1 INTRODUCTION
Quantum mechanics, in its traditional formulation, is based on Hilbert space [28]. More precisely, it is based on mappings of Hilbert spaces. In the simplest telling, states are unit vectors, evolving along certain linear maps between Hilbert spaces. Taking into account probabilities, states become operators, and dynamics certain linear maps between the resulting algebras. Either way, it is fair to say that the category of Hilbert spaces and linear maps is crucial to quantum theory.
This model gives accurate results, but the framework of Hilbert spaces is hard to interpret physically from first principles [24]. Many reconstruction programmes try to reformulate the framework mathematically, so that the primitive assumptions rely on fewer physically unjustified details [9]. For example, instead of starting with Hilbert space, they start from operator algebras [20], orthomodular lattices [23], generalised probabilistic theories [25] or from categories [13].
In this final case, the situation comes down to the following. Somebody hands you a category. How do you know it is (equivalent to) that of Hilbert spaces? The answer depends on which morphisms between Hilbert spaces you choose exactly. Previous work [12] settles the case of bounded linear maps, by giving axioms characterising that category precisely. Importantly, those axioms are purely categorical, and do not mention probabilities, complex numbers, or other analytical details that you may think are fundamental to Hilbert space.
Nevertheless, the story does not end with that answer, because not all bounded linear maps are physical. Quantum systems may only evolve along unitary linear maps. Taking quantum measurement into account allows the larger class of linear contractions [1, 26]. But no quantum-theoretical process can be described by a bounded linear map that is not a contraction. This article settles the case for the physical choice of morphisms being linear contractions.
The main idea is the following: The subcategory of Hilbert spaces and linear contractions generates the category of Hilbert spaces and all bounded linear maps as a monoidal category with invertible non-zero scalars (where scalars in a monoidal category are endomorphism on the tensor unit). Specifically, if you formally adjoin inverses for all non-zero scalars to the former category, you get the latter. We find properties of the former category that guarantee that its completion satisfies the axioms of [12] and hence is the category of Hilbert spaces and bounded linear maps. Finally, using the nature of the completion, we show that the former category must consist of exactly the linear contractions.
The presented axioms are natural from a categorical point of view. They progress in a modular way, so that fragments are compatible with other research programmes. For example, biproducts and equalisers are used as in abelian categories and regular categories [3], the tensor product is used as in categorical quantum mechanics [13] and the string diagrams of ZX-calculus [6] and the combination of daggers and kernels is used as in quantum logic [10]. Overall, our axioms begin with the structure of dagger rig categories, which form the starting point for a complete graphical model of mixed-state quantum theory [5] as well as for both classical and quantum reversible programming [4, 11]. Moreover, these axioms enable categorical methods such as the use of universal properties in quantum information theory [14].
In addition to being an important characterisation in its own right, this theorem is also a stepping stone towards trying to characterise related categories: Hilbert spaces and completely positive morphisms, going towards foundations of quantum information theory; Hilbert spaces and unitaries, going towards foundations of quantum computing; and Hilbert modules and adjointable morphisms, going towards unitary representations and foundations of quantum field theory [2].
The rest of this article is structured as follows. We start by stating the axioms in Section 2. Their meaning is discussed in Section 3. In particular, the category of Hilbert spaces and linear contractions is defined there, and it is shown how it satisfies the axioms. Next, Section 4 derives basic properties from the axioms. The real work begins in Section 5, which defines a completion by which we can abstractly recognise linear contractions among all bounded linear maps. Section 6 puts everything together to prove the main theorem.
2 THE AXIOMS
We will consider the following properties of a locally small category that is equipped with a contravariant endofunctor
and two symmetric monoidal structures
and
.
- (1) The contravariant endofunctor satisfies for all objects and for all morphisms . A category equipped with such a dagger is also called a dagger category. If , we call a dagger monomorphism, and if additionally , we call it a dagger isomorphism.
- (2) The category is a dagger rig category: the unitors, braidings and associators of the monoidal structures and are dagger isomorphisms, and they satisfy and for all morphisms and , and there are natural dagger isomorphisms
which satisfy the coherence conditions (I)–(XXIV) of [18].
- (3) The unit 0 is initial, and thus a zero object. We write for the unique morphism that factors through the object 0. It follows that there are natural morphisms:
This property of makes the category so-called affine or semicartesian.
- (4) The injections and are jointly epic: as soon as and .
- (5) Mixture occurs: there exists a morphism with .
- (6) The unit is dagger simple: any dagger monomorphism is either invertible or zero but not both.
- (7) The unit is a monoidal separator: as soon as for all and .
- (8) Any two parallel morphisms have a dagger equaliser, that is, an equaliser that is a dagger monomorphism.
- (9) Any dagger monomorphism is a dagger kernel, that is, a dagger equaliser of some morphism and .
- (10) Subobjects are determined by positive maps: monomorphisms and satisfy for some isomorphism if and only if .
- (11) Any directed diagram has a colimit.
Observe that these axioms are all elementary properties of categories: none of them refer to notions such as complex amplitudes, probabilities, norm, convexity, continuity or metric completeness. These are natural axioms from a categorical point of view, and thus, we hope that they can be interpreted as natural assumptions about quantum information processes. For example, one might interpret axiom (1) in terms of time reversal symmetry and axioms (2)–(5) in terms of superposition. However, this task is beyond the scope of this work, and we leave it to future research. Furthermore, we conjecture that these axioms are all necessary for our result and independent. For example, the category of sets and partial injections satisfies all the axioms except (5).
We end this section with a brief summary of the main ideas of previous work [12] that we will rely on. That work assumed similar axioms as here. Axioms (1) and (7)–(9) are literally the same. Axioms (2)–(5) resemble the axiom there stipulating dagger biproducts. Axiom (6) here concerns dagger monomorphisms but there concerns all monomorphisms, and similarly, axiom (11) here concerns colimits in the category itself but there concerns colimits in the wide subcategory of dagger monomorphisms. Axiom (10) here has no analogue there. A category satisfying those axioms is there shown to be equivalent to the category of Hilbert spaces as follows. Just as vectors of a Hilbert space are recovered as linear maps in the complex case, the idea is that an object can be turned into a concrete Hilbert space by taking the set , which intuitively consists of the vectors of , and equipping it with the structure of a complete inner product space over the field . The endomorphisms always form a commutative monoid, but due to the axioms imposed on , it turns into an involutive field [12, Lemma 1]. In fact, by Solèr's theorem, this field must be either or [12, Proposition 5]. The main result then showed that there is a functor from to the category of Hilbert spaces that defines an equivalence of dagger rig categories [12, Theorem 8].
3 THE CATEGORY
We define our main model. Recall that a Hilbert space is a vector space over or , together with an inner product whose induced norm makes Cauchy-complete.
Definition 1.A linear function between Hilbert spaces is bounded when there exists a constant such that for all ; write for the infimum of such . The function is called a contraction when , that is, when for all . Such linear functions are also known as non-expansive or short maps. We write and for the categories of Hilbert spaces and bounded linear maps, and and for the categories of Hilbert spaces and short linear maps, respectively, over the real and complex numbers. When the distinction does not matter, we will simply write and .
Let us discuss how
satisfies the axioms.
- (1) The dagger is provided by adjoints: for any bounded linear function , there is a unique linear map satisfying for all and . This is a well-defined functor on because . The dagger monomorphisms are the isometries, and the dagger isomorphisms are the unitaries. We shall use both terminologies, depending on whether we are working with an abstract dagger category or are in the concrete setting of Hilbert spaces.
- (2) The monoidal structures and are produced by tensor products and direct sums of Hilbert spaces, respectively. This is well defined in because and . The unit is the one-dimensional Hilbert space given by the base field, and the unit 0 is the zero-dimensional Hilbert space .
- (3) Linear maps must preserve the 0 vector, so the object 0 is indeed initial.
- (4) The injections and have images that linearly span , so are clearly jointly epic.
- (5) The vector in the two-dimensional Hilbert space satisfies (5), so this axiom may be read as saying that non-trivial superpositions of qubits exist.
- (6) The only vector subspaces of the base field are the zero-dimensional one and the whole space itself.
- (7) Linear contractions are continuous, and the linear span of pure tensor elements is dense in , so morphisms in out of a tensor product are indeed determined by their action on pure tensor elements.
- (8) If are linear contractions, then is a closed subspace of and hence an object of , and the inclusion is automatically an isometry.
- (9) Any closed subspace is a kernel of the orthogonal projection onto its orthocomplement, which is a contraction.
The last two axioms are perhaps a bit less standard, and we spell out proofs.
Lemma 2.The category satisfies (10).
Proof.Observe that any isomorphism in is unitary: if and , then for all vectors , so is an isometry and hence unitary. This makes one direction of (10) clear: if for unitary , then . The other direction follows directly from Douglas' lemma [7].
Lemma 3.The category satisfies (11).
Proof.See also [22, Proposition 3.1]. Let be a directed partially ordered set, and be a diagram . Let be the vector space , where and are identified when for some . Writing for the equivalence class of , define ; it is routine to verify that this limit exists and defines a seminorm on . (Recall that a seminorm is a function that satisfies all properties of a norm except positive definiteness; so, need not imply .) This seminorm satisfies the parallelogram law. So, if , then is an inner product space. Call its completion , define by and observe that this is a cocone.

Suppose that
is another cocone. It defines a function
that respects the equivalence, and so lifts to a linear function
. Furthermore, if
satisfies
, then
. Thus,
lifts to a linear function
. Similarly, this function is a contraction, and so extends to a linear contraction
. By construction,
for all
, and so,
is a mediating map from the cocone
to the cocone
. Finally, this mediating map is unique, because the
have jointly dense range in
by construction.
Axioms (10) and (11) are the only ones that hold in but not in . The behaviour following from these two axioms accounts for the difference between the two categories.
We end this section by determining the subobjects of in . Contrast the following lemma to the situation , where the subobjects of are ; the crucial difference is that scalars are invertible in but not in .
Lemma 4.There is an order isomorphism between the subobjects of in and the real unit interval [0,1].
Proof.Consider a monomorphism of into the base field . It is an injective linear contraction . The kernel of is zero, and so or . Hence, a subobject of is represented by the unique morphism or by an injective contraction . Up to isomorphism of subobjects, the latter morphisms are scalars in the interval (0,1]. Hence, the subobjects of in are canonically in bijection with the closed unit interval [0,1]. This bijection is clearly an order isomorphism.
4 THE BASIC LEMMAS
From now on, we assume a (locally small) category that satisfies the axioms (1)–(11), and set out to prove that . In this section, we derive some basic properties of . We start by recalling a factorisation that already follows from (8) and (9) alone [10]. We summarise a proof here for convenience.
Lemma 5.Any morphism factors as an epimorphism followed by a dagger monomorphism .
Proof.By (1), is a cokernel of , so that , and always factors through via some morphism . It remains to show that is an epimorphism.
We first prove that if , then . Observe that . Hence, factors through via some .
But
, so
, and so
.
Next, we prove that is an epimorphism. Suppose that for morphisms . Let be a dagger equaliser of and . Then factors through , and, writing for a cokernel of , we infer .
But by the property we proved earlier, that means
, and so
is invertible. We conclude that
.
Next we notice a consequence of axiom (6). A scalar in a monoidal category is a morphism , and we can multiply any morphism by it to obtain a morphism that we denote [13, 2.1]. For a scalar and morphisms and , it follows from the bifunctoriality of the tensor product that . In particular, for scalars . Instead of , we may write 1 for the identity scalar.
Lemma 6.Every non-zero scalar is monic and epic.
Proof.Let be non-zero. Lemma 5 factors it as an epimorphism followed by a dagger monomorphism. Axiom (6) guarantees that dagger monomorphism is either 0 or invertible. By assumption, it cannot be zero, and so, is epic. Similarly, is epic, and so, is also monic.
It follows that every dagger monic scalar is invertible, for it cannot be zero, so must be epic as well as split monic.
Lemma 7.For all non-zero scalars and morphisms , if , then .
Proof.By two applications of axiom (7), it suffices to prove that for all and . But , and the scalar is monic by Lemma 6.
The following two lemmas are consequences of axiom (10).
Lemma 8.Any isomorphism is a dagger isomorphism. Any split monomorphism is a dagger monomorphism.
Proof.Let be an isomorphism. Taking and in (10) shows that . Hence, .
Now suppose that is a split monomorphism. Factor for a dagger monomorphism and an epimorphism . Then the epimorphism is itself split monic. But in any category, a split monic epimorphism is an isomorphism, and hence, is a dagger isomorphism. Therefore, is a dagger monomorphism.
Lemma 9.If a scalar and a morphism satisfy , then for a dagger monomorphism .
Proof.Factor for a dagger monomorphism and an epimorphism . Then . Now (10) and Lemma 8 provide a dagger isomorphism such that , and so for the dagger monomorphism .
5 THE COMPLETION
This section contains the main construction of the proof. Lemmas
4,
6 and
8 show that scalars in
correspond to the unit disc:
Observe that non-zero scalars that are not on the unit circle have no inverse in
. But they do in
. In fact,
is the localisation of
at non-zero scalars. We now abstractly construct the localisation of a monoidal category at its non-zero scalars in a way that respects the axioms (1)–(11). Write
for the scalars of
.
Proposition 10.There is a category with the same objects as , where a morphism consists of a non-zero scalar and a morphism in modulo the following equivalence relation:
The identity on
is
, and composition is given by:
There is an embedding
defined by
.
Proof.The relation is clearly reflexive and symmetric. To see that it is transitive, suppose that and . Then and . It follows that . Lemma 7 now shows that , that is, . Hence, is an equivalence relation.
To see that the composition is well defined, suppose that and . Then , so . It is clearly associative and satisfies the identity laws. If and are non-zero scalars, then is non-zero because is monic by Lemma 6.
The assignment is functorial, injective on objects and faithful.
The previous proposition concretely describes a quotient category [19, II.8]: writing for the one-object category of non-zero scalars of , take the quotient of under the congruence relation .
Example 11.There is an isomorphism of categories. It is defined as the identity on objects, and as on morphims. It is faithful by construction, because implies and so . To see that it is full, let be a bounded linear function. If , then . If , then is a scalar in and is a contraction, so .
Lemma 12.The category inherits monoidal structure from , and the functor is strict monoidal for .
Proof.Observe that the equivalence relation of Proposition 10 is a monoidal congruence: if and , then .
Before we show that it satisfies the axioms of [12] one by one, let us first establish the universal property that justifies the notation . Notice that any morphism in can be factored as .
Proposition 13.Any functor that is strong monoidal for such that is invertible for all non-zero scalars factors through a unique functor that is strong monoidal for via the functor .
Proof.Define the functor by on objects, and by on morphisms. This is the only functor making the triangle commute, because it is completely determined by its values at and for all morphisms and all scalars in .
Thus is the localisation of at all non-zero scalars: it formally adjoins inverses for all non-zero scalars to . The concrete description of Proposition 10 simplifies the general construction for localisation [17].
Lemma 14.The category has a dagger, and the embedding preserves it. The embedding restricts to an isomorphism between the wide subcategories of dagger monomorphisms in and .
Proof.Set , which is clearly well defined. By Proposition 10, the embedding sends an object to itself, and sends a morphism to . It is faithful by Lemma 7. The embedding clearly preserves daggers, hence dagger monomorphisms, and so restricts to the wide subcategories of dagger monomorphisms. This restricted functor is still bijective on objects and faithful. To see that it is full, let be a dagger monomorphism in . The assumption means , so Lemma 9 implies for a dagger monic in . But then is the image of a dagger monic in under the embedding.
An object is simple when any monomorphism into it must be 0 or invertible. This requirement on is stronger than axiom (6), which only concerns dagger monomorphisms.
Lemma 15.The tensor unit in is simple.
Proof.First, we will show that a monomorphism represents the same subobject of I as if and only if . Suppose is a monomorphism in . It is always true that as subobjects, so is the minimum subobject if and only if . This happens when there is a morphism such that , that is, . Because , by Lemma 7, this means exactly that . Thus, is the minimum subobject if and only if .
Next, we will show that a monomorphism represents the same subobject of as if and only if . Similarly to the last paragraph, as subobjects, so is the maximum subobject if and only if . This happens when there is a morphism such that , that is, . We may choose and to see that this is true as soon as . This final inequality is equivalent to because is monic in this case.
Now, either or . If , then is the minimum subobject, and if , then is the maximum subobject. Therefore, is simple in .
Lemma 16.The tensor unit is a monoidal separator in .
Proof.Let , and suppose for all and and in . Then is equal to . By (7), . Taking gives , that is, .
A dagger biproduct of and in a dagger category with a zero object 0 is a product such that and are dagger monomorphisms and .
Lemma 17.The category has dagger biproducts, given by and 0, and the functor is strict monoidal for .
Proof.For all morphisms and , define a morphism by . This produces a functor . The embedding induces coherence isomorphisms, making a symmetric monoidal category. To show that it is cartesian monoidal, it suffices to prove that is the product of and with projections and , because 0 is clearly terminal [8]. It then follows that is a dagger biproduct of and because and are dagger monomorphisms.
Axiom (5) provides a map in such that the scalars and are non-zero. We will use this to show that forms a product in with projections and . Let and in . Define to be the following composite:
Here
is made up from
and isomorphisms provided by axiom (2). We will prove that the following diagram commutes in
:
The left triangle commutes because the following diagram commutes in
:
The right triangle commutes similarly.
Moreover, the mediating map is unique. For suppose both satisfy and . Then and . Because and are jointly epic by axiom (4), we have that . Thus, .
Lemma 18.The category has dagger equalisers, and the functor preserves them.
Proof.Let . Let be a dagger equaliser of and in . Then . Suppose that also for a non-zero scalar and . Then , and hence, there is a unique morphism in with . Thus, , that is, .


To see that
is the unique such morphism, suppose that we also have
, that is,
. We have that
. The morphism
is dagger monic because it is a dagger equaliser, so
. By Lemma
7,
, and so
.
Lemma 19.Any dagger monomorphism in is a kernel.
Proof.Consider a dagger monomorphism in . By Lemma 14, we may assume that it is of the form , where is a dagger monomorphism in . Now (9) gives for some in . We claim that in .
Firstly, clearly because . Next, suppose that for some . Then , so by Lemma 7. Thus, factors through via some . Then because .
Finally, to see that
is unique, suppose that
. Then
. It follows that
because
is dagger monic, and hence
.
Lemma 20.The wide subcategory of dagger monomorphisms in has directed colimits.
Proof.Consider a directed partially ordered set indexing a diagram of dagger monomorphisms in . By Lemma 14, we may assume that the diagram is represented as for dagger monomorphisms in for . Axiom (11) provides a colimiting cocone in . By Lemma 14, for to be a colimiting cocone in the wide subcategory of dagger monomorphisms of , it suffices to prove that each is dagger monic in .
Fix an index
, and restrict both the diagram and the cocone to indices
. This restricted cocone is still colimiting. Another cocone on the restricted diagram is given by
. Indeed, for
, we have:
Thus, we obtain a mediating morphism
satisfying
for all
. In particular,
. Therefore,
is a dagger monomorphism by Lemma
8.
Having established that all necessary axioms hold, we can now apply previous results and establish an equivalence with . We will describe this equivalence explicitly shortly.
Proposition 21.There is an equivalence of dagger rig categories .
Proof.Use Lemmas 12–20 to apply [12, Theorem 8].
6 THE THEOREM
The previous section showed that if a category satisfies (1)–(11), then the completion is equivalent to as a dagger rig category. Here, is an involutive field that is canonically isomorphic either to or to up to a choice of imaginary unit [12, Proposition 5]. To be explicit, the equivalence is implemented by the functor . The fact that this is an equivalence of dagger rig categories means the following. First, the functor is full and faithful, and any Hilbert space is unitarily isomorphic to for some object of . Also, the functor preserves the dagger: . Moreover, the functor is strong monoidal for , meaning that there are coherent isomorphisms and . Similarly, that the functor is strong monoidal for means that there are coherent isomorphisms and . Finally, modulo these coherence morphisms, the functor maps the distributors of of axiom (2) to the canonical distributors in .
To complete the proof of the main theorem, in this section, we will determine the image of in . The same proof applies to both the real and the complex cases. To simplify the presentation, we identify with its canonical image in .
Firstly, we characterise .
Lemma 22.We have that .
Proof.Let be an element of the unit interval. The morphism with components and is a dagger monomorphism and thus in . It follows that is also in . We conclude that . We also know that because contains all isometries. Therefore, .
For the reverse inclusion, recall that for any set , we have a directed diagram in that assigns an object to each finite , namely the biproduct of many copies of . The diagram consists of dagger monomorphisms for finite . For its colimit among the dagger monomorphisms of , write .
Because the morphisms of the diagram are dagger monomorphisms, it is also a diagram in . Write for the colimit of this diagram in . A priori, the objects and may not be isomorphic. We now specialise to the case .
Let . Without loss of generality, assume that the objects , for finite , are constructed using the canonical monoidal product on . Construct a natural transformation in such that is scalar multiplication by for each . In the colimit, we obtain a morphism such that the following square always commutes:
As in [
12],
for each
. Thus,
is a vector in the Hilbert space
. The wide subcategory of
with dagger monomorphisms is a subcategory of
, and thus,
factors through
. Therefore, the vector
is non-zero.
The operator satisfies . In other words, is an eigenvector of with eigenvalue . Thus, for all . But is bounded, so we must have . Therefore, this shows that . Altogether, we have equality.
Our next goal is to show that cannot contain morphisms of with . To engineer a counterexample, we first examine the situation in .
Example 23.Let be an increasing sequence in (0,1], and let be its supremum. The following is a directed diagram in :
This diagram admits two cocones, among others: firstly, a cocone into
whose edges are simply the scalars
; secondly, a cocone into
whose edges are the scalars
. This gives the following diagram:
It is routine to verify that
is a colimiting cocone.
The following lemma demonstrates the same phenomenon in .
Lemma 24.The functor restricts to a functor . Furthermore, for all objects and all sequences with , the following is a colimiting cocone in :
Proof.Let be a morphism in . Assume that , and let be such that and . Now , so is dagger monic and hence in . Furthermore, . But is a scalar in , contradicting Lemma 22. We conclude that , and that restricts to a functor .
Let be an increasing sequence of numbers with supremum 1. The morphisms are all in by Lemma 22. They form a directed diagram, which has a colimiting cocone . Now forms another cocone. Thus, there is a unique mediating morphism in :
The functor
maps the cocone
to a colimiting cocone in
. Thus, there is a mediating bounded operator
:
The universal property now implies that
is the identity on
. Because both
and
are contractions,
must be an isometry. It follows that there is a dagger monomorphism
in
, and hence in
, with
.
Thus, the morphisms and in satisfy . Because the functor is faithful, . In the same way, we find that , so , concluding that . Since is dagger monic, both and are dagger isomorphisms that are inverse to each other. Thus, the cocone is colimiting in .
Finally, we show that the image of in contains all contractions.
Lemma 25.For all objects and , .
Proof.It remains to show that morphisms of with are in . Firstly, suppose in satisfies . It follows that is a convex combination of unitary operators on the Hilbert space : there exist unitary operators and positive real numbers such that and [21, Theorem 13]. The operator
then satisfies
for all operators
on
. Furthermore,
is an isometry. Now [
12, Theorem 8] provides dagger isomorphisms
in
such that
for each
and a dagger isometry
in
such that
. By Lemma
14,
, and
are morphisms in
, and thus,
is a morphism in
. Also,
, so
. We conclude that
is in
and, more generally, that any endomorphism
of
with
is in
.
Next, suppose in satisfies . Using polar decomposition, any operator in factors as , where and are isometries and is a bounded operator on the domain of such that . Because is an equivalence of monoidal dagger categories, factors as , where and are dagger monomorphisms and is an endomorphism such that . Thus, is in . Therefore, any morphism in with is in .
Finally, suppose in satisfies . Let be a sequence of numbers with supremum 1. For each , we calculate that ; hence, . Therefore, the following diagram lies entirely in :
Lemma
24 provides a unique morphism
making this diagram commute. Thus,
. Since
, we conclude by Lemma
7 that
. Hence,
is in
. Therefore, any morphism
of
with
is in
.
We have arrived at the main theorem of this article.
Theorem 26.A dagger rig category satisfying axioms (1)–(11) is equivalent to or , and the equivalence preserves daggers and is strong symmetric monoidal for both and and preserves the distributors of axiom (2).
Proof.Combine Proposition 21 and Lemma 25.
All of the isomorphisms of axiom (2) are preserved, making any dagger rig category satisfying axioms (1)–(11) equivalent to or as a dagger rig category. The equivalence of Theorem 26 is a weak equivalence, that is, a full, faithful, and essentially surjective functor . In fact, it is essentially surjective in the stronger sense that every object of is dagger isomorphic to one in its image by Lemma 8.
Let be a category. If there exist a contravariant endofunctor and symmetric monoidal structures and that satisfy axioms (1)–(11), then there exists an equivalence , as an immediate consequence of Theorem 26. The converse also holds. For any equivalence , there exist a dagger and dagger symmetric monoidal structures and on such that the equivalence preserves all of these structures; we combine [27, Lemma V.1], [16, Corollary 3.1.6] and [15, Theorem 5] and use the axiom of global choice.
Thus, a category is equivalent to or if and only if there exist a dagger and symmetric monoidal structures and on satisfying axioms (1)–(11). In this way, these axioms serve to characterise and as categories with no additional structure.