states when all atoms are equally represented~\\cite{hines2003entanglement}. As a measure of the entanglement, we plot in Fig.~\\ref{fig:entropy} the ratio of $E_{\\rm vn}$ to its corresponding maximum value $E=E_{\\rm vn}\/E_{\\rm max}$ for various numbers $N$ and external perturbation strength $V_0$~\\cite{hines2003entanglement, xie2006quantum, byrnes2012quantum}. The values of $E$ range from $0$ to $1$. Note that the ratio $E$ only represents how much the entanglement of the state $|\\Psi_{\\rm BEC}\\rangle$ with total $N$ atoms is less than its corresponding maximum entanglement, $\\log(N+1)$. It is meaningless to compare the values of $E$ for different $N$. As a comparison, the dynamical evolution of the spin polarization $\\langle s_z \\rangle$ is also shown using the same parameter $V_0$. \n\\begin{figure*}[t]\n\t\\centering\n\t\\includegraphics[width=0.245\\textwidth]{sz_06V}\n\t\\includegraphics[width=0.245\\textwidth]{sz_0999V}\n\t\\includegraphics[width=0.245\\textwidth]{sz_1001V}\n\t\\includegraphics[width=0.245\\textwidth]{sz_14V}\\\\\n\t\\includegraphics[width=0.245\\textwidth]{entropy_06V}\n\t\\includegraphics[width=0.245\\textwidth]{entropy_0999V}\n\t\\includegraphics[width=0.245\\textwidth]{entropy_1001V}\n\t\\includegraphics[width=0.245\\textwidth]{entropy_14V}\n\\caption{The evolution of entropic entanglement $E$ is related to the evolution of pseudospin $\\langle s_z\\rangle$. From left to right, the external perturbation strength is (a)(e) $V_0 = 0.6V_{0,{\\rm crit}}$, (b)(f) $0.999V_{0,{\\rm crit}}$, (c)(g) $1.001V_{0,{\\rm crit}}$, and (d)(h) $1.4V_{0,{\\rm crit}}$. For each situation, we plot the entropic entanglement for various numbers of atoms $N=1$ (blue), $N=10$ (green), and $N=100$ (red solid). Other parameters are used as $g_{s}n=1.0k_{0}^{2}$ and $\\Omega=0.3k_{0}^{2}$.}\n\\label{fig:entropy}\n\\end{figure*}\n\nFrom Fig.~\\ref{fig:entropy}, it is clear that the entropic entanglement evolves with time periodically, and the period is related to that of the spin oscillation. In the regime where the perturbation strength is below the critical value $V_{0}0$. In such a case, the oscillatory period of the two-mode entanglement is as same as that of the spin oscillation. When all the atoms are at the magnetized state $k_x=\\pm k_m$, we have $\\langle s_z\\rangle=\\pm1$ and $E=0$, i.e., the two-mode entanglement doesn't exist. When atoms distribute equally in the two modes, i.e., $\\langle s_{z}\\rangle$ approaches zero, the two-mode entanglement reaches a maximal value. On the other hand, in the regime where the external perturbation strength exceeds the critical value, as shown in Figs.~\\ref{fig:entropy}(c)(d)(g)(h), the spin evolves on the entire spherical surface of the Bloch sphere for one oscillation period $T_R$. A complete transition of atoms from one magnetized phase to another occurs, i.e., $\\langle s_z\\rangle=-1$, at $T_R\/2$, where $E=0$ ends one complete cycle of entanglement motion. Therefore, for this case the period of entanglement oscillation is half of that of the spin evolution. \n\\begin{figure}[t]\n\t\\centering\n\t\\includegraphics[width=0.4\\textwidth]{entropy_mean_N}\n\t\\caption{The time-averaged entropic entanglement measure $\\bar{E}$ calculated as a function of perturbation strength $V_0\/V_{0,{\\rm crit}}$ (left axis). Different lines represent different total numbers of atoms $N=1$ (blue), $N=10$ (green), and $N=100$ (red). The entanglement displays a sharp peak at the critical value $V_{0,{\\rm crit}}$ where the DPT occurs as characterized by the order parameter $\\bar{M}$ (black dashed, right axis). This observation suggests that the critical behavior of entropic entanglement measure can be used to identify the DPT. Other parameters are used as those in Fig.~\\ref{fig:entropy}.}\n\t\\label{fig:time-average_entropy}\n\\end{figure}\n\nThe spin DPT across $V_{0,{\\rm crit}}$ can be characterized by the order parameter $\\bar{M}$ defined in Eq.~(\\ref{order-parameter}), which is depicted by the dashed black line in Fig.~\\ref{fig:time-average_entropy}. To reveal the connection between DPT and entanglement, we define the time-averaged entropic entanglement measure as\n\\begin{equation}\n\t\\bar{E}(V_{0})=\\frac{1}{T_{R}}\\int_{0}^{T_{R}}E(t)dt,\n\t\\label{time-average_entropy}\n\\end{equation}\nwhich is plotted as a function of perturbation strength for different particle numbers by solid lines in Fig.~\\ref{fig:time-average_entropy}. Apparently, the entanglement measure ${\\bar E}$ features a sharp peak at the critical point where the order parameter ${\\bar M}$ distinctly changed from $\\bar{M}>0$ (dynamical magnetized phase) to $\\bar{M}=0$ (dynamical unmagnetized phase), indicating that one can achieve maximal entanglement in the vicinity of the critical point of DPT. We also notice that the entanglement peak at DPT is less sharp with increasing particle number $N$. This is because when we define the entanglement measure $E$, the maximal entanglement $E_{\\rm max}$ in the denominator increases logarithmically with $N$. Thus, when $N$ goes larger, the normalization leads to a flatter peak of $E$, making the usage of entropic entanglement measure as an indicator of DPT rather ambiguous. \n\n\\section{Correlation-based entanglement measure for two-mode system}\\label{Sec:HZ}\n\nAlthough the entropy of entanglement is a useful measure to characterize the DPT, it is an entanglement measure only for pure states and therefore cannot account for the effects of finite temperatures. In addition, measuring entropic entanglement requires reconstruction of the quantum states via tomography, which demands the state-of-art technique in experiments even for special systems~\\cite{islam2015measuring} and remains challenging for general quantum systems. To circumvent these difficulties, in the following we use an experimentally feasible correlation-based entanglement criterion to detect entanglement undergoing a DPT and discuss the influence of thermal excitations. \n\nA sufficient entanglement criterion for a two-mode system is the operator product measure $|\\langle ab^{\\dagger}\\rangle|^2>\\langle a^\\dagger ab^\\dagger b\\rangle$ given by Hillery and Zubairy (HZ)~\\cite{hillery2006entanglement}, where $a$ and $b$ denote the annihilation operators of the two modes. \nFor double-well BEC systems, a spin version of HZ criterion has also been developed~\\cite{he2011einstein, he2012einstein}. Specifically, the state is entangled if\n\\begin{equation}\n\tE_{\\rm HZ}=\\frac{\\Delta^2 J_x+\\Delta^2 J_y}{\\langle \\hat{N} \\rangle\/2}<1,\n\t\\label{HZ_J}\n\\end{equation}\nwhere the spin operators $J_{x}=(\\psi_{R}^{\\dagger}\\psi_{L}+\\psi_{R}\\psi_{L}^{\\dagger})\/2$, $J_{y}=(\\psi_{R}^{\\dagger}\\psi_{L}-\\psi_{R}\\psi_{L}^{\\dagger})\/(2i)$, $J_{z}=(\\psi_{R}^{\\dagger}\\psi_{R}-\\psi_{L}^{\\dagger}\\psi_{L})\/2$ with canonical commutation relations $[J_{x},J_{y}]=iJ_{z}$ (and cyclic permutations). The variance of measurements of $J_{x,y}$ is defined as $\\Delta^2 J_{x,y}\\equiv \\langle J^2_{x,y}\\rangle-\\langle J_{x,y}\\rangle^2$, and the expectation values for $\\hat{N}=\\psi_{R}^{\\dagger}\\psi_{R}+\\psi_{L}^{\\dagger}\\psi_{L}$ are fixed at $N$ for state~(\\ref{two-mode_wave_function}). It is convenient to quantify entanglement using spin-operator methods, and this type of spin-operator variance has been measured experimentally by expanding the two condensates and measuring the absorption imaging average fringe visibility~\\cite{esteve2008squeezing, ji2014experimental}. \nWe emphasize that the variances $\\Delta^2 J_{x,y}$ for the state~(\\ref{two-mode_wave_function}) are proportional to the number of atoms, so that the value of $E_{\\rm HZ}$ is independent of $N$ as shown in spatial double-well BEC systems~\\cite{he2012einstein}.\n\nTo get the evolution of the entanglement parameter $E_{\\rm HZ}$, we consider the Heisenberg equation for $\\langle \\psi_{R}^{\\dagger}\\psi_{R} \\rangle$, $\\langle \\psi_{R}^{\\dagger}\\psi_{L} \\rangle$, $\\langle\\psi_{L}^{\\dagger}\\psi_{L}\\rangle$, $\\langle \\psi_{R}^{\\dagger}\\psi_{R}\\psi_{L}^{\\dagger}\\psi_{L} \\rangle$, etc. Due to the nonlinear interaction terms in the Hamiltonian, the sets of equations can't be closed. Thus we truncate the set of equations at the fourth order of operators with mean field approximation, leading to\n\\begin{widetext}\n\\allowdisplaybreaks\n\\begin{eqnarray}\n\t\\frac{d}{dt}\\langle\\psi_{R}^{\\dagger}\\psi_{R}\\rangle &=& -iV_{p}\\left[ \\langle\\psi_{R}^{\\dagger}\\psi_{L}\\rangle-\\langle\\psi_{L}^{\\dagger}\\psi_{R}\\rangle \\right] , \\nonumber\\\\\n\t\\frac{d}{dt}\\langle\\psi_{L}^{\\dagger}\\psi_{L}\\rangle &=& -iV_{p}\\left[ \\langle\\psi_{L}^{\\dagger}\\psi_{R}\\rangle-\\langle\\psi_{R}^{\\dagger}\\psi_{L}\\rangle \\right] , \\nonumber\\\\\n\t\\frac{d}{dt}\\langle\\psi_{R}^{\\dagger}\\psi_{L}\\rangle &=& -i\\left[ V_{p}( \\langle\\psi_{R}^{\\dagger}\\psi_{R}\\rangle-\\langle\\psi_{L}^{\\dagger}\\psi_{L}\\rangle )+2(E_{s}-E_{m})( \\langle \\psi_{R}^{\\dagger}\\psi_{R}^{\\dagger}\\psi_{R}\\psi_{L} \\rangle - \\langle \\psi_{R}^{\\dagger}\\psi_{L}^{\\dagger}\\psi_{L}\\psi_{L}\\rangle ) \\right] , \\nonumber\\\\\n\t\\frac{d}{dt}\\langle\\psi_{R}^{\\dagger}\\psi_{R}\\psi_{L}^{\\dagger}\\psi_{L}\\rangle &\\approx& -iV_{p}\\left[ \\langle \\psi_{R}^{\\dagger}\\psi_{R}\\psi_{R}\\psi_{L}^{\\dagger}\\rangle+\\langle\\psi_{R}^{\\dagger}\\psi_{L}^{\\dagger}\\psi_{L}\\psi_{L}\\rangle-\\langle\\psi_{R}^{\\dagger}\\psi_{R}^{\\dagger}\\psi_{R}\\psi_{L}\\rangle-\\langle\\psi_{R}\\psi_{L}^{\\dagger}\\psi_{L}^{\\dagger}\\psi_{L}\\rangle \\right] , \\nonumber\\\\\n\t\\frac{d}{dt}\\langle\\psi_{R}^{\\dagger}\\psi_{R}^{\\dagger}\\psi_{R}\\psi_{L}\\rangle &\\approx& -i\\left[ V_{p}( \\langle\\psi_{R}^{\\dagger}\\psi_{R}^{\\dagger}\\psi_{R}\\psi_{R}\\rangle+\\langle \\psi_{R}^{\\dagger}\\psi_{R}^{\\dagger}\\psi_{L}\\psi_{L}\\rangle -2\\langle\\psi_{R}^{\\dagger}\\psi_{R}\\psi_{L}^{\\dagger}\\psi_{L}\\rangle)\\right.\\nonumber\\\\\n\t&&\\ \\ \\ \\ \\ \\ \\left.+2(E_{s}-E_{m})\\langle\\psi_{R}^{\\dagger}\\psi_{R}^{\\dagger}\\psi_{R}\\psi_{L}\\rangle( \\langle\\psi_{R}^{\\dagger}\\psi_{R}\\rangle-\\langle\\psi_{L}^{\\dagger}\\psi_{L}\\rangle+1) \\right] , \\nonumber\\\\\n\t\\frac{d}{dt}\\langle\\psi_{R}^{\\dagger}\\psi_{R}^{\\dagger}\\psi_{L}\\psi_{L}\\rangle &\\approx& -i\\left[ 2V_{p}( \\langle\\psi_{R}^{\\dagger}\\psi_{R}^{\\dagger}\\psi_{R}\\psi_{L}\\rangle-\\langle\\psi_{R}^{\\dagger}\\psi_{L}^{\\dagger}\\psi_{L}\\psi_{L}\\rangle )+4(E_{s}-E_{m})\\langle\\psi_{R}^{\\dagger}\\psi_{R}^{\\dagger}\\psi_{L}\\psi_{L}\\rangle( \\langle\\psi_{R}^{\\dagger}\\psi_{R}\\rangle-\\langle\\psi_{L}^{\\dagger}\\psi_{L}\\rangle ) \\right] ,\\nonumber\\\\\n \\frac{d}{dt}\\langle\\psi_{R}^{\\dagger}\\psi_{R}^{\\dagger}\\psi_{R}\\psi_{R}\\rangle&=&-2iV_{p}\\left[ \\langle\\psi_{R}^{\\dagger}\\psi_{R}^{\\dagger}\\psi_{R}\\psi_{L}\\rangle-\\langle\\psi_{R}^{\\dagger}\\psi_{R}\\psi_{R}\\psi_{L}^{\\dagger}\\rangle \\right] .\n\\end{eqnarray}\n\\allowdisplaybreaks[0]\n\\end{widetext}\nOther terms can be derived directly by considering symmetry and conjugate properties of the equations.\n\\begin{figure*}[t]\n\t\\centering\n\t\\includegraphics[width=0.245\\textwidth]{EHZ_compare_06V}\n\t\\includegraphics[width=0.245\\textwidth]{EHZ_compare_0999V}\n\t\\includegraphics[width=0.245\\textwidth]{EHZ_compare_1001V}\n\t\\includegraphics[width=0.245\\textwidth]{EHZ_compare_14V}\n\t\\caption{A comparison between the evolution of HZ entanglement parameter $E_{\\rm HZ}$ (blue dashed) and the entropic entanglement $1-E$ with total number of atoms $N=10$ (green solid) and $N=100$ (red solid). The strength of external perturbation is (a) $V_{0}=0.6V_{0,{\\rm crit}}$, (b) $0.999V_{0,{\\rm crit}}$, (c) $1.001V_{0,{\\rm crit}}$, and (d) $1.4V_{0,{\\rm crit}}$, respectively. The HZ entanglement parameters behaves similarly to the entropic entanglement for different $N$. Other parameters are used as those in Fig.~\\ref{fig:entropy}.}\n\t\\label{fig:HZ-compare}\n\\end{figure*}\n\nBy solving the equations above numerically, we show in Fig.~\\ref{fig:HZ-compare} the evolution of entropic and HZ entanglement signature. To make a direct comparison, the entropic entanglement measure is plotted as $1-E < 1$.\nWe find that the evolution of HZ entanglement parameter exhibits the very same qualitative behavior as that of the entropic entanglement measure. Specifically, if the system is in a magnetized state where $|\\alpha|=0,\\, 1$ and $\\langle s_z\\rangle\\rightarrow \\pm 1$, the HZ entanglement parameter $E_{\\rm HZ}\\rightarrow 1$ showing zero entanglement. On the other hand, the best entanglement would be obtained when the system is in a stripe state where atoms are distributed equally in the two modes with $|\\alpha|^2=0.5$ and $\\langle s_z\\rangle\\rightarrow 0$. \n\nIn order to characterize the DPT, we introduce the time-averaged HZ entanglement parameter\n\\begin{equation}\n\t\\bar{E}_{\\rm HZ}=\\frac{1}{T_{R}}\\int_{0}^{T_R} E_{\\rm HZ}(t)dt .\n\t\\label{HZ-t}\n\\end{equation}\nOur results for the entropic entanglement measure $1-\\bar{E}$ and the correlation-based HZ entanglement parameter $\\bar{E}_{\\rm HZ}$ are depicted in Fig.~\\ref{fig:HJ-mean}, showing that the HZ measure is an excellent proxy for entropic entanglement measure. The time-averaged HZ entanglement parameter presents a sharp dip in the vicinity of the critical point, which can be used as an indicator for the DPT. Comparing to the entropic measure discussed in Sec.~\\ref{Sec:entropy}, the HZ parameter benefits not only from the experimental feasibility, but also from the fact that the dip is not affected by the total numbers of atoms, and thus can identify the DPT in systems with large particle numbers.\n\n\n\\section{Thermal Effects}\\label{Sec:thermal}\n\nSo far we have studied how the two-mode entanglement can characterize the DPT at zero temperature. In the practical experiments, there are inevitable thermal excitations due to the finite temperature. In general, thermal excitations reduce entanglement because they will cause decoherence and degrade the purity. In the present system, an increasing finite temperature reduces the condensate density $n_{C}$ and consequently changes the interaction energies of the condensate $g_sn_{C}$, as well as the critical perturbation strength $V_{0,{\\rm crit}}$. This may change the dynamics of pseudospin and induce new critical points for DPT. We explore how these new critical points at finite temperatures connect with the two-mode entanglement examined by the HZ entanglement criterion. \n\\begin{figure}[t]\n\t\\centering\n\t\\includegraphics[width=0.4\\textwidth]{EHZ_compare_mean}\n\t\\caption{The time-averaged HZ entanglement parameter $\\bar{E}_{\\rm HZ}$ (blue dashed) calculated as a function of perturbation strength $V_{0}\/V_{0,{\\rm crit}}$, in comparison with the time-averaged entropic entanglement measure $1-\\bar{E}$ with $N=10$ (green solid) and $N=100$ (red solid). In the vicinity of the critical point $V_{0,{\\rm crit}}$, $\\bar{E}_{\\rm HZ}$ shows a discontinuity in the first-order derivative, which signifies the occurrence of DPT. Other parameters are used as those in Fig.~\\ref{fig:entropy}.}\n\t\\label{fig:HJ-mean}\n\\end{figure}\n\nIn order to study the effects of thermal excitations, we solve the quasiparticle spectra by using Hartree-Fock-Bogoliubov theory with Popov approximation~\\cite{pethick2002bose,liao2014spin,chen2015collective,PhysRevA.96.013625}. The general wave function $\\Psi_{\\pm k_{m},\\sigma}$ with $\\sigma=\\uparrow, \\downarrow$ is given by\n\\begin{equation}\n\t\\Psi_{\\pm k_{m},\\sigma}(x,t)=e^{-i\\mu t\\pm ik_{m}x}\\left[ \\Phi_{\\pm k_{m},\\sigma}+\\delta\\Phi_{\\pm k_{m},\\sigma}(x,t) \\right] ,\n\t\\label{general_wave_function}\n\\end{equation}\nwhere $\\mu$ is the chemical potential, $\\Phi_{\\pm k_{m}, \\sigma}$ are the condensate wave functions, and $\\delta\\Phi_{\\pm k_{m},\\sigma}(x,t)$ are the fluctuations with the following form\n\\begin{eqnarray}\n\t\\delta\\Phi_{\\pm k_{m},\\sigma}&=&\\psi_{\\pm k_{m}+q,\\sigma}e^{-i\\omega t}+\\phi_{\\pm k_{m}-q,\\sigma}^{\\dagger}e^{i\\omega t}.\n\\end{eqnarray}\nHere, $q$ is the quasi momentum and $\\omega$ is the frequency. It is convenient to solve the Bogoliubov spectrum by expanding $\\psi_{\\pm k_{m}+q,\\sigma}$ and $\\phi_{\\pm k_{m}-q,\\sigma}$ in the Bloch form~\\cite{li2013superstripes,poon2016quantum} with basis $\\psi_{\\tilde{q}+2 \\ell k_{m},\\sigma}$ and $\\phi_{-\\tilde{q}+2 \\ell k_{m},\\sigma}$, \n\\begin{eqnarray}\n\t\\psi_{\\pm k_{m}+q,\\sigma}&=&\\sum_{\\ell}\\psi_{\\tilde{q}+2 \\ell k_{m},\\sigma} ,\\nonumber\\\\\n\t\\phi_{\\pm k_{m}-q,\\sigma}&=&\\sum_{\\ell}\\phi_{-\\tilde{q}+2 \\ell k_{m},\\sigma} ,\n\\end{eqnarray}\nwhere $\\ell$ is an integer and $|\\tilde{q}|0$ switching to $\\bar{M}=0$ at different temperatures. This observation suggests that the DPT can be characterized by the time-average two-mode entanglement even under finite temperatures. \n\nWe also notice from Fig.~\\ref{fig:EHZ_mean}(b) that at the sharp dip, the lowest value of the parameter $\\bar{E}_{\\rm HZ}$ increases with elevated temperature, indicating that thermal effect is in general detrimental to entanglement. However, since the transition point also shifts with temperature, for a fixed perturbation strength $V_0 < V^{T=0}_{0,{\\rm crit}}$ one may find that a lower value of $\\bar{E}_{\\rm HZ}$ can be obtained with increasing temperature. Physically, this corresponds to the fact that although a finite temperature will inevitably introduce thermal fluctuations and hence degrade entanglement, in certain circumstances it can also bring the system closer to a phase transition point, which features maximal entanglement. \n\n\n\\section{Conclusion}\\label{Sec:conclusion}\n\nIn summary, we have studied the possibility that using two-mode entanglement in the synthetic (i.e., spin) space to characterize the dynamical phase transitions in a BEC with spin-orbit coupling. By adding an additional lattice potential in the system Hamiltonian as perturbation, we show that the time-averaged entropic entanglement reaches a maximal value at the critical values of perturbation strength where the system is driven from dynamic magnetized phase to a dynamic nonmagnetized phase. The sharp peak of entropic entanglement measure can be used to identify the existence of the DPT. On the other hand, this provides inspiration for generating the maximal entanglement between two modes. Then, considering the difficulty of measuring entropic entanglement in experiments, we have also examined another correlation-based entanglement criterion which is more feasible for experimental test. Our results shows that the time-averaged HZ entanglement parameter is an excellent substitute for entropic entanglement measure, which can not only determine the existence, but also qualitatively characterize the extent of entanglement. Furthermore, it can also be used to account for the effects of thermal excitations induced by finite temperatures. We find that the thermal effects will change the critical point of the DPT, which can be revealed by the shift of the sharp dips of the time-averaged HZ entanglement parameter. \nThis work may broaden the understanding of the connection between quantum correlations and dynamical phase transitions in the SOC systems with interactions. \n\nIn the end, we would briefly comment about the experimental feasibility of the study. The synthetic spin-orbit coupling in ultracold atomic gases for pseudospin-1\/2 systems has been realized experimentally~\\cite{lin2011spin, wang2012spin, cheuk2012spin, ji2014experimental}. The periodic perturbation $V_{\\rm ex}$ is simply a lattice potential which can be generated by standing-wave lasers with wave vector $k_{m}$. The time-averaged two-mode entanglement parameter $\\bar{E}_{\\rm HZ}$ can be detected by measuring the evolution of pseudospin variance of the BEC which has been realized in many experiments for quantum squeezing and metrology~\\cite{esteve2008squeezing,riedel2010atom,luo2017deterministic}. Thus, we expect that the system can be readily prepared and investigated with present experimental technique. \n\n\n\\begin{acknowledgements}\n\nWe acknowledge illuminating discussions with X.-J. Liu, T. F. Jeffrey Poon, and H. Zhai. This work is supported by the Ministry of Science and Technology of China (Grant No. 2016YFA0301302), National Natural Science Foundation of China (Grants No. 11622428, No. 11274025, No. 61475006, No. 11434011, No. 11522436, and No. 11774425), the Research Funds of Renmin University of China (Grants No. 10XNL016 and No. 16XNLQ03), and the National 973 Program (Grant No. 2014CB921403).\n\n\\end{acknowledgements}\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\n\\label{intro}\nIn recent years, deep neural networks have been shown to be highly effective in solving many real world problems, creating an increased demand for their deployment in a variety of applications. However, such deep neural networks often require a large amount of computational resources for both training and inference purposes, which limits the adoption and spread of this technology in scenarios where large computational resources are not available. This issue is likely to persist in the near future, since recent trends have demonstrated that increased model size is often correlated with improved performance \\cite{GPT3, tan2019efficientnet, BiT}. To mitigate this issue, recent efforts have been focused on developing specialized hardware to support the computational demands \\cite{hardware_survey} as well as model compression methods in order to reduce them \\cite{compression_survey}. These include broad families of techniques such as pruning \\cite{pruning_survey}, knowledge distillation \\cite{distillation_survey}, neural architecture search (NAS)\\cite{nas_survey} and as is the focus of this paper, quantization \\cite{hubara2017quantized}.\n\nQuantization methods enable the computations performed by neural networks to be carried out with fixed point operations, rather than floating point arithmetic.\nThis improves the computational efficiency of neural networks and reduces their memory requirements. However, as with other compression methods, this typically comes at the cost of reduced performance \\cite{compression_survey}. Recent efforts in the field have focused on improving the trade-offs between model compression and performance, by proposing a plethora of quantization schemes tailored for different scenarios.\n\n\\subsection{Quantization Methods}\nQuantization schemes can first be divided into post training and quantization aware training schemes. Post training schemes decouple the tasks of model training and the quantization of its weights and\/or activations. These methods are most suitable when the training data is no longer available when compressing the network, and only the trained network is available \\cite{soudry1, postq1}. On the other hand, quantization aware training schemes perform both optimization tasks together, and do require training data, which tends to provide better performance.\n\nIn addition, quantization schemes can be further divided into methods which perform uniform or non-uniform quantization. Uniform quantization divides the real-valued domain into equally sized bins, whereas non-uniform methods do so with bins of varying sizes. The former method tends to be more resource efficient and compatible with modern hardware accelerators, while the latter tends to provide better model performance \\cite{li2019additive, jung2019learning}.\n\nQuantization schemes can be further subdivided into per-channel methods, which utilize different quantization parameters for different channels within each layer, as opposed to per-layer methods, which use the same parameters for all channels in each layer, but different parameters for different layers. However, as with non-uniform quantization, per-channel methods can be difficult to implement efficiently on standard hardware components, making per-layer methods more feasible for deployment on edge devices.\n\nFor the reasons stated above, in this work we focus on per-layer, uniform quantization aware training with a different number of bits allocated per layer, referred to as dynamic quantization.\n\\subsection{Dynamic Quantization}\\label{dynamicmotivation}\n\\begin{figure}\n\\vskip 0.1in\n\\begin{center}\n\\centerline{\\includegraphics[width=\\columnwidth]{Figure_1}}\n\\caption{Our proposed training scheme, which uses iterative optimization of the model weights and bit allocation. Given a fixed allocation of bits for each layer, the weights are optimised using a quantization aware training procedure. Then, the bit allocation is optimized for those fixed weights and the process is repeated.}\n\\label{fig:OptScheme}\n\\end{center}\n\\vskip -0.1in\n\\end{figure}\nDynamic quantization aims to further improve the trade-offs between performance and computational requirements of QNNs for deployment on specialized edge devices. However, the proper allocation of bits between layers is combinatorial in nature and is hard to optimize. Recent efforts to tackle this problem include the use of reinforcement learning \\cite{elthakeb2018releq, wang2019haq}, Hessian analysis \\cite{dong2019hawqv1, dong2019hawqv2}, quantizer parametrization \\cite{mpd}, and differentiable NAS approaches \\cite{dnas_dynamic, li2020efficient, edmips}. Among these methods, only the NAS approaches account for dependencies between bit allocations in different layers, by forming a super network that includes multiple branches for each precision at each layer. NAS approaches, however, are more expensive to train due to the multiple network branches. Here we offer an alternative approach to achieve the same goal.\n\n\\subsection{Our Contributions}\nIn this work we wish to optimize for the bit allocation of the network as a whole, and address the full interactions between different components of QNNs. To this end, we propose to utilize gradient-free optimization algorithms for the bit allocation. Such algorithms are known to perform well in difficult scenarios with complex dependencies between variables, while maintaining excellent sample efficiency during optimization \\cite{conn2009introduction, rios2013derivative}. In particular, we use the algorithm Covariance Matrix Adaptation Evolution Strategy (CMA-ES) \\cite{hansen2003reducing}, which aims to achieve the mentioned goals, though other alternatives may also be suitable for the task.\n\nWe propose a novel quantization aware training procedure for dynamic QNNs. In this procedure, CMA-ES is used for computing the optimal (per-layer) dynamic bit allocation for the weights and activations of the network. This method is interchangeably applied with gradient-based methods. That is, the network weights are updated by a gradient-based method, while the bit allocation is updated using the gradient-free method CMA-ES---see Fig. \\ref{fig:OptScheme}.\n\nThe advantages of our approach are as follows:\n\\begin{itemize}\n\\item Our training scheme for the dynamic bit allocation optimizes for the network as a whole. That is, it takes into account dependencies between layers of the neural network, as well as dependencies between weights and their bit allocation.\n\\item Our method for bit allocation is gradient-free, hence it can handle multiple (possibly non-differentiable) hardware constraints. This enables the QNN to be tailored to the resources of the specific edge devices it is to be deployed on.\n\\item The systematic combination of gradient-based and gradient-free optimization algorithms can be utilized in other applications and scenarios, e.g. systematic search of the network's other hyper-parameters.\n\\end{itemize}\n\n\\section{Preliminaries}\n\\subsection{Quantization Aware Training}\\label{qat}\nThe quantization schemes of weights or activations in neural networks that we consider here can typically be broken into three steps: clip, scale and round. In the first step, the real values of the weights or activations are clipped to be within a range of values: i.e., $[-\\alpha, \\alpha]$ for symmetric signed quantization (typically weights) and $[0, \\alpha]$ for unsigned quantization (typically activations, outputs of ReLU operations). Then, the range is mapped to the target integer range $[-2^{b-1} + 1, 2^{b-1} - 1]$ and $[0, 2^{b} - 1]$, for signed and unsigned quantization respectively, where $b$ is the number of bits. The values $\\alpha$ are referred to as ``scales'' or clipping parameters and take on different values for different layers. The clipping parameters are optimized throughout the training procedure (along with the weights), so that they yield the lowest possible generalization error when quantization applied to the models' weights and\/or activations.\n\nSeveral approaches have been proposed in recent years, in order to optimize for the scales. In \\cite{zhang2018lq} the scales are chosen based on quantization error minimization. \\cite{choi2018pact} proposes differentiable clipping parameters which are learned through backpropagation together with the model's weights. This approach is further improved in \\cite{li2019additive}, which also supports a computationally efficient \\emph{non-uniform} quantization scheme. In this work we base our uniform quantization aware training scheme (for the weights and activations) on \\cite{li2019additive}, while interchangeably utilizing CMA-ES for the dynamic bit allocation.\n\n\\begin{figure}\n\\vskip 0.1in\n\\begin{center}\n\\centerline{\\includegraphics[width=7cm, height=3cm]{Quantized_tensor}}\n\\caption{Example of a signed 4-bit quantized weights tensor and $\\alpha=0.16$ (original-left, quantized-right). Values which are lesser or greater than the clipping parameters $\\pm \\alpha$ are clipped to $-\\alpha$ or $+\\alpha$ respectively, then scaled and rounded to the signed integer range. The value of 0 is preserved in this process.}\n\\label{fig:quantized_weights_example}\n\\end{center}\n\\vskip -0.1in\n\\end{figure}\n\nTo formally introduce the point-wise quantization operations we first define the quantization operator:\n\\begin{equation} \\label{quantoperator}\n Q_b(x) = \\frac{\\mbox{round}((2^b - 1) \\cdot x)}{2^b - 1},\n\\end{equation}\nwhere $x$ is a real-valued tensor in [-1, 1] or [0, 1] for signed or unsigned quantization, respectively. $b$ is the number of bits that are used to represent $x$.\nGiven this operator, we use the reparameterized clipping function \\cite{li2019additive} to define the quantized weights and activations.\n\\begin{eqnarray} \\label{quantweights}\n W_b &=& \\alpha_W Q_{b-1}(\\mbox{clip}(\\frac{W}{\\alpha_W}, -1, 1))\\\\\n X_b &=& \\alpha_X Q_{b}(\\mbox{clip}(\\frac{X}{\\alpha_X}, 0, 1)) \\label{quantacts}.\n\\end{eqnarray}\nHere, $W, W_b$ are the real-valued and quantized weight tensors, $X, X_b$ are the real-valued and quantized input tensors, and $\\alpha_W, \\alpha_X$ are their associated scale (or clipping) parameters, respectively.\nAn example of Eq. \\eqref{quantweights} being applied to a weights tensor, using 4-bit signed quantization (without multiplying by $\\alpha_W$), can be seen in Figure \\ref{fig:quantized_weights_example}.\n\nWe note that equations \\eqref{quantoperator}-\\eqref{quantacts} are used for training only. During inference, weights and activations are dynamically quantized, and all operations are performed using integers in mixed precision, while taking the scales into account.\n\nIn this quantization aware training scheme, both weights and activations are quantized during the forward pass, and during the backward pass, the Straight Through Estimator (STE) \\cite{STE} is used to optimize the weights. That is, we ignore the derivative of $Q_b$ in the backward pass, and iterate on the floating point values of $W$ using the derivatives w.r.t $W_b$ in the SGD optimization. Given Eq. \\eqref{quantweights}-\\eqref{quantacts} the STE can also be used to calculate the gradients with respect to $\\alpha_W, \\alpha_X$ \\cite{li2019additive}.\nThis enables the quantized network to be trained in an end-to-end manner using backpropogation.\nTo improve stability, \\cite{li2019additive} also use weight normalization before quantization:\n\\begin{equation} \\label{eq:normalize}\n \\hat W = \\frac{W - \\mu}{\\sigma + \\epsilon}.\n\\end{equation}\nHere, $\\mu$ and $\\sigma$ are the mean and standard deviation of the weight tensor, respectively, and $\\epsilon=10^{-6}$.\n\n\\subsection{CMA-ES}\\label{prelim_cma}\nCovariance Matrix Adaptation Evolution Strategies (CMA-ES) \\cite{hansen2003reducing}, is a population based gradient-free optimization algorithm. It is known to be highly versatile and has been applied to a large variety of settings, such as RL \\cite{heidrich2008evolution}, placement of wave energy converters \\cite{neshat2019ahybrid}, hyper-parameter optimization \\cite{loshchilov2016cma}, evolving game levels in Mario \\cite{volz2018evolving}, adversarial attacks \\cite{8917642} and more. It is designed to work in $d$-dimensional continuous spaces and optimize discontinuous, ill-conditioned and non-separable objective functions, in a black box optimization setting \\cite{cma_bbob}.\n\nAt a high level, the optimization process of CMA-ES is as follows. At the $g$-th generation, a set of $\\lambda$ $d$-dimensional samples are drawn from a multivariate normal distribution $\\mathcal{N} (m^{(g)}, \\mathcal{C}^{(g)})$:\n\\begin{equation}\\label{cmaes}\n x_k^{(g+1)} \\sim m^{(g)} + \\sigma^{(g)} \\mathcal{N} (0, \\mathcal{C}^{(g)}), \\mbox{ for }k = 1,...,\\lambda\n\\end{equation}\nWhere $m^{(g)}$, $\\mathcal{C}^{(g)}$ are the mean and covariance matrrix of the population at the previous generation, respectively. $\\lambda$ is the population size and $\\sigma^{(g)}$ is the step-size.\n\nOnce the samples are drawn from this distribution, they are evaluated and ranked based on the objective function.\nThen, keeping only the top $\\mu$ samples with the best objective values, $m^{(g+1)}$, $\\mathcal{C}^{(g+1)}$ and $\\sigma^{(g+1)}$ are calculated using a set of update rules. Please refer to \\cite{hansen2016cma} for more details.\nThese are then used to draw samples for the next generation using Eq. \\eqref{cmaes}.\nThis process is repeated until one of several convergence criteria are full-filled, or until the algorithm exceeds its predefined budget of allowed samples.\n\nThough CMA-ES is an effective gradient-free optimization algorithm, it has several downsides. First, its computational complexity is $O(d^2)$ in space and time \\cite{hansen2003reducing}, where $d$ is the dimension of the parameter vector to be optimized.\nSecond, its convergence rate is often also $O(d^2)$ \\cite{cma_bbob}, making it inefficient for solving high dimensional problems where $d$ is larger than a few thousands.\n\n\\section{The GradFreeBits Method}\\label{method}\nIn this section we describe the components of our optimization procedure for dynamic QNNs.\n\n\\subsection{Motivation: CMA-ES for Dynamic Precision.}\nWe argue that CMA-ES is highly compatible with the problem of bit-allocation in QNNs. Here we assume the objective function is the differentiable loss function used during training, with additional possibly non-differentiable constraints related to computational requirements (exact details are provided below).\n\nAs recent evidence suggests \\cite{wang2019haq, dong2019hawqv1}, the optimization landscape of the bit-allocation problem is likely to be discontinuous, ill-conditioned and amenable for optimization using gradient-free optimizers. Since the constraints may be non-differentiable, they can be sampled in a black-box setting, as is done in gradient-free optimization (CMA-ES) and reinforcement learning \\cite{wang2019haq}.\nAdditionally, as shown in \\cite{dong2019hawqv1, dong2019hawqv2}, the Hessian eigenvalues show large variations for different layers in QNNs, meaning that certain layers are typically more sensitive to changes in bit-allocation than others. This is in part what motivated us to choose CMA-ES for this work, as it is capable of adapting to high variations in the Hessian eigenvalues, and is therefore considered to be one of the best and widely used gradient-free methods.\nLastly, since neural networks typically have an order of tens to hundreds of layers, the problem of bit optimization falls well within the range of dimensionallity where CMA-ES is known to perform well, as discussed in the previous section.\n\n\n\\subsection{Setting the Stage for CMA-ES}\\label{bitaloc}\nTo enable the gradient-free optimization algorithm to efficiently optimize the bit-allocation of a QNN, two items must be defined: its search space and the objective function.\n\n\\paragraph{Search Space}\nWe define the search space as a vector containing the bit allocation of the weights and activations in all the layers of the network, asides the first and the last layer, as these are quantized using a fixed allocation of 8 bits.\nWe found it beneficial to optimize the logarithm (base 2) of this vector, rather than the vector itself.\nThus, the vector to be optimized by CMA-ES is the log-precision vector:\n\\begin{equation}\\label{cma_search_space}\n \\mathbf{v} = [\\mathbf{v_W}, \\mathbf{v_X}] = \\log_2 ([\\mathbf{r_W}, \\mathbf{r_X}]),\n\\end{equation}\nwhere $\\mathbf{r_W}, \\mathbf{r_X}$ are the bit allocations of the network's weights and activations respectively, and $[\\cdot, \\cdot]$ is the concatenation operation.\n\n\\paragraph{Objective Function}\nAs mentioned in Section \\ref{prelim_cma}, the objective function to minimize is a combination of the network's performance measure, the differentiable loss function, subject to a number of non-differrentiable computational constraints. More formally:\n\\begin{equation}\n \\min_\\mathbf{v} \\textit{ } \\mathcal{L}(\\mathbf{v}; \\mathbf{\\theta}), \\\\\n \\hspace{4pt} \\mbox{ s.t. } h_j(\\mathbf{v}) \\leq C_j \\hspace{4pt} \\mbox{ for } j=1,...,M\n\\end{equation}\nWhere $\\mathcal{L}(\\mathbf{v}; \\mathbf{\\theta})$ is the loss function over the training set, parameterized by network parameters $\\mathbf{\\theta}$, which are assumed to be fixed during the bit-allocation optimization stage.\nFurthermore, $h_j(\\mathbf{v})$ are the computational requirements for a given precision vector $\\mathbf{v}$ (e.g., model size, inference time etc.), $C_j$'s are the target requirements that we wish to achieve, and $M$ is the number of constraints.\nIn our framework, we combine the constraints into the objective function using the penalty method \\cite{penalty}:\n\\begin{equation}\\label{objective}\n \\min_\\mathbf{v} \\textit{ } \\mathcal{L}(\\mathbf{v}; \\mathbf{\\theta})\n + \\sum_{j=1}^M \\rho_j \\max(0, h_j(\\mathbf{v}) - C_j)^2,\n\\end{equation}\nWhere $\\rho_j$ are balancing constraint parameters. This is similar to the approach taken in \\cite{mpd}, but here it is applied to gradient-free optimization.\n\nWe define the computational constraints by comparing the requirements between the dynamic and static precision networks. For example, we wish that the model size will be equal to that of a static 4 bits allocation.\nSpecifically, we use constraints on the model size for the weights entries $\\mathbf{v_W}$ and the mean bit allocation for the activation entries $\\mathbf{v_X}$:\n\\begin{equation} \\label{constraint1}\n\t h_1({\\mathbf{v}}) = MB({\\mathbf{v_W}}), \\hspace{4pt} C_1 = \\beta_1 MB({\\mathbf{v}^S_{\\textbf{W}}})\n\\end{equation}\n\\begin{equation} \\label{constraint2}\n\t h_2({\\mathbf{v}}) =\n\t \\frac{1}{L} \\sum_{i=1}^L v_{X_i}, \\hspace{4pt} C_2 = \\frac{\\beta_2}{L} \\sum_{i=1}^L v^S_{X_i}\n\\end{equation}\nWhere $MB(\\cdot)$ calculates the model size given weight entries $\\mathbf{v_W}$, $\\mathbf{v}^S$ is the log-precision vector of the target static precision we wish to achieve, $L$ is the number of relevant layers, and $\\beta_1, \\beta_2>0$ control the target compression rates of the weights and activations, respectively.\n\n\nThese constraints are designed to limit the computational requirements, while allowing the gradient-free optimization algorithm to explore non-trivial solutions which satisfy them.\nIt is important to note that though these constraints are only related to memory requirements, other constraints can easily be incorporated into our framework - such as power usage or inference time measurements, chip area, etc.\n\n\\subsection{Variance Reduction}\\label{variance_reduction}\nVariance reduction has been shown to improve the convergence rate of optimization algorithms \\cite{variance_reduction}. The main source of variance in our objective function (Eq. \\eqref{objective}) is in the first term, related to the performance of the model for different bit-allocations. There are two main causes of variance in this term: sub-sampling noise, caused by using small mini-batches of randomly selected samples, and sensitivity to quantization errors, which networks are typically not robust to.\nTo mitigate these issues and improve the convergence rate of CMA-ES we adopt two strategies for reducing the variance in the objective function, utilizing moving super-batches and quantization aware pretraining, which will be described promptly.\n\n\\begin{figure}\n\\vskip 0.1in\n\\begin{center}\n\\centerline{\n\\includegraphics[width=7.0cm, height=4.0cm]{Moving_superbatch}}\n\\caption{Batch replacement scheme used in moving super-batches, as compared to standard mini-batch replacement. At each iteration a single mini-batch is replaced, in a queue-like manner, creating a high overlap of samples between consecutive super-batches.}\n\\label{fig:moving_superbatch}\n\\end{center}\n\\vskip -0.1in\n\\end{figure}\n\n\\paragraph{Moving Super-batches}\\label{moving_superbatch_section}\nWe define a moving super-batch as a set of mini-batches which are replaced in a queue-like manner. That is, in each iteration of the super-batch we replace part of the mini-batches within it. During each objective evaluation of CMA-ES, the entire super-batch is run through the model in order to calculate the first term of \\eqref{objective}. The queue-like replacement scheme enables CMA-ES to encounter new data samples in each objective evaluation, but with a larger overlap of data samples as compared to standard SGD, where the entire mini-batch is re-sampled at each iteration. The process is illustrated in Figure \\ref{fig:moving_superbatch}. Several strategies for the frequency of replacement can be considered, such as replacing one or more mini-batches after each objective evaluation, or doing so every fixed number of evaluations. These different settings are explored in the ablation study in Section \\ref{ablation_study_section}.\n\n\\paragraph{Pretraining}\\label{pretraining}\nAnother way to reduce variability in performance, is by reducing sensitivity to quantization errors. This can be done by pretraining the network, using the quantization aware training scheme, as described in section \\ref{qat}. The bit allocation for this stage is static, meaning the weights and activations in all conv layers are quantized to the same target precision, except the first conv and last full-connected layers, which are always quantized to 8-bit precision.\n\n\\subsection{Gradient-free Steps}\nWe define gradient-free steps, as steps in which the CMA-ES algorithm optimizes the bit allocation (given the network's weights) according to the objective function in Eq. \\eqref{objective}. In each step, multiple generations of samples of the log-precision vector $\\mathbf{v}$ are evaluated on \\eqref{objective}. At each objective evaluation, the sample of $\\mathbf{v}$ is used to quantize the weights and activations of the model. Since CMA-ES typically operates in continuous spaces, the bit allocations of the weights and activations (which are positive integers) are extracted from $\\mathbf{v}$ using the following formula:\n\\begin{equation} \\label{extracting_precision}\n \\mathbf{r_W} = \\ceil{2^{\\mathbf{v_W}}}, \\mathbf{r_X} = \\ceil{2^{\\mathbf{v_X}}}\n\\end{equation}\nWhere $\\ceil{x}$ is the \"ceil\" operator, which rounds its argument upwards to the nearest integer. Once the bit allocations have been inserted into the model, the loss \\eqref{objective} is calculated over all the mini-batches in the super-batch, yielding the value of the objective for each of the sampled bit-allocations. This process gradually minimizes the value of the objective function and enables non-trivial bit-allocations to be found. We define each gradient-free step to include a predefined number of objective evaluations. (this a hyper-parater, typically set to 512 evaluations). It is important to note that gradient-free steps require significantly less computational resources than traditional epochs, even if the number of mini-batches are matched. This is because they do not require backpropogation and the gradient-free optimizer computations are negligible compared to those performed by the model.\n\n\n\\subsection{Iterative Alternating Retraining}\nThe primary innovation in this work is the combination of gradient-based training sessions of the model weights and gradient-free training sessions of the bit-allocation, which we refer to as iterative alternating retraining.\n\nAfter the pretraining stage, the model is passed to the gradient-free optimizer CMA-ES to optimize its bit allocation for a session of a predefined number of steps $N_{GF}$. This adapts the bit allocation to the model weights, which are fixed at this stage in their floating point values. This maximizes the performance of quantized networks, subject to the computational constraints (Eq. \\eqref{objective}), using the optimization process as described in section \\ref{prelim_cma}.\n\nOnce the gradient-free session is completed, the best bit allocation found by CMA-ES is passed to the gradient-based optimizer, for a session of a predefined number of epochs $N_{GB}$. This adapts the model weights to the bit allocation, which is kept fixed, using the quantization aware training scheme described in Section \\ref{qat}. Once this session is completed, the model is returned to the gradient-free optimizer to further improve the bit allocation, given the new model weights. This cycle is then repeated a number of times, or until the performance and computational requirements are satisfactory. The process is illustrated in Fig. \\ref{fig:OptScheme}.\n\nTo increase stability in the handover process between the optimization algorithms, we found it beneficial to restart CMA-ES after each gradient-based session and handover only the best model weights and bit allocation (with lowest loss\/objective values) found across all sessions, as opposed to handing over the best versions found in each session.\n\n\\section{Experiments}\n\n\\begin{figure}\n\\vskip 0.1in\n\\begin{center}\n\\centerline{\n\\includegraphics[width=\\columnwidth, height=3.8cm]{Res56_cifar10}\n}\n\\caption{(a) Convergence plot for a 3 and 4-bit dynamically quantized ResNet56 on CIFAR10, during the retraining process. Grey regions correspond to gradient-free steps, while white regions correspond to gradient-based epochs. (b) The bit allocations for the weights (blue, left) and activations (green, right) of the 4-bit ResNet56.}\n\\label{fig:res56_cifar10}\n\\end{center}\n\\vskip -0.1in\n\\end{figure}\n\nTo quantitatively compare our dynamic quantization scheme (GradFreeBits) to other related works, we apply it to several neural network architectures for image classification tasks. Throughout all the experiments, we consider similar numbers of bits (on average, in the dynamic case) for the weights and activations. Since we use the ReLU activation in all the settings, the activations are quantized using unsigned integers. The weights are quantized symmetrically using signed integers, so for example, 2-bit quantization will result in ternary weights in $\\{-1,0,1\\}$.\n\nFurthermore, in this section we also perform an ablation study, to identify the effects that different settings of the proposed system have on the performance metrics of the resulting dynamically quantized models.\n\nWe compare our approach to the following related works that use uniform quantization, with either static (S) or dynamic (D) bit-allocation schemes:\nLQ-Nets(S) \\cite{zhang2018lq}, PACT(S) \\cite{choi2018pact}, APoT(S) \\cite{li2019additive}, DSQ(S) \\cite{gong2019differentiable}, BSGD(S) \\cite{bcgd}, TQT(D) \\cite{tqt}, MPD(D) \\cite{mpd}, EBS(D)\\cite{li2020efficient}, EDMIPS(D) \\cite{edmips}, HAQ(D) \\cite{wang2019haq}, HAWQ-V1(D) \\cite{dong2019hawqv1}, HAWQ-V2(D) \\cite{dong2019hawqv2}, WNQ(S) \\cite{wnq}, RES(S) \\cite{res}, DoReFa-Net(S) \\cite{drfn}. Lastly, we denote our our proposed method GradFreeBits as GFB(D).\n\n\n\\subsection{CIFAR 10\/100} \\label{cifar10_section}\nThe CIFAR10 and CIFAR100 image classification benchmarks \\cite{cifar10} have 10 and 100 classes respectively, both containing 50k train and 10k test 32x32 RGB images. For these datasets, pretraining \\textit{from scratch} was conducted for 400 epochs, using a mini-batch size of 128, SGD optimizer with momentum of 0.9, cosine learning rate decay, with a maximum learning rate of 0.1 and warm-up of 10 epochs. Data augmentation used during the pretraining and iterative retraining stages included: random horizontal flips and crops, and mixup \\cite{mixup}. To approximately satisfy the constraints of \\eqref{objective}, we chose $\\beta_1=\\beta_2=0.7$ and $\\rho_1=\\rho_2=0.5$. This allows the penalty constraints to be slightly violated, while keeping model size and mean precision below their actual targets. We also limit the search space, such that all values of $\\textbf{v}$ are in $[0, 3]$.\n\nGiven the pretrained model, we applied 5\/3 rounds of the iterative alternating retraining algorithm for CIFAR10\/100 respectively. In each round, we applied 4 steps in the gradient-free session and 4\/16 epochs in the gradient-based session for CIFAR10\/100 respectively. At each gradient-free step of CMA-ES we applied 512 objective evaluations on super-batches, each containing 32 mini-batches of 128 data samples. After each of the 512 objective evaluations we swap only one mini-batch in the super-batch as illustrated in Fig. \\ref{moving_superbatch_section}.\n\n\n\n\n\n\\begin{table}[t]\n\\caption{Top1 accuracy of quantized network on CIFAR10\/100. (S) and (D) denote static and dynamic quantization, while \\textbf{*} denotes experiments which use 2-bit weights and 4-bit activations.}\n\\label{tab:cifar10}\n\\vskip 0.15in\n\\begin{center}\n\\begin{small}\n\\begin{sc}\n\\begin{tabular}{lccr}\n\\toprule\nModel & Method & 3\/3 & 4\/4 \\\\\n\\midrule\n\nRes.20 & PACT(S) & 91.1 & 91.7 \\\\\nCIFAR10 & LQ-Nets(S) & 91.6 & - \\\\\nFP 93.3 & BCGD(S) & 91.2 * & 92.0 \\\\\n & TQT(D) & 90.4 * & - \\\\\n & MPD(D) & 91.4 * & - \\\\\n & EBS(D) & 92.7 & 92.9 \\\\\n \\rowcolor{gray!20}\n & GFB(D) (ours) & \\textbf{93.3} & \\textbf{93.6} \\\\\n \\rowcolor{gray!20}\n & GFB(D) (ours) & \\textbf{93.4}* & - \\\\\n\\hline\n\\hline\n\nRes.56 CIFAR10 & EBS(D) & 94.1 & 94.3 \\\\\n\\rowcolor{gray!20}\nFP 95.1 & GFB(D) (ours) & \\textbf{94.8} & \\textbf{94.9} \\\\\n\\hline\n\\hline\n\nRes.20 & DoReFa-Net(S) & 68.4 & 68.9 \\\\\nCIFAR100 & Res(S) & 68.3 & 68.7 \\\\\nFP 70.35 & LQ-Nets(S) & 68.4 & 69.0 \\\\\n & WNQ(S) & 68.8 & 69.0 \\\\\n \\rowcolor{gray!20}\n & GFB(D) (ours) & \\textbf{69.6} & \\textbf{70.6} \\\\\n\\bottomrule\n\\end{tabular}\n\\end{sc}\n\\end{small}\n\\end{center}\n\\vskip -0.1in\n\\end{table}\n\nThe results for the CIFAR10 dataset are presented in Table \\ref{tab:cifar10}. Our method out-performs the previous state of the art EBS(D) for both ResNet20 and ResNet56 models, at both dynamic precision settings: e.g. +0.7\\% for 4-bit ResNet20 and +0.6\\% for 4-bit ResNet56. Table \\ref{tab:cifar10} also includes the results for the CIFAR100 dataset. For this benchmark, our method also outperforms all other related works: by 1.6 for 4-bit ResNet20 and 0.8 for fore 3-bit ResNet20.\n\n\n\n\n\\subsection{ImageNet}\\label{imagenet_section}\nThe ImageNet \\cite{imagenet} image classification benchmark has 1K classes, 1.2M train and 150K test RGB images. For all models, we used pretrained weights from TorchVision \\cite{torchvision}.\nAdditional quantization aware pretraining was then conducted for 30 epochs, using mini-batches of size 128, SGD optimizer with momentum of 0.9, and a cosine learning rate decay with a maximum of 0.001 and a warm-up of 3 epochs.\n\nData augmentations are identical to those used in the CIFAR10\/100 experiments (above), with the addition of image resize to $256 \\times 256$ and random crops to $224 \\times 224$. These are used for both pretraining and iterative alternating retraining stages. Furthermore, $\\beta_1=\\beta_2=0.7$, $\\rho_1=\\rho_2=0.5$ are used to satisfy the constraints and the search space for $\\textbf{v}$ is limited to $[0, 3]$. Three rounds of iterative alternating retraining were applied. In each round, we applied 4 steps in the gradient-free session and 4 epochs in the gradient-based session.\nIn each gradient-free step of CMA-ES we applied 512 objective evaluations on super-batches, each containing 8 mini-batches of 128 data samples, swapping one-minibatch in the super-batch after each objective evaluation.\nThe entire procedure requires objective evaluations that are comparable to 47 epochs (30 gradient-based pretraining epochs, 12 gradient-free steps, and 12 gradient-based epochs).\n\\begin{table}[t]\n\\caption{Top1 accuracy of quantized networks on ImageNet. (S) and (D) denote static and dynamic quantization, ($\\cdot$) denotes model size, measured in MB. $\\approx$ is used when only graphical results were available for comparison. \\textbf{*} denotes experiments which use 2-bit weights and 4-bit activations.}\n\\label{tab:ImageNet}\n\\vskip 0.15in\n\\begin{center}\n\\begin{small}\n\\begin{sc}\n\\resizebox{\\columnwidth}{!}{%\n\\begin{tabular}{lccr}\n\\toprule\nModel & Method & 3\/3 & 4\/4 \\\\\n\\midrule\nRes.18 & BCGD(S) & - & 67.4(-) \\\\\nFP 70.7 & LQ-Nets(S) & 68.2(6.1) & 69.3(7.4) \\\\\n(46.8) & DSQ(S) & 68.7(6.1) & 69.6(7.4) \\\\\n & PACT(S) & 68.1(6.1) & 69.2(7.4) \\\\\n & APoT(S) & 69.4(4.6) & - \\\\\n & EBS(D) & 69.5(-) & 70.2(-) \\\\\n & EDMIPS(D) & $\\approx$67(-) & $\\approx$68(-) \\\\\n & TQT(D) & - & 69.5(5.6) \\\\\n & MPD(D) & - & 70.1(\\textbf{5.4}) \\\\\n \\rowcolor{gray!20}\n \n & GFB(D) (ours) & 69.2(\\textbf{3.6}) & \\textbf{70.3}(\\textbf{5.4}) \\\\\n\n\\hline \\hline\nRes.50 & DoReFa-Net(S) & 69.9(16.6) & 71.4(19.4) \\\\\nFP 76.4 & LQ-Nets(S) & 74.2(16.6) & 75.1(19.4) \\\\\n(97.5) & PACT(S) & 75.3(16.6) & \\textbf{76.5}(19.4) \\\\\n & EDMIPS(D) & $\\approx$72.5(-) & $\\approx$74.0(-) \\\\\n & HAQ(D) & 75.4(9.2) & 76.1(\\textbf{12.1}) \\\\\n & HAWQ-V1(D) & 75.5(\\textbf{8.0}) \\textbf{*} & - \\\\\n & HAWQ-V2(D) & \\textbf{75.8}(\\textbf{8.0}) \\textbf{*} & - \\\\\n \\rowcolor{gray!20}\n & GFB(D) (ours) & 75.7(9.6) & 76.1(12.8) \\\\\n\\bottomrule\n\\end{tabular}}\n\\end{sc}\n\\end{small}\n\\end{center}\n\\vskip -0.1in\n\\end{table}\n\n\n\\noindent The results for the ImageNet dataset are presented in Table \\ref{tab:ImageNet}. For the Renet18 model, our method slightly outperforms the previous state of the art EBW(D)\\cite{li2020efficient} for 4-bits dynamic precision $+0.1\\%$, though slightly under-performs for 3-bits $-0.3\\%$. However, both dynamically quantized models achieve the smallest model size in their category.\n\n\nFor the 3-bit weights and 3-bit ResNet50, our method achieves similar Top1 accuracy to that of HAWQ2(D) \\cite{dong2019hawqv2} with 2-bit weights and 4-bit activations, though with a larger model size $+1.6MB$. However, our 4-bit ResNet50 falls behind PACT(S) \\cite{choi2018pact} in performance $-0.4\\%$, though with a significantly smaller model size $-6.6MB$. This is on par with the previous state of the art method HAQ(D) \\cite{wang2019haq}, even though it uses \\emph{non-uniform}, quantization which can achieve a higher accuracy than uniform.\n\nFor these experiments we can observe the advantage of the dynamically quantized models compared to the statically quantized ones. Even though the dynamically quantized models require less memory, they tend to obtain higher or comparable Top1 accuracy scores.\n\n\\subsection{Ablation Study}\\label{ablation_study_section}\nIn this section we examine the effects that different system settings have on the performance of our dynamically quantized models. More specifically, we examine the effects of pretraining, iterative alternating retraining, the different settings of moving super-batches and the number of mini-batches they contain. To enable efficient comparison, all experiments are conducted with 4-bit dynamically quantized ResNet20 models on CIFAR100 with the same hyperparameters used in \\ref{cifar10_section}. (learning rate, number of gradient-based\/free epochs, etc.).\n\nWe consider the following settings for replacing mini-batches in the super-batch with new batches of randomly selected training samples: \\textbf{NR} - no mini-batches are replaced in the super-batch, \\textbf{EB} - single mini-batch replacement after each gradient-free step, \\textbf{EF} - replacement of all mini-batches after each gradient-free step, \\textbf{SB} - single mini-batch replacement after each objective evaluation, \\textbf{SF} - replacement of all mini-batches after each objective evaluation. The test without iterative alternating retraining runs all gradient-free steps first, followed by all the gradient-based epochs. The test without pretraining, uses the full-precision model as the initial weights.\n\n\n\\begin{table}[t]\n\\caption{Ablation study of a 4-bit dynamically quantized ResNet20 on CIFAR100, for various system settings. We use the shorthand: \"SS.\" for super-batch setting, \"NMB.\" for number of mini-batches in the super-batch, \"IT.-RET.\" for iterative alternating retraining, \"PRET.\" for pretraining and \"ACC.\" for accuracy.}\n\\label{ablation_system_settings}\n\\vskip 0.15in\n\\begin{center}\n\\begin{small}\n\\begin{sc}\n\\begin{tabular}{lccccr}\n\\toprule\nvariable & ss. & nmb. & it.-ret. & pret. & Acc. \\\\\n\\midrule\nBaseline & SB & 32 & $\\surd$ & $\\surd$ & \\textbf{70.61} \\\\\n\\hline\nComponents & SB & 32 & $\\surd$ & $\\times$& 66.99 \\\\\n & SB & 32 & $\\times$& $\\surd$ & 70.43 \\\\\n\\hline\nSuper-batch & NR & 32 & $\\surd$ & $\\surd$ & 70.21 \\\\\nSetting & EB & 32 & $\\surd$ & $\\surd$ & 70.36 \\\\\n & EF & 32 & $\\surd$ & $\\surd$ & 70.26 \\\\\n & SF & 32 & $\\surd$ & $\\surd$ & 70.39 \\\\\n\\hline\nSuper-batch & SB & 4 & $\\surd$ & $\\surd$ & 69.68 \\\\\nSize & SB & 8 & $\\surd$ & $\\surd$ & 70.25 \\\\\n & SB & 16 & $\\surd$ & $\\surd$ & 70.18 \\\\\n& SB & 64 & $\\surd$ & $\\surd$ & 70.26 \\\\\n\\bottomrule\n\\end{tabular}\n\\end{sc}\n\\end{small}\n\\end{center}\n\\vskip -0.1in\n\\end{table}\n\n\n\n\n\\noindent The results of the ablations study are presented in Table \\ref{ablation_system_settings}. The $-3.6\\%$ accuracy degradation demonstrates that pretraining plays a crucial role in reducing performance degradation due to changes in bit-allocation. It also seems that iterative alternating retraining, as opposed to separating the bit-optimization and weight-optimization stages, leads to a small $+0.2\\%$ increase in performance, demonstrating the added value of this approach. Regarding the super-batch settings, it seems that the optimal setting is to use 32 mini-batches and replace a single batch after each objective evaluation (SB), also leading to a performance increase of $+0.2\\%$. For these reasons, we used these settings used in the rest of the experiments.\n\n\\section{Discussion}\nThough the proposed method achieves favorable trade-offs between performance and model size, several items need be taken into account when applying it to new problems.\n\nFirst, since CMA-ES is designed to operate in continuous search spaces and the search space considered here is discrete, rounding operations are required to ensure compatibility (Eq. \\eqref{extracting_precision}). This creates plateaus of constant values in the objective function, which may harm the convergence rate. To tackle this, we may use other gradient-free algorithms, which are more suitable for discrete optimization, and may lead to even better results using our framework.\n\nSecondly, though this work shows that the CMA-ES optimizer is highly effective in a variety of QNNs, applying it to larger networks (such as EfficientNet B7 \\cite{tan2019efficientnet}) may require more generations and objective evaluations in order to achieve the comparable improvements in the performance and compression rate. This is because the convergence rate of CMA-ES is typically $O(L^2)$ (see section (\\ref{prelim_cma})), where $L$ is the number of layers.\nThat said, if these consideration are taken into account, the combination of gradient-based and gradient-free optimization methods can be applied to achieve highly competitive results.\n\n\\section{Conclusion}\nWe proposed GradFreeBits, a novel framework for optimizing the bit-allocation in dynamically quantized neural networks, which enables neural networks to be customized to meet multiple hardware constraints. The framework is based on the combination of a gradient-based quantization aware training scheme for the weights and gradient-free optimization of the bit-allocation, based on CMA-ES. By benchmarking our method on multiple image classification tasks, we find that our method often outperforms several related dynamic and static quantization methods, in both accuracy and model size. We believe our method will help accelerate the deployment of low-precision neural network on resource constrained edge-devices, by enabling the network compression to be tailored to the specific hardware requirements. Additionally, the proposed framework for combining gradient-free and gradient-based optimization in an iterative alternating retraining scheme is quite general, making it likely easy to apply to other applications.\n\nFuture work in this direction includes utilizing additional constraints, such as FLOPS count and measurements from hardware simulators (such as BISMO \\cite{bismo}) - including inference time, power consumption and chip area. Furthermore, it is important to evaluate this method on new tasks such as object detection and semantic segmentation, as these tasks are more difficult than image classification, perhaps making the models more sensitive to changes in bit allocation.\nWe hope this framework will provide a foundation for future works related to the combination of gradient-free and gradient-based neural network training, and are looking forward to more contributions in this direction.\n\n\\small\n\n\\section{Appendix: CMA-ES}\\label{prelim_cma}\n\nIn this section we provide a very brief overview of the main derivations that are used in CMA-ES, without including any theoretical background for the structure of these update rules, or the choices of hyperparameters. For more details regarding these topics, we refer the curious reader to \\cite{hansen2016cma}---we follows the same notation as this paper here. \n\nCovariance Matrix Adaptation Evolution Strategies (CMA-ES) \\cite{hansen2003reducing}, is a population based gradient-free optimization algorithm.\nAt a high level, the optimization process of CMA-ES is as follows. At the $g$-th generation, a set of $\\lambda$ $d$-dimensional samples \n$\\boldsymbol{x}_{k} \\in \\mathbb{R}^d$\nare drawn from a multivariate normal distribution $\\mathcal{N} (m^{(g)}, \\mathcal{C}^{(g)})$:\n\\begin{equation}\\label{cmaes2}\n \\boldsymbol{x}_{k}^{(g+1)} \\sim \\boldsymbol{m}^{(g)}+\\sigma^{(g)} \\mathcal{N}\\left(\\mathbf{0}, \\boldsymbol{C}^{(g)}\\right), \\mbox{ for }k = 1,...,\\lambda\n\\end{equation}\nWhere $\\boldsymbol{m}^{(g)}$, $\\boldsymbol{C}^{(g)}$ are the mean and covariance matrix of the population at the previous generation, respectively. $\\lambda$ is the population size and $\\sigma^{(g)}$ is the step-size. \n\nOnce the samples are drawn from this distribution, they are evaluated and ranked based on their objective function values. These ranked samples $\\boldsymbol{x}_{i: \\lambda}^{(g+1)}$ are used to calculate $m^{(g+1)}$, $\\mathcal{C}^{(g+1)}$ and $\\sigma^{(g+1)}$ of the next generation, using a set of update rules which are provided below. \n\n\n\n\\subsection{Hyperparameters}\n\n\nCMA-ES uses several hyperparameters in order to perform optimization \\cite{hansen2016cma}. These include a dampling parameter $d_{\\sigma}$ and $c_{1}, c_{\\mu}, c_{\\mathrm{c}}, c_{\\mathrm{m}}, c_{\\sigma}$ which are \"momentum\"-like parameters, which control the amount of information retained from previous generations. Furthremore, $w_i$ are known as the \"recombination weights\", which are used in most update rules. They are typically chosen such that\n\n\\begin{equation}\\label{cmaes}\n\\sum_{j=1}^{\\lambda} w_{j} \\approx 0, \\quad \\sum_{i=1}^{\\mu} w_{i}=1, \\quad w_{1} \\geq \\cdots \\geq w_{\\mu}>0.\n\\end{equation}\n\nThese are also used to calculate the effective population size for recombination: $\\mu_{\\mathrm{eff}}=\\left(\\sum_{i=1}^{\\mu} w_{i}^{2}\\right)^{-1}$. For more details regarding the specific choices of these hyparamerameters, please refer to \\cite{hansen2016cma}.\n\n\\subsection{Mean Update Rule}\nAs mentioned above, several update rules are employed in CMA-ES. The first of those is the update of the mean $\\boldsymbol{m}^{(g+1)}$:\n\\begin{equation}\\label{cmaes}\n \\boldsymbol{m}^{(g+1)}=\\boldsymbol{m}^{(g)}+c_{\\mathrm{m}} \\sum_{i=1}^{\\mu} w_{i}\\left(\\boldsymbol{x}_{i: \\lambda}^{(g+1)}-\\boldsymbol{m}^{(g)}\\right).\n\\end{equation}\n\n\\subsection{Covariance Matrix Update Rule}\nThe covariance matrix requires auxiliary vectors in order to construct its update rule, these are calculated using:\n\n\\begin{eqnarray} \\label{auxilary_vectors}\n \\boldsymbol{p}_{\\mathrm{c}}^{(g+1)} &=& \\left(1-c_{\\mathrm{c}}\\right) \\boldsymbol{p}_{\\mathrm{c}}^{(g)} + \\tilde c_{\\mathrm{c}} \\frac{\\boldsymbol{m}^{(g+1)}-\\boldsymbol{m}^{(g)}}{\\sigma^{(g)}}\\\\\n \\boldsymbol{y}_{i: \\lambda}^{(g+1)} &=& \\left(\\boldsymbol{x}_{i: \\lambda}^{(g+1)}-\\boldsymbol{m}^{(g)}\\right) \/ \\sigma^{(g)},\n\\end{eqnarray}\nwhere $\\tilde c_{\\mathrm{c}} = \\sqrt{c_{\\mathrm{c}}\\left(2-c_{\\mathrm{c}}\\right) \\mu_{\\mathrm{eff}}}$. These auxiliary vectors are then used to construct the rank-$\\mu$ and rank-$1$ update matrices:\n\\begin{eqnarray} \\label{auxilary_matricies}\n \\boldsymbol{C}_{\\mu}^{(g+1)} &=& \\sum_{i=1}^{\\lambda} w_{i} \\boldsymbol{y}_{i: \\lambda}^{(g+1)} \\left(\\boldsymbol{y}_{i: \\lambda}^{(g+1)}\\right)^{\\boldsymbol{\\top}}\\\\\n \\boldsymbol{C}_{1}^{(g+1)} &=& \\boldsymbol{p}_{\\mathrm{c}}^{(g+1)} \\boldsymbol{p}_{\\mathrm{c}}^{(g+1)^{\\top}}.\n\\end{eqnarray}\nFinally, by defining $c_{old} = \\left(1-c_{1}-c_{\\mu} \\sum_{i=1}^{\\lambda} w_{i}\\right)$, we get the update rule for the covariance matrix:\n\\begin{equation}\\label{covariance_matrix_update}\n\\boldsymbol{C}^{(g+1)} = c_{old} \\boldsymbol{C}^{(g)} + c_1 \\boldsymbol{C}_{1}^{(g)} + c_{\\mu} \\boldsymbol{C}_{\\mu}^{(g+1)}.\n\\end{equation}\n\n\n\n\n\\subsection{Step Size Update Rule}\nThe last update rule is for the step size, which also requires use of an auxiliary vector: \n\\begin{equation}\n\\boldsymbol{p}_{\\sigma}^{(g+1)}=\\left(1-c_{\\sigma}\\right) \\boldsymbol{p}_{\\sigma}^{(g)}+\\tilde c_{\\sigma} \\boldsymbol{C}^{(g)^{-\\frac{1}{2}}} \\frac{\\boldsymbol{m}^{(g+1)}-\\boldsymbol{m}^{(g)}}{\\sigma^{(g)}},\n\\end{equation}\nWhere $\\tilde c_{\\sigma} = \\sqrt{c_{\\sigma}\\left(2-c_{\\sigma}\\right) \\mu_{\\mathrm{eff}}}$. This is then used to construct the last update rule, which is for the step size:\n\\begin{equation}\n\\sigma^{(g+1)}=\\sigma^{(g)} \\exp \\left(\\frac{c_{\\sigma}}{d_{\\sigma}}\\left(\\frac{\\left\\|\\boldsymbol{p}_{\\sigma}^{(g+1)}\\right\\|}{E\\|\\mathcal{N}(\\mathbf{0}, \\mathbf{I})\\|}-1\\right)\\right),\n\\end{equation}\nwhere ${E\\|\\mathcal{N}(\\mathbf{0}, \\mathbf{I})\\|} \\approx \\sqrt{n}+\\mathcal{O}(1 \/ n)$ is the expectation of the $l_2$ norm of samples drawn from $\\mathcal{N}(\\mathbf{0}, \\mathbf{I})$.\n\n\n\\subsection{Next Generation}\nOnce $\\boldsymbol{m}^{(g+1)}$, $\\boldsymbol{C}^{(g+1)}$ and $\\sigma^{(g+1)}$ have been calculated, they are inserted into \\eqref{cmaes2}, so that the next generation of samples can be drawn and the process repeated, until the convergence criteria are full-filled.\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}}