Algebraic Path Finding

21 March 2023

There’s something magical about polymorphic algorithms. Really, isn’t it awesome to write a piece of code once and then just tweak the meaning of some inputs to solve another seemly unrelated problem? Let’s see an example of those with applications ranging from regular expressions to matrix inversion, with a lot of useful queries on graphs along the way.

Today we will continue our explorations in the land of dynamic programming but with a much more abstract view. As it stands, dynamic programming in a finite horizon isn’t necessarily concerned with solving optimization problems. It can be abstracted to any kind of problem where we have well-behaved notions of combining and aggregating values along paths, and the solution methods remain exactly the same. This way, we go from the optimization view of the shortest path problem to the generic view of the algebraic path problem, which subsumes a lot of classical algorithms from computational mathematics.

By the way, be alert! Because it’s possible that your favorite algorithm is also an alternative view on finding shortest paths.

{-# LANGUAGE PackageImports,      DeriveTraversable   #-}
{-# LANGUAGE DataKinds,           KindSignatures      #-}
{-# LANGUAGE TypeApplications,    AllowAmbiguousTypes #-}
{-# LANGUAGE ScopedTypeVariables, RankNTypes          #-}
import         Control.Applicative (liftA2)
import         Data.Ord            (Down(..))
import         Data.Semigroup      (Min(..), Sum(..), Product(..))
import         Data.List           (foldl')
import         Data.Proxy          (Proxy(Proxy))
import         GHC.TypeNats        (KnownNat, Nat, natVal)
import "array" Data.Array          (Array, array, Ix(range), (!), (//), bounds)

Classical Shortest Paths

Consider a graph with weights on its edges. Like the one in the figure, for example. One way to represent this graph is via an adjacency matrix A whose component A(s, t) is the weight of the edge between nodes s and t. Notice that since there may not be edges between all vertices (after all, what’s the fun in a complete graph?), A is a partial function. We can solve this by adjoining a value \infty to the possible weights and assigning it as the weight of the non-existent edges. In case this solution seems like a dirty trick to you, a possibility is to think of this process as the same as returning a Maybe value to make a partial function total, with \infty playing the role of Nothing.1

Given this graph, a common question to ask is what are best ways to travel between its vertices. That is, if you want to go from node s to node t, which sequence of edges should you choose? If you’ve read what I have been writing about dynamic programming, you are probably already smelling the sweet scent of value functions and Bellman equations in the air.

Let \mathcal{S} be the (finite) set of nodes in the graph, and let’s define V \colon \mathcal{S}\times \mathcal{S}\to (-\infty, \infty] as the value function which tells us the length of the shortest path between two vertices. With a bit of dynamic programming, we can find out that V satisfies a Bellman equation:

\begin{aligned} V(s, s) &= 0 \\ V(s, t) &= \min_{q \in \mathcal{S}}\, A(s, q) + V(q, t), \; \forall s \ne t. \end{aligned}

The first line is the base case for the recursion and represents the 0-step paths. It costs nothing to just stay where you are, but it is in turn impossible to reach any other vertex in zero steps. This may seem really simple, but let’s nevertheless define a matrix I to represent this 0-step reachability relation. Why we want to do that will soon be clear.

I(s, t) = \begin{cases} 0 ,& s = t \\ \infty ,& s \ne t. \end{cases}

We can use I to write the recurrence more compactly:

V(s, t) = \min \left\{ I(s, t),\, \min_{q \in \mathcal{S}} A(s, q) + V(q, t) \right\}

What the above equation means is that the minimal distance between two vertices is the shortest among staying in the vertex (an empty path) and recursively following the shortest path with at least one edge.

Every problem is linear if you squint hard enough

The equation above is indeed rather ugly. However, by looking at it with care, one can the structure unveiling. First, the term

\min_{q \in \mathcal{S}}\, A(s, q) + V(q, t)

looks a lot like some kind of composition where we aggregate over the middle index q. Even more: if you squint your eyes enough, it looks a lot like the formula for matrix multiplication.

Well, let’s try to discover where this idea takes us. Even if it is just for the sake of abstract nonsense, it could be interesting. In general, when talking about real numbers, we use the usual arithmetic structure with addition and multiplication. Today we’re going to be eccentric and define a new kind of addition and multiplication:

\begin{aligned} a \oplus b &= \min\{a,\, b\}, \\ a \otimes b &= a + b, \\ \mathbf{0} &= \infty, \\ \mathbf{1} &= 0. \end{aligned}

So, what is happening here? The thing is: the real numbers extended with \infty satisfy all the axioms for a Semiring structure if we take sum to be \min and product to be +. This is called the min-plus or Tropical semiring and, surprisingly, satisfies a lot of the axioms we expect from a sum and product The main difference is that instead of having inverses, \oplus is idempotent.

Before further exploring the properties of the Tropical semiring, or of semirings in general, let’s take advantage of our new notation to rewrite the ugly equation

V(s, t) = I(s, t) \oplus \bigoplus_{q \in \mathcal{S}} A(s, q) \otimes V(q, t)

into the much more elegant (and index-free!) matrix format

V = I \oplus (A \otimes V).

By now, the choice of the letter I for the 0-step becomes clear: it is the identity matrix in the Tropical algebra!

What is cool is that in this form, finding the shortest path becomes the same as solving a linear system. Can we transfer the methods from classical linear algebra to this semiring setting? Unfortunately the answer is in general negative… This is why we will have to require a further piece of structure from our semirings: a closure operator denoted ^* which returns the fixed point of the mapping x \mapsto \mathbf{1} \oplus a \otimes x, that is, we require that

x^* = \mathbf{1} \oplus a \otimes x^* = \mathbf{1} \oplus x^\star \otimes a.

Now, you may call me a cheater since this is the scalar version of the shortest path equation. Nevertheless, we will see that while such an operator tends to be pretty simple to define for the scalar field, it is also the missing piece for solving the Algebraic Path Problem. That is, an algorithm capable of solving the Bellman equation in a way that is polymorphic on the semiring.

Alright, that was a pretty mathematical introduction. It’s far from time for this post to become more computational. Let’s take the above discussion to Haskell land by defining a Semiring class.

class Semiring r where
 zero,  one   :: r
 (|+|), (|*|) :: r -> r -> r
 closure      :: r -> r

infixl 6 |+|
infixl 7 |*|
infixr 8 |^|

For later use, let’s also define an exponentiation operator using the classic divide-and-conquer formula.

(|^|) :: (Semiring r, Integral n) => r -> n -> r
x |^| 0 = one
x |^| n | even n    = x |^| div n 2 |*| x |^| div n 2
        | otherwise = x |*| x |^| (n-1)

As usual, there are also a bunch of laws that an element of this class must obey. Since those are all pretty standard, I will just direct you to the Wikipedia article on the topic.

The Tropical Semiring and Shortest Paths

One can think of a semiring as a structure with an operation for combining two values (\otimes) and another for aggregating values (\oplus). Since the min-plus semiring was our choice for intuition, let’s start the implementation with it.

data Tropical a = Finite a | Infinity
  deriving (Eq, Functor)

One thing to notice about our motivation is that we only used two operations on the real numbers: the capacity of combining weights along a path with addition (our product) and the capacity of aggregating weights of different paths by choosing the smallest among them (our addition). Thus, the Tropical structure works for any monoid endowed with a total order. We will define the operations in this more general setting2 however, I advise you to keep the Real numbers in mind for the intuition. You will shortly see why this generality pays off.

instance (Ord a, Monoid a) => Semiring (Tropical a) where
 zero = Infinity       -- Weight of nonexistent path
 one  = Finite mempty  -- Weight of empty path

For addition, Infinity works as identity since we are minimizing.

 Infinity |+| x        = x
 x        |+| Infinity = x
 Finite x |+| Finite y = Finite (min x y) -- Choose the smallest among paths

For the product, we combine two values in the usual way with the extension that Infinity is absorbing: adding an infinite quantity to any number, no matter how small, always produces an infinite result.

 x        |*| Infinity = Infinity
 Infinity |*| y        = Infinity
 Finite x |*| Finite y = Finite (x <> y)  -- Combine weights along a path

Now, this next one is interesting. The closure of x is equivalent to the shortest distance in a graph with a single node. If there are no negative cycles, this should equal the weight of an empty path, which is the identity to the product.

 closure _ = one -- the distance taken by not moving

Of course, we are not interested only in single node graphs. Let’s take a look at weighted adjacency matrices.

Matrices over a semiring are a semiring

Since working wht Haskell arrays may be described as anything but pleasent, let’s define some simple wrappers to ease our life a little bit. We will go with some dependentish square matrices, but nothing that makes our code too complicated.

type Edge = (Nat, Nat)

-- | n x n square matrix with components in r
newtype Matrix (n :: Nat) a = Matrix (Array Edge a)
  deriving (Eq, Functor, Foldable, Traversable)

To ease our life, let’s also define some methods to get a cleaner interface. First of all, a smart constructor that transforms a function on pairs of indices into a Matrix.

-- | Extract a type-level natural to the term level
nat :: forall n. KnownNat n => Nat
nat = natVal (Proxy @n)

matrix :: forall n a. KnownNat n => (Edge -> a) -> Matrix n a
matrix f = Matrix $ array ends [(x, f x) | x <- range ends]
  where ends = ((1, 1), (nat @n, nat @n))

And another smart constructor that turns a list of edges into the respective adjacency matrix.

adjacency :: forall n r. (KnownNat n, Semiring r)
          => [(Edge, r)] -> Matrix n r
adjacency es = let Matrix o = one :: Matrix n r
               in Matrix (o // es)

It is also nice to have an Applicative instance, in order to aide us during future lifts. Notice that since our matrices are a type with fixed size, we can lift operations by simply matching the components.

instance KnownNat n => Applicative (Matrix n) where
 pure x = matrix (const x)
 (Matrix f) <*> (Matrix y) = matrix $ \(s,t) -> (f ! (s,t)) (y ! (s,t))

With these tools, we can keep away from the Array low-level API, with the exception of indexing. Now we can properly define the Semiring instance for a matrix.

Addition and zero are simply defined pointwisely.

instance (KnownNat n, Semiring r) => Semiring (Matrix n r) where
 -- These are the pointwise lifts of their scalar versions
 zero  = pure zero
 (|+|) = liftA2 (|+|)

For multiplication, we define one as the identity matrix with ones on the diagonal and zero elsewhere (the same as in the introduction), and use the traditional definition to build the product.

 -- Identity matrix
 one  = matrix $ \(i, j) -> if i == j then one else zero
 -- Matrix multiplication using Semiring operators
 Matrix x |*| Matrix y = matrix contract
  where
   contract (s, t) = add [x ! (s, q) |*| y ! (q, t) | q <- [1..nat @n]]
   add             = foldl' (|+|) zero

Finally, it is time for our main algorithm: how to calculate the closure of a matrix. Let’s take another look at the equation that A^\star must obey in order to gather some intuition.

A^\star = I \oplus (A \otimes A^\star)

Let’s try an experiment: what happens if we repeatedly substitute the A^\star term on the right-hand side with the entire right-hand side? After all, they’re equal… By doing that, we get a representation of the closure as a power series.

\begin{aligned} A^\star &= I \oplus A \oplus A^2 \oplus A^3 \oplus A^4 \oplus \ldots \\ &= \bigoplus_{k = 1}^\infty A^k \end{aligned}

What does this view contributes to us? Let’s take a look in the Tropical Semiring. Each power A^k of the adjacency matrix has at its components A^k(s, t) the distance of the shortest path from s to t with exacly k edges. Thus, while the recursive Bellman equation decomposes the cost of an edge plus the cost of a smaller path, this power series formulation, on the other hand, is saying that the optimal cost among all paths may be decomposed by path length. The minimum among all path is the minimum among all paths of any fixed length.3

Since there are just n vertices and we are assuming no negative cycles in order for the problem to be well-posed, looking at any power A^{n+1} is redundant. Because, without negative cycles, it is never worth it to pass through any vertex more than once.

A way to calculate the closure is with a variation the Floyd-Warshall or the Gauss-Jordan algorithm that is adapted to work on any closed semiring.4

 closure x = one |+| foldl' step x [1..nat @n]
  where step (Matrix m) k = matrix relax
         where relax (i, j) = let c = closure (m ! (k, k))
                 in m ! (i, j) |+| m ! (i, k) |*| c |*| m ! (k, j)

To see that it is in general necessary to cross n vertices, consider a graph consisting of a single path passing through all nodes. It takes exactly n iterations to discover that all vertices are reacheable from the first.

Everything you ever wanted to know about your graphs

Great, we have all the necessary tools to solve the Bellman Equation on the tropical semiring! Let’s write a little wrapper that turns an association list of weighted edges into the matrix of shortest distance between vertices. It works by converting the weights into a Tropical semiring where we use Haskell’s Sum monoid to indicate that it should combine values with addition.

shortestDistances :: (KnownNat n, Ord r, Num r)
                  => [(Edge, r)] -> Matrix n (Tropical (Sum r))
shortestDistances = closure . adjacency . fmap tropicalize
 where tropicalize = fmap (Finite . Sum)

This matrix represents the problem’s optimal value function, and, from it we can reconstruct the optimal paths. But wouldn’t it be cooler if the algorithm could already form the paths for us? It is, in some sense, walking over the graph after all… As you may guess, there is a semiring capable of just that!

Thanks to our generic Tropical semiring and Haskell’s default instances, this addition is almost an one-liner.

-- | Perform an optimization problem storing the optimal paths
solveWithPath :: (KnownNat n, Ord (f r), Monoid (f r))
              => (a -> f r) -> [(Edge, a)] -> Matrix n (Tropical (f r, [Edge]))
solveWithPath f = closure . adjacency . fmap withEdge
 where withEdge (e, w) = (e, Finite (f w, [e]))

shortestPath :: (KnownNat n, Ord r, Num r)
             => [(Edge, r)] -> Matrix n (Tropical (Sum r, [Edge]))
shortestPath = solveWithPath Sum

Well, how does the above work? The List type has a default Monoid instance given by [] and concatenation, and pairs of monoids are also monoids with pointwise operations. Furthermore, Haskell provides default Òrd instances for pairs and lists using the lexicographic order. Thus, in the type Tropical (Sum r, [Edge]), the product sums the weights in the first component and concatenate the paths in the second, while the addition takes the lexicographic minimum: keep the pair with smallest weight or, if they are equal, disambiguate the paths lexicographically. Since we only want one shortest path, any disambiguation criterion works, and we happily receive our path.

Of course, this technique works for many more operations. By using the Down constructor, which inverts an Ord instance, we can also solve maximization problems! For example, by changing a single constructor, we can query for the largest path.

largestPath :: (KnownNat n, Ord r, Num r)
            => [(Edge, r)] -> Matrix n (Tropical (Down (Sum r), [Edge]))
largestPath = solveWithPath (Down . Sum)

Keep in mind that in this case, our previous assumption of no negative cycles becomes the dual, but for some reason much less agreeable, assumption of positive cycles.

Now suppose that you heavy a very important letter to send to somebody you know. Unfortunately, the roads (or the network links) are each day more dangerous and finding a route that will actually deliver your message is no easy task. One can model this situation as a graph whose weights this edge’s reliability, that is, the probability of the message crossing that edge without being compromised. Assuming independence, the reliability of a path is the product of its edge’s probabilities and we can find the most reliable path by maximizing among all of them. At this point, you can already guess: there is a semiring for that. Use take \max to be the sum and \times for the product et voilà, were’ done.

mostReliablePath :: (KnownNat n, Ord r, Num r)
                 => [(Edge, r)] -> Matrix n (Tropical (Down (Product r), [Edge]))
mostReliablePath = solveWithPath (Down . Product)

Suppose now that you’re not worried about the message failing to be delivered, but with its size. You want to send a really big package and most know whether it can pass on the road. In order to guarantee the delivery, you decide to find the widest path on the graph. To find it, we define the width of a path as the minimum width among its edges (the bottleneck) and then aggregate among all paths by choosing the one with the largest bottleneck. That’s a semiring with \oplus = \max, and \otimes = \min.

widestPath :: (KnownNat n, Ord r, Bounded r)
           => [(Edge, r)] -> Matrix n (Tropical (Down (Min r), [Edge]))
widestPath = solveWithPath (Down . Min)

Besides those, a lot of other queries on graphs are realizable in terms of semirings, and specially some variation of the tropical semiring. A book with a lot of examples in the applied field of networks is the one by John S. Baras and George Theodorakopoulos5.

Transitive Closures of Relations

Let’s now divert our attention to another topic that is nevertheless closely related: finite relations. Think about the common binary relations that we work with everyday: equality, order (\le, <, \ge, >), equivalence relations… All of them share two important properties: transitivity and reflexivity.

So, if you’re going to work with some relation, a typical process is to complete it in order to turn this relation transitive and reflexive. Of course, your relation is just for the VIP and you don’t want to add every arbitrary pair of elements during this operation. You want to find the smallest reflexive-transitive relation containing yours.

This is called the reflexive-transitive closure of a relation. Now, how do we calculate this thing? As you may imagine, there is a semiring made just for that and the algorithm is exactly the same as for shortest paths. All we have to do is to use the Boolean semiring.

instance Semiring Bool where
 one  = True
 zero = False
 (|+|) = (||)
 (|*|) = (&&)
 closure _ = one  -- Reflexivity for a single vertex

Using the old trick of representing a finite relation as a Boolean matrix, the operation we need is exaclty the closure for the matrix semiring.

reflexiveTransitive :: KnownNat n => Matrix n Bool -> Matrix n Bool
reflexiveTransitive = closure

Interestingly, the above can also be comprehended as a common operation on graphs. By translating the boolean matrix into the (unweighted) adjacency matrix of a directed graph, this closure returns the graph whose edges are all paths on the original graph. If one can go from s to t in G, then G^\star(s, t) = \mathtt{true}. In the figure below, we illustrate this relationship.

Free as in Regular Expressions

When we study some kind of algebraic structure, we tend to find out some kind of datastructure that is intimately related to it. For example, monoids have lists, magmas have binary trees, and vector spaces have fixed-length boxes of numbers. These datatypes are a way to represent the free version of the algebraic structure. The actual definition of something being free in mathematics is a bit too technical for this blog’s tone, but the important to us is that, in general, a free structure can act as a skeleton (or syntax tree) representing a computation and any other instance of the algebraic structure can be recovered with a suitable interpreter.6

Well, do you have any guess for what is the free closed semiring? Perhaps surprisingly, it consists of Regular Expressions, at least when we assume the sum to be idempotent.7 Yep, regular expressions, those strange grawlixes that perl programmers love so much and that renders your queries completely unreadable a couple of weeks after you’ve written them. Since I don’t want this page to be worthy of a comic book cursing context, we will follow a more disciplined approach and represent the regular expressions as its own a datatype.8

data Regex a = Nope                      -- Empty Set: matches nothing
             | Empty                     -- Empty String: matches anything
             | Literal a                 -- A single character from the alphabet
             | Union (Regex a) (Regex a) -- matches either an argument or the other
             | Join  (Regex a) (Regex a) -- Matches one argument followed by the other
             | Many  (Regex a)           -- Kleene star: Matches zero or more instances of something

Notice that we’re limiting ourselves to truly regular expressions, those that represent some kind of Regular Language.

The semiring instance is straighforward, since the constructors closely resemble the class methods. We will only implement a couple simplifications in order to not get expressions with redundant parts.

instance Eq a => Semiring (Regex a) where
 zero    = Nope
 one     = Empty
 -- Union with some eliminations for empty elements
 Nope |+| e    = e
 e    |+| Nope = e
 -- Remove empties when possible
 Empty    |+| (Many e) = Many e  -- Star contains the empty string
 (Many e) |+| Empty    = Many e
 Empty    |+| Empty    = Empty
 x        |+| y        = Union x y
-- Concatenation and some simplifications
 Nope  |*| x     = Nope          -- Annihilation
 x     |*| Nope  = Nope
 Empty |*| x     = x             -- Identity
 x     |*| Empty = x
 x     |*| y     = Join x y
 -- Kleene star
 closure Nope     = Empty        -- the closure is at least empty
 closure Empty    = Empty        -- many instances of something empty
 closure (Many e) = Many e       -- idempotence of Kleene star
 closure e        = Many e

The previous definition was long but rather mechanical and, most important of all, gave us a shining new semiring to play with! Alright, what does the closure of a matrix of regular expressions? First, a graph labeled with regular expressions is exactly a finite state machine (more precisely, a ε-NFA), and since in this case \oplus is union and \otimes is concatenation, we get from the power series interpretation that the closure A^* is the component-wise union of all fixed length paths one can follow in this automaton. Hence, each component A^*(s, t) is a regular expression representing the language accepted by this automaton with initial state s and accepting state t.

What is the Floyd-Warshall implementation, when specialized to this context? Here it becomes Kleene’s algorithm for converting a finite state machine into a regular expression.

As we discussed earlier, regular expressions are the free closed (idempotent) semiring. This means that any operation on semirings, such as finding shortest-paths, may be first calculated on regexes and then converted to the proper type. We achieve this with an interpreter function.

interpret :: Semiring r => (a -> r) -> (Regex a -> r)
interpret f Nope        = zero
interpret f Empty       = one
interpret f (Literal a) = f a
interpret f (Union x y) = interpret f x |+| interpret f y
interpret f (Join  x y) = interpret f x |*| interpret f y
interpret f (Many  e)   = closure (interpret f e)

This is similar to how Haskell’s foldMap works for monoids or the eval function that we defined for calculus expressions on another post. All of them are some kind of “realizations of typeclass”.

The usefulness of this interpreter comes from the fact that it commutes with the matrix operations. Thus, whenever we want to perform many queries in a same graph with idempotent aggregations (such as shortest paths, largest paths, most reliable paths, widest paths etc.), we may first calculate the regular expressions representing these paths and then collapse them separately for each semiring.

Classical Matrix Inversion

To wrap up this post, let’s take a look at a semiring that is not idempotent: the classical field structure on the real/complex numbers. Or, to be technicality accurate, one of these fields complete with an extra point to amount for the non-invertibility of zero.

data Classical a = Field a | Extra
  deriving Eq
instance (Eq a, Fractional a) => Semiring (Classical a) where
 zero = Field 0
 one  = Field 1
 Extra   |+| _       = Extra
 _       |+| Extra   = Extra
 Field x |+| Field y = Field (x + y)

 Field 0 |*| _       = Field 0
 _       |*| Field 0 = Field 0
 Extra   |*| _       = Extra
 _       |*| Extra   = Extra
 Field x |*| Field y = Field (x * y)

For a field, the closure operation has, in general, a closed form solution:

x^\star = 1 + x\cdot x^* \implies x^* = (1 - x)^{-1}.

Great, I must admit that I had already started missing being able to subtract and divide our numbers… Unfortunately, not everything is flowers and the number 1 is non-invertible. That’s why we had to add an extra point to the structure.

 closure Extra     = Extra
 closure (Field 1) = Extra
 closure (Field x) = Field $ recip (1 - x)

Since matrices have themselves a notion of inverse, their closure also satisfies the equation B^* = (I - B)^{-1} whenever the inverse on the right makes sense. With a change of variables I - B \to A, we arrive at a formula for matrix inversion:

A^{-1} = (I - A)^*.

Furthermore, this works whenever the right-hand side has no Extra terms, meaning that we’ve arrived at a safe inversion formula for a matrix.

inv :: (Fractional a, Eq a, KnownNat n) => Matrix n a -> Maybe (Matrix n a)
inv = traverse toMaybe . closure . conj
 where
  conj a = one |+| fmap (Field . negate) a   -- A -> I - A
  toMaybe Extra     = Nothing
  toMaybe (Field a) = Just a

To which classical algorithm is this equivalent? By looking attentively at the definition of closure, you will see that this is exactly the Gauss-Jordan elimination, where in our case multiplying by the closure takes the same role as dividing by a row in the classical presentation.

Farewell

This was a lot of fun, but every post must come to an end. Most of it, I’ve learned from Baras and Theodorakopoulos9, which also contains a lot of examples in the context of networks. If you want to go deeper on tropical mathematics and the theory of closed semirings, the best book I know is Michel Gondran and Michel Minoux10. A lot of the implementations on this site where inspired by those from Stephen Dolan11’s paper and the very interesting blog post by Russel O’Connor12.

Farewell, and have fun spotting shortest paths where you would never imagine before. Heads up, there’s a lot of semirings lurking around.


  1. Assuming you don’t consider Kleisli categories are as dirty as arithmetic with infinities.↩︎

  2. As if this posts wasn’t abstract enough.↩︎

  3. This is very similar to how a linked list can be viewed either as a pair containing a value and a reference to another list or as a vector that can have any possible length.↩︎

  4. There is also a pretty elegant alternative that, unfortunately, only works for idempotent semirings. It is a little slower (O(n^3 \log n) versus O(n^3)) but it is worth showing here. If R is idempotent, then all cross-terms in a binomial cancel out and we get (I \oplus A)^n = \bigoplus_{k=0}^n A^k = A^\star. From this equation follows a pretty sleek one-liner:

    closure a = (one |+| a) |^| nat @n

    Sadly, it requires idempotence…↩︎

  5. Path Problems in Networks, Synthesis Lectures on Learning, Networks, and Algorithms (Cham: Springer International Publishing, 2010), https://doi.org/10.1007/978-3-031-79983-9.↩︎

  6. As always in mathematics there are some technicalities about when this holds. That’s a cool topic related to monads and CT, but too out of scope for this discussion. I suggest the books: Paolo Perrone, “Notes on Category Theory with Examples from Basic Mathematics,” February 2021, chap. 5, https://doi.org/10.48550/arXiv.1912.10642 and Emily Riehl, Category Theory in Context, n.d., chap. 5, https://math.jhu.edu/~eriehl/context/.↩︎

  7. To be precise, regular expressions are the free Kleene Algebra.↩︎

  8. For a primer on regular expressions in Haskell, see Pedro Vasconcelos’ A regular expression matcher.↩︎

  9. Path Problems in Networks.↩︎

  10. Graphs, Dioids and Semirings, vol. 41, Operations Research/Computer Science Interfaces (Boston, MA: Springer US, 2008), https://doi.org/10.1007/978-0-387-75450-5.↩︎

  11. “Fun with Semirings: A Functional Pearl on the Abuse of Linear Algebra,” in Proceedings of the 18th ACM SIGPLAN International Conference on Functional Programming (Boston Massachusetts USA: ACM, 2013), 101–10, https://doi.org/10.1145/2500365.2500613.↩︎

  12. “A Very General Method of Computing Shortest Paths,” n.d., https://r6.ca/blog/20110808T035622Z.html.↩︎

Spread the Word