Skip to main content

A generalized Robinson-Foulds distance for labeled trees

Abstract

Background

The Robinson-Foulds (RF) distance is a well-established measure between phylogenetic trees. Despite a lack of biological justification, it has the advantages of being a proper metric and being computable in linear time. For phylogenetic applications involving genes, however, a crucial aspect of the trees ignored by the RF metric is the type of the branching event (e.g. speciation, duplication, transfer, etc).

Results

We extend RF to trees with labeled internal nodes by including a node flip operation, alongside edge contractions and extensions. We explore properties of this extended RF distance in the case of a binary labeling. In particular, we show that contrary to the unlabeled case, an optimal edit path may require contracting “good” edges, i.e. edges shared between the two trees.

Conclusions

We provide a 2-approximation algorithm which is shown to perform well empirically. Looking ahead, computing distances between labeled trees opens up a variety of new algorithmic directions.Implementation and simulations available at https://github.com/DessimozLab/pylabeledrf.

Background

Phylogenic trees represent the evolutionary relationship between sets of genetic elements or taxa, where the elements of a set are in one-to-one relationship with the leaves of the corresponding tree [1]. Different phylogenetic inference methods may lead to different trees, and each method, typically exploring a large space of trees, can also result in multiple equally likely solutions for the same dataset. It follows that comparing trees is an essential task for finding out how inferred trees are far from one another, or how an inferred tree is far from a simulated tree or from a gold standard tree for the same datasets.

Designing appropriate tree metrics is a widely explored branch of research. A variety of measures have been designed for different types of trees, rooted or unrooted, some restricted to comparing tree shapes [2], others considering multilabeled trees, i.e. trees with repeated leaf labels [3] and yet others considering information on edge length [4]. In particular, a large number of pairwise measures of similarity or dissimilarity have been developed for comparing two topologies on the same leafset. Among them are the methods based on counting the structural differences between the two trees in terms of path length, bipartitions or quartets for unrooted trees, clades or triplets for rooted trees [57], or those based on minimizing a number of rearrangements that disconnect and reconnect subpieces of a tree, such as nearest neighbour interchange (NNI), subtree-pruning-regrafting (SPR) or Tree-Bisection-Reconnection (TBR) moves [810]. While the latter methods are NP-hard [11], the former are typically computable in polynomial time. In particular, the Robinson-Foulds (RF) distance, defined in terms of bipartition dissimilarity for unrooted trees, and clade dissimilarity for rooted trees [12], can be computed in linear [13], and even sublinear time [14].

Despite several drawbacks such as lack of robustness (a small change in a tree may cause a disproportional change in the distance), skewed distribution [1517], and a lack of biological rationale, RF remains the most widely used measure, not only in phylogenetics, but also in other fields such as in linguistics. To increase robustness, improved versions of the RF distance have also been developed [11, 18].

In addition of being efficiently computable, RF has the merit of being a true metric. It was originally defined on unrooted trees, in terms of edit operations on the tree edges: the minimum number of edge contraction and extension needed to transform one tree into the other [19]. Interestingly, the same metric, expressed in terms of node deletion and insertion, has been widely used in the context of data featuring hierarchical dependencies, modeled as trees with labeled nodes. In this case, the standard Tree Edit Distance (TED) is defined in terms of a minimum cost path of node deletion, node insertion and node relabeling (label substitution) transforming one tree to the other, for two trees sharing the same set of node labels (i.e. each label is present exactly once in each tree). While the less constrained version of the problem on unordered labeled trees is NP-complete [20], most variants are solvable in polynomial time [2123].

Even though this kind of hierarchical node labeling has limited applicability for phylogenetic trees, other types of labeling can be used in the context of genetic data comparison. In the case of gene trees, it is important to identify the evolutionary event (duplication, speciation, transfer, etc) that has led to a given bifurcation. For example, information on duplication and speciation node labeling is provided for the trees of the Ensembl Compara database [24] (reconciled with TreeBest [25]). Therefore, being able to compare labeled phylogenies is important in the context of gene tree reconstruction and analysis.

This paper is the first effort towards extending the RF distance to labeled trees involving, in addition to edge contraction and extension (operations that can alternatively be defined as node insertion and deletion), a node substitution or “relabeling” operation. Importantly, our extended RF remains a metric in the mathematical sense.

While the formulation of the RF distance in terms of edit operations is known, the bipartition and clade formulations are often those that are used in the literature. Though similar, the three formulations present some differences depending on whether the trees are rooted or unrooted. We begin by making these differences explicit. We then explore some properties of the extended RF distance in the case of two labels (e.g. speciation and duplication). In particular, we show that, in contrast to the RF distance for unlabeled trees, an optimal edit path for labeled trees may involve contracting good edges, i.e. edges representing common bipartitions of the two compared trees, which makes the extended RF much harder to compute than the basic RF. We then explore various avenues for computing the extended RF. We give an exact algorithm for contracting “mixed subtrees”, i.e. subtrees with alternating labels, and a bounded heuristic for general trees that achieves a factor 2 approximation. In the following section, the heuristic is shown, on simulated datasets, to be efficient, by plotting the number of tree edits against the computed RF distance. Finally, we explore some avenues for improvement. All proofs are given in the Appendix.

Methods

We first start with notations and concepts, and then describe the Robinson Foulds distance and the extension to labeled trees.

Let T be a tree with a node set V(T) and an edge set E(T). Given a node x of T, the degree of x is the number of edges incident to x. We denote by L(T)V(T) the set of leaves of T, i.e. the set of nodes of T of degree one. A node of V(T)L(T) is called an internal node. A tree with a single internal node is called a star tree. An edge connecting two internal nodes is called an internal edge; otherwise, it is a terminal edge. Moreover, a rooted tree admits a single internal node r(T) considered as the root.

Let x and y be two nodes of a rooted tree T; y is an ancestor of x if y is on the path from x to the root (possibly y itself); y is a descendant of x if y is on the path from x to a leaf (possibly y itself) of T. For a rooted tree, we may write (x,y) for an edge between x and y where x is closer to the root. We say that y is a child of x. If T is unrooted, we call the set {y:{x,y}E(T)} the set of children of x (this is an unusual definition, but defining a notion of children for both rooted and unrooted trees will be useful later). For a rooted or an unrooted tree T, we denote by Ch(x) the set of children of an internal node x of T.

A tree T representing the evolution of a set \({\mathcal {L}}\/\) of entities (usually taxa or genes) is a tree with a one-to-one mapping between L(T) and \({\mathcal {L}}\/\). We simply write \({\mathcal {L}}\/ = L(T)\) and say that T is a tree for \({\mathcal {L}}\). An internal node represents an ancestral event (classically a speciation or a duplication) leading from one to many different entities. Moreover rooting a tree amounts to determining the common ancestor of all entities, i.e. determining the direction of evolution. Accordingly, internal nodes of an evolutionary tree (which are the trees considered in this paper) should be of degree at least 3, except the root which is of degree at least 2. An internal node xr(T) of a tree T is binary if and only if x is of degree 3 and r(T) is binary if and only if r(T) is of degree 2. A tree T is said binary if and only if all its internal nodes are binary.

A subtreeS of T is a tree such that V(S)V(T),E(S)E(T) and any edge of E(S) connects two nodes of V(S). A chain of T is a subtree C with a node set V(C)={x1,,xk} and an edge set E(C)={e1,,ek−1} such that for each 1≤ik,ei is incident to xi and xi+1.

If T is an unrooted tree, an unrooted version of T can just be T ignoring the root status of r(T). To avoid having nodes of degree two, we rather define the unrooted tree T’ corresponding to T as the unrooted tree obtained from T by adding a dummy leaf R and an edge e=(r(T),R).

For a rooted tree T, we denote by Tx the subtree of T rooted at xV(T), i.e. the subtree of T containing all the descendants of x. We call L(Tx) the clade of x. A clade is non-trivial if it corresponds to an internal node of T. We denote by \({\mathcal {C}}(T)\) the set of non-trivial clades of T. It can be seen as a subset of the power set of \({\mathcal {L}}\).

The bipartition of an unrooted tree T corresponding to an internal edge e={x,y} is the unordered pair of clades L(Tx) and L(Ty) where Tx and Ty are the two subtrees rooted respectively at x and y obtained by removing e from T. A bipartition is non-trivial if it corresponds to an internal edge of T, and trivial otherwise. We denote by \({\mathcal {B}}(T)\) the set of non-trivial bipartitions of T. Note that bipartitions are sometimes called splits in the literature.

The Robinson-Foulds distance

Definition 1

(edit operations) Two edit operations on the edges of a tree T (rooted or unrooted) are defined as follows:

  • Let e={x,y} be an internal edge of E(T). An edge contraction Cont(T,e) is an operation transforming the tree T into the tree T obtained from T by removing the edge e of T and identifying x and y; in other words, T is obtained by adding the edge {x,z} for each zCh(y){x}, and then removing y and its incident edges (including {x,y}).

  • Let x be a non-binary internal node of V(T) and \(X = \left \{y_{1}, \cdots, y_{t}\right \} \subsetneq Ch(x)\) be a subset of Ch(x) such that |X|≥2. A node extension Ext(T,x,X) is an operation transforming the tree T into the tree T obtained from T by removing the edges {x,yi}, for 1≤it, creating a node y and a new edge e={x,y} adjacent to x, and creating new edges {y,yi}, for 1≤it.

The function δ(T1,T2) assigning to each pair of rooted or each pair of unrooted trees the length of a minimum sequence of edit operations transforming T1 into T2 has been shown to be a metric, called the Edit distance or Robinson-Foulds distance between T1 and T2 [19].

For unrooted trees T1 and T2, this distance corresponds to the symmetric difference between the bipartitions of the two trees. More precisely, \(\delta \left (T_{1},T_{2}\right) = |{\mathcal {B}}\left (T_{1}\right) \setminus {\mathcal {B}}(T_{2})| + |{\mathcal {B}}(T_{2}) \setminus {\mathcal {B}}(T_{1})|\). In fact, to transform T1 into T2, edit operations are needed on bad edges representing bipartitions which are not shared by the two trees, i.e. edges of T1 (respec. T2) defining bipartitions in T1 (respec. T2) which are not in \({\mathcal {B}}(T_{2})\) (respec. in \({\mathcal {B}}(T_{1})\)). An edge which is not bad is said to be good. Terminal edges are always good.

In the case of rooted trees T1 and T2, the Robinson-Foulds distance, that we denote in this case δR(T1,T2), is usually defined in the literature as the symmetric difference between the clades of the two trees. More precisely, for two rooted trees T1 and \(T_{2}, \delta _{R}(T_{1},T_{2}) = |{\mathcal {C}}(T_{1}) \setminus {\mathcal {C}}(T_{2})| + |{\mathcal {C}}(T_{2}) \setminus {\mathcal {C}}(T_{1})|\).

The link between the distance defined in terms of clades (that we write δR) and the edit distance (that we write δ) has been established through the defined relation between the bipartition system (or split system) and the clade system (or cluster system) [26].

Although our extended distance is more likely useful for rooted trees, algorithmic analyses are simpler for unrooted trees, as in this case all internal nodes can be treated in the same way. Here, we make the link between the rooted and unrooted case, and then focus, for the rest of the paper, on unrooted trees.

Let Tr be a rooted version of an unrooted tree T, with a binary root. Denote by e1,e2 the two edges adjacent to r(Tr). As e1 and e2 define the same bipartition of \({\mathcal {B}}(T)\), these edges are either both good or both bad. These notations are used in the following lemma.

Lemma 1

(Link between rooted and unrooted trees) Let T1,T2 be two rooted trees, and T1′, T2′ be the corresponding unrooted trees. Then

$${\kern1.7cm}\delta_{R}(T_{1},T_{2}) = \delta(T'_{1},T'_{2}) $$

The edit distance between two trees (rooted or unrooted) can be computed in linear time with the algorithm proposed by Day [13] in 1984. Our goal is to extend this distance to labeled trees.

Labeled trees

Given a finite set of labels Λ, T is labeled if and only if each internal node x of T has a unique label λ(x)Λ.

Contraction and extension operations are generalized to labeled trees as follows: The node y created from an edge extension Ext(T,x,X) is such that λ(y)=λ(x); an edge contraction is only defined on edges {x,y} for which λ(x)=λ(y). It follows that a third edit operation should be introduced for labeled trees. Let x be a node of a labeled tree T with label λ=λ(x). A node flip Flip(x,λ) is an operation assigning a new label λ to x, i.e. a label λΛ such that λλ. Those operations are depicted in Fig. 1.

Fig. 1
figure 1

The three edit operations defined for labeled trees. From left to right: Flip, Contraction and Extension

A node flip is required before contracting a mixed edge, i.e. an edge with its two extremities being differently labeled. A tree is said to be a mixed tree if all its edges are mixed edges.

Let \({\mathcal {T}}\) be the set of trees on \({\mathcal {L}}\), all trees being of the same type, i.e. all rooted or unrooted, all labeled or unlabeled. The following lemma (holding for all these cases) shows that introducing the flip operation does not prevent δ from being a distance.

Lemma 2

(Edit distance) The function δ(T1,T2) assigning to each pair \(\left (T_{1}, T_{2}\right) \in {\mathcal {T}}^{2}\) the minimum length of a sequence of edit operations transforming T1 into T2 defines a distance on \({\mathcal {T}}\).

In this paper, Λ is restricted to two labels. They are illustrated by a circle and a square in Fig. 2. The two labels can, for example, represent speciation and duplication events. Notice however that labeling is not constrained to be consistent with a species tree [27, 28]. In other words, the intermediate trees in an optimal path transforming a tree to another are not required to be feasible according the speciation/duplication labeling. Algorithmic analyses are made independently of the nature of the two node labels. However, for notation purpose, we write Λ={Spe,Dup}.

Fig. 2
figure 2

Two unrooted and labeled trees T and T on \({\mathcal L} = \{A, B, C, D, E, F, G\}\). The square and circle symbols represent the two possible labels for an internal node. Bad edges are red and good ones are black. {S1,S2} are the maximal bad subtrees of T and {S1′,S2′} the corresponding subtrees of T

Results

Consider \({\mathcal {T}}\) as the set of unrooted and labeled trees on \({\mathcal L}\). The goal is to compute the edit distance δ(T,T) for any pair T,T of trees of \({\mathcal {T}}\), that is the number of operations in an optimal sequence, i.e a sequence of edit operations of minimum length transforming T into T.

Note that although we focus on unrooted trees, our results can then be easily extrapolated to rooted trees using Lemma ??,

Reduction to maximal bad subtrees

Let S be a subtree of T. Let {ei={xi,yi}, for 1≤ik} be the set of terminal edges of S, with each yi being a leaf of S, and {Xi,Yi} being the bipartition corresponding to ei. Each leaf yi of S is said to be mapped to Yi. Notice that \(\cup _{1 \leq i \leq k}Y_{i} = {\mathcal L}\).

We say that S is a bad subtree of T if and only if S contains only bad edges, except the terminal edges of S which are all good edges of T. In other words, S is maximal in the sense that no more bad internal edges can be added into it. Intuitively, S can be obtained by taking a subtree with only bad edges, and adding edges adjacent to bad edges of S iteratively until the process stops. As a result, every terminal edge ei of S will be good, i.e. there is an edge \(e^{\prime }_{i} = \left \{x'_{i}, y'_{i}\right \}\) in T corresponding to ei={xi,yi}, that determine the same bipartition {Xi,Yi}. Note that a maximal bad subtree may contain no bad edge at all (i.e. it is a star tree centered on good edges).

Lemma 3

(Pairs of maximal bad subtrees) Let S be a maximal bad subtree of T with the set {ei}1≤ik of terminal edges, and let {ei′}1≤ik be the corresponding set of edges in T. Then the subtree S of T, containing all \(e^{\prime }_{i}\) edges as terminal edges, is unique. Moreover, it is a maximal bad subtree of T.

Let {S1,S2,,Sk} be the set of maximal bad subtrees of T and {S1′,S2′,,Sk′} be the corresponding subtrees of T (see Fig. 2 for an example). For 1≤im, let \({\mathcal {P}}_{i}\) be an optimal sequence transforming Si into \(S^{\prime }_{i}\). Then the sequence \({\mathcal {P}}\) obtained by performing consecutively \({\mathcal {P}}_{1}, {\mathcal {P}}_{2}, \cdots, {\mathcal {P}}_{m}\) transforms T into T.

Although the traditional RF distance can be deduced from the above observation, in our case such a sequence is not necessarily optimal. In fact, in contrast with unlabeled trees, optimal sequences for labeled trees may involve contracting good edges, as illustrated in Fig. 3.

Fig. 3
figure 3

Example where the minimal edit path requires contracting a good edge: if we contract the internal good edge of T (the bold one), then the 3 subtrees of T can be handled together, requiring 6 node flips and 18 edge contractions to reduce T into a star tree, and then 18 edge extensions to reach T, leading to 42 operations in total. By contrast, if we do not contract the good edge of T, then the two subtrees of T separated by this edge should be handled separately, requiring 9 flips, 17 edge contractions and 17 edge extensions to reach T, leading to 43 operations in total. The first scenario is the better one

Reduction to mixed bad subtrees

In the next section, we will describe an exact algorithm for optimally contracting a mixed tree. Before reaching this step, the question is how to obtain such a tree. The next lemma shows that non-mixed bad edges can be contracted first. The idea of the proof is that any optimal solution must eventually contract a non-mixed bad edge {x,y}. We can thus contract {x,y} first into a single node z, and “reproduce” all the events of the optimal solution by treating z as either x or y.

Lemma 4

(Contract non-mixed bad edges) Let e be any non-mixed bad edge of T, and let Tc be the tree obtained from T by contracting e. Then δ(Tc,T)=δ(T,T)−1.

According to this lemma, we can safely start by contracting all non-mixed bad edges of T and T first, since there is always an optimal sequence of edit operations that also does this. The resulting trees Tc and \(T^{\prime }_{c}\) can then be subdivided into pairs of maximal bad subtrees, all such bad subtrees being mixed subtrees.

Algorithms

We first consider a general framework which entails performing all required edge contractions first, and then all node extensions.

This general framework leads to the following upper bound for δ(T,T).

Lemma 5

(Upper bound δ) Let T and T be two unrooted and labeled trees with n internal nodes each and let e (resp. e) be the number of internal bad edges of T (resp. T). Then δ(T,T)≤e+e+n.

Notice that if both T and T are binary, then e=e. Moreover, in this case 2e+n is a tight bound as it can be reached in some cases (see an example in Fig. 4).

Fig. 4
figure 4

A pair of unrooted mixed trees (T,T), both with eight internal edges and nine internal nodes. All their internal edges are bad edges (red edges). Here δ(T,T)=25=2·8+9=2·e+n

The first step of Methodology 1 leads to a star tree T. Instead of then extending nodes to reach T, a symmetric way would be to transform T into a star tree \(T^{\prime }_{*}\). The difference between T and \(T^{\prime }_{*}\) may be in the label of the single node of each of these trees, which would then need an additional flip operation to reconstruct a corresponding path from T to T. This second methodology is given below, where Contract-Tree (T,T) takes as input a tree T and returns a sequence of operations contracting a tree T, i.e. transforming T into a star tree, and the star tree T resulting from this optimal contraction.

Methodology 2 is clearly simpler to handle and will be explored in the next section. The next lemma shows that it may overestimate an optimal sequence returned by Methodology 1 by at most one operation for each pair of maximal bad subtrees.

Lemma 6

(Compare Meth.1 and Meth.2) Let S and S be a pair of maximal bad subtrees of Tc and \(T^{\prime }_{c}\), obtained similarly by Methodology 1 and Methodology 2. Let M1(S,S) (respec. M2(S,S)) be the number of operations performed by the for loop of Methodology 1 (respec. Methodology 2). Moreover, let S (respec. \(S^{\prime }_{*}\)) be the star tree returned by Contract-Tree on S (respec. on S).

  1. 1

    If S=S′ (same node label), then M2(S,S)=M1(S,S);

  2. 2

    Otherwise, M1(S,S)≤M2(S,S)≤M1(S,S)+1

An optimal algorithm for contracting a tree

The remaining problem is the one of finding an optimal sequence of contraction and flip operations contracting a mixed tree T. For any such sequence, the number of contraction operations is just the number of internal edges of T. Therefore, the problem reduces to finding the minimum number of flip operations ϕ(T) in such an optimal sequence. Notice that the problem does not reduce to performing the minimum number of flips leading to the same label for all nodes, which would just be min{nbspe,nbdup} with nbspe (respec. nbdup) being the number of Spe (respec. Dup) nodes of T. For example, for the tree T of Fig. 3, min{nbspe,nbdup}=9. However, proceeding by an alternating sequence of flip and contraction operations (the top node flipped to Dup, then the three top edges contracted, then the next top node flipped to a Spe node, then the three top edges contracted, etc.) leads to a total of 6 flips rather than 9.

We will proceed iteratively by starting a sequence of contraction operations from the center of a tree T, i.e. the midpoint of the longest mixed chain of T. The diameter, denoted diam(T), of a tree T is the length of its longest chain (determined in terms of the number of edges). Note that any longest chain in a tree has two leaves at its extremities, as otherwise we could extend the chain. Assume that T has at least two terminal edges, so that diam(T)≥2. We show that ϕ(T) is equal to diam(T)/2−1. For a node v, let eccT(v) denote the maximum distance from v to a leaf of T (this is known as the eccentricity of v).Footnote 1

Lemma 7

(Optimal path contracting a mixed tree) The minimum number of flips in an optimal sequence of operations transforming a mixed tree T into a star tree is diam(T)/2−1.

Lemma ?? immediately lead to Algorithm Contract-Tree. The fact that the algorithm contracts T into a star tree using ϕ(T) flips follows from the proof of Lemma ??.

Theorem 1

For T being a mixed tree, Algorithm Contract-Tree returns the length of an optimal sequence of operations contracting T.

One should note that if T has even diameter, then there are two possible midpoints, i.e. two nodes with minimum eccentricity. This means that it is possible to choose the label of the internal node of the resulting star tree. This guarantees that when contracting a pair of bad subtrees T and T, we can always avoid a final flip by choosing the appropriate final label if either T or T has even diameter. We cannot guarantee that this final flip is avoidable if both subtrees have odd diameter.

We now show that Methodology 2 has a guaranteed approximation ratio of 2 when using Algorithm Contract-Tree as a subroutine. The idea behind the approximation is to show that any optimal solution must contract all the bad edges and perform at least one flip or good edge contraction per bad subtree. Our algorithm only contracts bad edges, and we can show that the number of flips performed is at most the number of bad edges plus twice the number of bad subtrees.

Theorem 2

(Upper bound Meth.2) Let d be the number of operations performed by Methodology 2 when tree contractions are done by Algorithm Contract-Tree. Then d≤2δ(T,T).

Experimental results

We implemented a heuristic following Methodology 2, using the Contract-Tree algorithm. To test it on simulated data, we retrieved the TP53 gene family from Ensembl release 96 (542 genes), including the speciation and duplication labels, and introduced an increasing number of random edit operations, on 30 replicates. A random edit was introduced as follows: with probability 0.3, the label of one random internal node was flipped; the rest of the probability mass function was evenly distributed among all internal edges connecting nodes of the same type (which could be potentially contracted) and all nodes of degree >3 (in which a new edge could potentially be expanded).

After each edit, we computed the classical RF distance and its extension to labeled trees using our heuristic (Fig. 5). Because it accounts for labels, the latter tracked more closely the true number of edits. At the same time, the estimated distances were never higher than the actual number of edits, which suggests that the heuristic can identify a minimum edit path when the total number of edit operations is relatively low. The implementation, including the function to mutate labeled trees, is available as an open source Python library (PyPI package pylabeledrf, also available at https://github.com/DessimozLab/pylabeledrf).

Fig. 5
figure 5

Empirical comparison of the distance inferred for an increasing number of random edit operations (contraction, extensions, and flips), using the classical Robinson-Foulds distance (left) and Contract-Tree algorithm (right). Because the former ignores node labels, it grossly underestimates the actual number of edits. Our algorithm tracks more closely the actual number of edits

Discussion

In this paper, we have considered what we thought was the simplest and most natural extension of the Robinson-Foulds distance to labeled trees. Although its theoretical complexity is unknown and remains an open problem, this extension appears to be much harder to compute than the classical RF distance for unlabeled trees.

Despite the optimality of Algorithm Contract-Tree for contracting a mixed tree, neither Methodology 1, nor Methodology 2 are guaranteed to lead to an optimal solution. This is due to two main reasons. The first one is that, as shown in Fig. 3, an optimal path contracting a tree T may require contracting good edges, i.e. edges common to both trees, which is not the case for unlabeled trees. The second reason is that an optimal path from a tree T to a tree T may not be one with all edge contraction events preceding all edge extension. An example, given in Fig. 6, shows that it may be better to convert a given bad edge into a good edge rather than contracting all bad edges. It can be observed from this example that going from T to T following the red path entails performing a nearest-neighbour interchange (NNI) operation on the edge e of T. A future direction for improving the algorithm will be to consider such “safe” edges, i.e. edges admitting an NNI leading to a bipartition of the target tree.

Fig. 6
figure 6

An optimal path from T to T following Methodology 1 is depicted by black arrows and involves 6 operations. It is not optimal as another path, depicted by red arrows, involves only five operations. The path of length 3 from T to Ts acts on the safe edge, represented in orange. This path involves an edge contraction, an edge extension and a flip, leading to the good edge (red edge) in Ts

Still, we have implemented a heuristic which constitutes a better baseline solution to quantifying differences between labeled tree topologies than the conventional RF measure, which is blind to labels. For instance, this implementation could be useful in the context of orthology benchmarking, to compare inferred labeled trees with reference curated ones [29].

Conclusion

Looking ahead, we envision several potential future directions. We see potential in identifying the good edges that should be contracted and characterizing classes of trees that may be resolved optimally. In particular, it would be interesting to restrict the study to the class of labeled trees consistent with a species tree (which is not the case of the trees of Fig. 3).

Another direction would be to consider an alternative extension of the RF distance. In this paper, edge contraction and edge extension, the two edit operations defining the classical RF, were re-defined in the context of labeled nodes, by constraining them to occur on edges with the same labels on their extremities. Another direction would be to consider edit operations on nodes, as for the Tree Edit Distance (TED) for hierarchical trees, i.e. node deletion, insertion and relabeling. In addition to the theoretical complexity and computational efficiency, it would be important to evaluate the robustness of these two RF extensions with respect to small changes in the topology or tree labeling. Although we do not expect robustness to be much better than the classical RF, knowing which extension is better can orient the study towards future improvements. Finally another direction would be to extend the study to an arbitrary set of possible labels.

More generally, we think that computing the distance between labeled trees conceals many new problems and opens a variety of new algorithmic directions.

Appendix

Proof of Lemma ?? (Link between rooted and unrooted trees)

Let T1 and T2 be two rooted trees and \(T^{\prime }_{1}\) and \(T^{\prime }_{2}\) be the corresponding unrooted trees, i.e. V(T1′)=V(T1){R},V(T2′)=V(T2){R},E(T1′)=E(T){(r(T1),R)} and E(T2′)=E(T){(r(T2),R)}.

We first show that any bad bipartition of \(T^{\prime }_{1}\), i.e. any bad edge of \(T^{\prime }_{1}\), corresponds to a bad clade of T1 (a clade which is not present in T2). Let \(e^{\prime }_{1}\) be a bad edge of \(T^{\prime }_{1}\). Then \(e^{\prime }_{1}\) should be a non-terminal edge of \(T^{\prime }_{1}\), thus different from (r(T1),R)), and therefore it has a corresponding edge e1=(x1,y1) in T1. Then, for one of the two nodes adjacent to \(e^{\prime }_{1}\) that we denote y1′, we have \(L(T'_{1y'_{1}}) = L(T_{1y_{1}})=C\). If e1′ is a bad edge of \(T^{\prime }_{1}\), then C should be a bad clade of T1 not present in T2. This is because otherwise C would be a non-trivial clade of T2 rooted at an internal node y2 adjacent to an edge e2=(x2,y2) and thus also equal to \(L(T'_{2y'_{2}})\) for a given edge e2′=(x2′,y2′). This contradicts the fact that \(e^{\prime }_{1}\) is a bad edge. Therefore, each bad bipartition of \(T^{\prime }_{1}\) corresponds to a bad clade of T1. Moreover, two disjoint bad bipartitions of \(T^{\prime }_{1}\) correspond to two different bad edges of \(T^{\prime }_{1}\), with the corresponding edges of T1 associated to two disjoint clades. Thus we have \(|\mathcal {B}(T'_{1})| \leq |\mathcal {C}(T_{1})|\).

Conversely, a bad clade C of T1 corresponds to an internal node y1 of T1. Let e1=(x1,y1) in T1, where x1 is the parent of y1. Then the corresponding edge \(e^{\prime }_{1}\) in \(T^{\prime }_{1}\) is a bad edge. Moreover, two disjoint clades of T1 correspond to two disjoint edges of T1′. It follows that \(|\mathcal {C}(T_{1})| \leq |\mathcal {B}(T'_{1})|\). Combining this result with the result above, we deduce that \(|\mathcal {C}(T_{1})| = |\mathcal {B}(T'_{1})|\). As T2 and \(T^{\prime }_{2}\) can be considered similarly, the result follows.

Proof of Lemma ?? (Edit distance):

The non-negative and identity conditions are obvious. For the symmetric condition, notice that we can reverse every edit operation in an optimal sequence from T1 to T2 to obtain a sequence from T2 to T1 with the same number of events, and vice-versa (extensions and contractions are inverses of each other, and any flip can be reversed by a flip). We thus have δ(T2,T1)≤δ(T1,T2) and δ(T1,T2)≤δ(T2,T1), and equality follows.

Finally, we prove the triangular inequality condition: for 3 trees T1,T2 and T3, to transform T1 into T2, we may take any edit sequence from T1 to T3, followed by any edit sequence from T3 to T2. It follows that δ(T1,T2)≤δ(T1,T3)+δ(T3,T2).

Proof of Lemma ?? (Pairs of maximal bad subtrees):

As iYi=Ł, \(\phantom {\dot {i}\!}\{e'_{i}\}_{1 \leq i \leq k}\) are the only terminal edges of any subtree S of T containing the set \(\phantom {\dot {i}\!}\{e'_{i}\}_{1 \leq i \leq k}\) as terminal edges. As T is a tree, for any 1≤ijk, there is only one possible path from \(x^{\prime }_{i}\) to \(x^{\prime }_{j}\). Uniqueness follows.

Suppose that such a subtree S is not a bad subtree. Then it contains an internal good edge e=(x,y). In other words, there is a non-trivial bipartition of {Yi}1≤ik which is also a bipartition in S. This contradicts the fact that S is a bad subtree of T. Finally, as all terminal edges of S are good edges of T, it follows that S is a maximal bad subtree of T.

Proof of Lemma ?? (Contract non-mixed bad edges):

We first introduce a definition that will be of use later in the proof. For two rooted trees S1 and S2, define the union of S1 and S2 as the tree obtained by identifying their roots, i.e. by removing the root of S2 and making all its children now children of the root of S1.

Let e={u,v} be a non-mixed bad edge and assume, without loss of generality, that both u and v have the label Spe (recall that Λ={Spe,Dup}). Notice that any sequence of operations turning T into T, at some point, must contract the {u,v} edge, as otherwise, the (bad) bipartition corresponding to {u,v} would remain in the transformed tree and we would not obtain T (noting that extensions cannot remove bipartitions). We now prove the Lemma by induction over δ(T,T). As a base case, suppose that δ(T,T)=1. Then {u,v} must be the only bad edge of T and the single operation is to contract it, proving the base case.

Now assume that for any tree \(\tilde {T}\) satisfying \(\delta (\tilde {T}, T') < \delta (T, T')\), contracting any non-mixed bad edge of \(\tilde {T}\) reduces its distance to T by 1. Let Q=(q1,…,ql) be an optimal sequence of operations transforming T into T (here each qi denotes either a contraction, extension or flip). Let qj be the event that contracts {u,v}. If q1=qj, then we are done, so assume otherwise. We make the assumption that whenever there is a contraction involving u prior to qj, the contracted node is still called u. Furthermore, we assume that if an extension prior to qj splits the neighbors of u, the node v is still a neighbor of u after the operation. All the same assumptions hold for v. This just changes the names we give to nodes and does not alter the scenario, but observe that this means that {u,v} is in every tree obtained before the first j operations.

For each i{1,…,l}, let Ti be the tree obtained after applying q1,…,qi on T, and define T0=T. Furthermore, for i{0,1,…,j−1}, denote by \(T^{u}_{i}\) and \(T^{v}_{i}\) the two trees obtained from Ti by removing the edge {u,v}, where u is in \(T^{u}_{i}\) and v is in \(T^{v}_{i}\). Define \(T^u = T^{u}_{0}\) and \(T^v = T^{v}_{0}\). We will assign u and v as the respective roots of each \(T^{u}_{i}\) and \(T^{v}_{i}\). Notice that for each i{1,…,j−1}, qi only modifies either the subtree \(T^{u}_{i-1}\) or \(T^{v}_{i-1}\). Therefore, if events qi and qi+1 modify \(T^{u}_{i-1}\) and \(T^{v}_{i}\), respectively, we could apply qi+1 before qi and Ti+1 would still be the same tree. This lets us assume that we may reorder events such that all events affecting Tu (prior to qj) occur before those affecting Tv. That is, there is some h such that q1,…,qh only affects the Tu subtree, qh+1,…,qj−1 only affects the Tv subtree, so that \(T^{u}_h = T^{u}_{h+1} = \ldots = T^{u}_{j-1}\) and \(T^v = T^{v}_1 = \ldots = T^{v}_{h}\).

Suppose first that u is labeled Spe in Th, and thus also in Tj−1. Then v is also labeled Spe in Tj−1 (and also in Th since v was untouched until qh+1). Let \(\hat {T}\) be the tree obtained after contracting {u,v} in T, and let z be the resulting node. Observe that if we interpret z as u, then we may apply the events q1,…,qh on \(\hat {T}\), since these events only affected the Tu subtrees. To be formal, we “reproduce” q1 through qh on \(\hat {T}\) by applying the events Q=(q1′,…,qh′) on \(\hat {T}\), defining \(\hat {T}_{i}\) as the tree obtained after the i-th event of Q, where each \(q^{\prime }_{i}\) in Q is defined as follows:

  • if qi contracts {x,y} in Ti−1, then \(q^{\prime }_{i}\) contracts {x,y} in \(\hat {T}_{i-1}\) if x,yu, otherwise if, say, x=u, then \(q^{\prime }_{i}\) contracts {z,y} (and calls the resulting node z);

  • if qi flips x in Ti−1, then \(q^{\prime }_{i}\) flips x in \(\hat {T}_{i-1}\) if xu, or flips z otherwise;

  • if qi is an extension and splits the neighborhood of x, then \(q^{\prime }_{i}\) does the same if xu (replacing u by z if needed). If x=u, then let X be the set of neighbors of v in Ti−1, excluding u. If Ch(u) is split into A and B by qi, where vB, then \(q^{\prime }_{i}\) splits the neighbors ABX of z into A and BX (and z is the neighbor of BX and the newly created node).

One can verify the following that the following invariant holds on each \(\hat {T_i}, i \in \{1, \ldots, h\}\): if we take Ti and contract the edge {u,v}, ignoring the labels and keeping the label of u, then we obtain \(\hat {T}_{i}\) (the invariant is also true for T and \(\hat {T}\)).

The resulting tree \(\hat {T}_{h}\) obtained from applying q1′,…,qh′ on \(\hat {T}\) will therefore contain z as a Spe node, and will be the union of \(T^{u}_{h}\) and \(T^{v}_{0}\). From this point, in a similar fashion, we may interpret z as v and apply qh+1,…,qj−1 on \(\hat {T}_{h}\), resulting a tree that is the union of \(T^{u}_h = T^{u}_{j-1}\) and \(T^{v}_{j-1}\). The corresponding events are the same as above, we omit the formal details. Since Tj is obtained from Tj−1 by contracting {u,v}, this means that \(\hat {T}_{j-1} = T_{j}\), which we have attained with j events but contracting {u,v} first, which proves this case.

Suppose instead that u is labeled Dup in Th. Then v is a Dup node in Tj−1. We may further assume that v is a Spe node in Th+1,…,Tj−2, since whenever we flip v into a Dup, we may assume by induction that {u,v} gets contracted. Therefore, qj−1 flips v from Spe to Dup, and for the first time. We may then do the following: first apply the events qh+1,…,qj−2 on \(\hat {T}\), interpreting z as v. The resulting tree \(\hat {T}'\) contains z as a Spe node, and is the union of \(T^{v}_{j-2}\) and \(T^{u}_{0}\). We may now apply q1,…,qh on \(\hat {T}'\) by interpreting u as z, resulting in a tree \(\hat {T}^{\prime \prime }\) that contains z as a Dup node and is the union of \(T^{u}_{h} = T^{u}_{j-1}\) and \(T^{v}_{j - 1}\). We have thus attained Tj, but this time without the qj−1 flip on v, contradicting the optimality of Q. This concludes the proof.

Proof of Lemma ?? (Upper bound δ):

Methodology 1 performs e contractions and e extensions. As for the number of flips, we have to flip at most all the nodes belonging to the smallest label group, which means at most half the nodes in each tree, and thus at most n flips in total.

Proof of Lemma ?? (Compare Meth.1 and Meth.2):

We denote by Cont(T) the minimum length of a sequence of operations contracting T, and by l() the length of a sequence of edit operations (Fig. 7).

Fig. 7
figure 7

Notations for the Proof of Lemma ??

Let 2 be an optimal sequence contracting S to S and 2′ be an optimal sequence contracting S to S′. As each operation is reversible, 2′ leads to a corresponding sequence 2′′ of the same length between S′ and S. Thus, 2, concatenated with a possible flip operation transforming S to \(S^{\prime }_{*}\), concatenated with 2′′ is a sequence from S to S following Methodology 1, and thus M1(S,S)≤M2(S,S) (R1).

Conversely, let be an optimal sequence following Methodology 1. Then this sequence can be subdivided into a sequence 1 from S to a star tree S1, and 1′ from S1 to S. As each operation is reversible, 1′ leads to a corresponding sequence 1′′ of the same length between S and S1. In other words, M1(S,S)=l(1)+l(1′)=l(1)+l(1′′)≥Cont(S)+Cont(S).

  1. 1.

    If S=S′, then M2(S,S)=Cont(S)+Cont(S) and thus M1(S,S)≥M2(S,S), and the result follows from (R1).

  2. 2.

    Otherwise, S and S′ are different and M2(S,S)=Cont(S)+Cont(S)+1. Thus M1(S,S)≥Cont(S)+Cont(S)=M2(S,S)−1, and thus M2(S,S)≤M1(S,S)+1.

Proof of Lemma ?? (Optimal path contracting a mixed tree):

We first show that at least diam(T)/2−1 flips are needed, by induction over the diameter of T. When diam(T)=2, T is a star tree and 0=diam(T)/2−1 flips are needed. For the induction step, we assume that any tree T with diam(T)<diam(T) requires at least diam(T)/2−1 flips. Take any optimal sequence of events S, and observe that in S, when we flip a node v of T, by Lemma ?? we may assume that S contracts all the incident edges to v until we obtain another mixed tree. Let T1,T2,…,Tk be the sequence of mixed trees encountered when applying S, i.e. each Ti is obtained after flipping a node and contracting its incident edges. Define T0=T. Let i be the smallest index such that diam(Ti)<diam(T). Then in Ti−1, there was a longest chain P=(u1,…,ul) of length diam(T). The flip-and-contract operations from Ti−1 to Ti can reduce the length of P by at most 2 since we flip one node and only its incident edges, of which there are at most two on P. Hence diam(Ti)≥diam(T)−2. We deduce by induction that the number of required flips is at least 1+(diam(T)−2)/2−1=diam(T)/2−1.

We now turn to the converse bound ϕ(T)≤diam(T)/2−1. Fix any node v of T, and suppose that we run the following procedure: as long as T is not a star tree, flip v and contract its incident internal edges. Since each flip-and-contraction iteration reduces the length from v to any leaf by 1 (except its neighbors), eccT(v) is reduced by 1 each round. We stop when eccT(v)=1, in which case only terminal edges remain, and in the end, this means that eccT(v)−1 flips are needed.

To see why this proves our bound, we show that there always exists a node with eccentricity diam(T)/2. Consider a longest chain P of T with nodes w1,…,wk. Observe that diam(T)=k−1 (recall that distances are counted in terms of edges). Consider a midpoint node w:=wk/2 on P. We claim that eccT(w)=diam(T)/2. It is easy to check that w has distance at most diam(T)/2 and at least diam(T)/2 to the leaves w1 and wk on P. Assume for contradiction that w is at distance at least diam(T)/2+1 from some leaf l of T not in P. Then either we can form a chain from w1 to w and then to l, or a chain from wk to w and then to l. This chain has length at least diam(T)/2+diam(T)/2+1>diam(T), a contradiction. This shows that eccT(w)=diam(T)/2 and concludes the proof.

Proof of Theorem ?? (Upper bound Meth.2):

Consider a given instance (T,T). Take any leaf of T and assign it as the root, and do the same for T. Although we have assumed roots of degree at least two so far, we use this rooting only for our analysis in order fix a parent-child relationship between nodes. Let Q be an optimal sequence of operations turning T into T. We may assume that Q first contracts every non-mixed edge, and our algorithm does the same. Therefore, we suppose that T and T contain no non-mixed edges. Assume for our purposes that whenever a contraction takes place in Q between a node u and a child v, the u node stays in the tree and v gets removed (here the notion of a child is in the rooted sense with respect to our rooting above). Also assume that when there is an extension splitting a node u, then the newly created node becomes a child of u and u retains the same parent. It is easily checked that this only alters the name of nodes and not the sequence itself.

Call an internal node v of T a good child if the edge between v and its parent is good. Note that v has a unique corresponding node in T which we denote v (i.e. v is the root of the same clade as the subtree rooted at v). Further, call v a bad-good child if v is a good child, but either the label of v differs from that of v, or v is incident to at least one bad edge (yes, children are capable of being both bad and good). Note that every bad subtree of T is rooted at a bad-good child, and observe that here we say that a bad-good child v that is incident to only good edges is a particular case of a bad subtree (i.e. v just has the wrong label).

We already know that δ(T,T) is at least the number of bad edges in T and T. Let Q be the set of operations of Q that are either flips, or contraction of good edges. We argue that |Q| is at least the number of bad-good children in T. To see this, let v be a bad-good child. Assume first that v is not incident to any bad edge. If we never flip v nor remove it by contracting its parent edge, then Q cannot transform T into T, as v and its underlying clade remain present in every tree from T to T, but with the wrong label (because a contraction not removing v cannot remove the v clade, and extensions can create clades but not remove them). So we may assume that v gets flipped or that its parent edge gets contracted. A flip must be in Q and, observing that at any point the parent edge of v must be good, a contraction removing v must also be in Q. Assume instead that v is incident to at least one bad edge {v,w}, with w a child of v. If v is never flipped nor removed owing to a contraction of its parent edge, then at some point w must be flipped so that the {v,w} edge gets contracted. Otherwise, if v gets removed, then its parent edge was contracted, again implying the contraction of a good edge. Either cases imply an operation in Q. Importantly, observe that the operations in Q identified above are all distinct, since each one implies a flip or a node removal of a node in a different bad subtree of T.

Now, let T1,…,Tk be the bad subtrees of T and T, and for each i{1,…,k}, let ti be the number of bad edges in Ti. Further denote \(b = \sum _{i=1}^k t_{i}\). Since bad subtrees form pairs, our arguments above imply that Q has at least k/2 operations (because |Q| is at least the number of bad trees in T, which is half the number of bad subtrees). The contraction of bad edges plus the operations of Q show that Q has at least \(\sum _{i = 1}^k t_i + k/2 = b + k/2\) operations. Our algorithm contracts b edges in total. To count the number of flips, take any bad subtree Ti. Then tidiam(Ti)−2 and the number of flips we perform is at most diam(Ti)/2−1=(diam(Ti)−2)/2ti/2+1. Note that this also holds when Ti contains no bad edge. Therefore, the number of operations that we perform is at most \(b + \sum _{i=1}^k (t_i/2 + 1) = 3b/2 + k\). Our approximation ratio is therefore \(\frac {3b/2 + k}{b + k/2} \leq \frac {2b + k}{b + k/2} = 2\).

Availability of data and materials

The software implemented during the current study is available in the pylabeledrf repository, https://github.com/DessimozLab/pylabeledrf.

Notes

  1. The radius of T is a well-known graph parameter and is defined as the minimum eccentricity of a node of T. In a tree, the radius turns out to be diam(T)/2.

Abbreviations

RF:

Robinson-Foulds

References

  1. Semple C, Steel M, et al. Phylogenetics vol. 24. Oxford: Oxford University Press on Demand; 2003.

    Google Scholar 

  2. Colijn C, Plazzotta G. A metric on phylogenetic tree shapes. Syst Biol. 2018; 67(1):113–26.

    Article  CAS  PubMed  Google Scholar 

  3. Lafond M, El-Mabrouk N, Huber KT, Moulton V. The complexity of comparing multiply-labelled trees by extending phylogenetic-tree metric. Theor Comput Sci. 2019; 760:15–34.

    Article  Google Scholar 

  4. Bryant D, Scornavacca C. An O(n logN) time algorithm for computing the path-length distance between trees. Algorithmica. 2019; 81(9):3692–706.

    Article  Google Scholar 

  5. Cardona G, Llabrés M, Rosselló F, Valiente G. Nodal distances for rooted phylogenetic trees. J Math Biol. 2010; 61(2):253–76.

    Article  PubMed  Google Scholar 

  6. Estabrook GF, McMorris F, Meacham CA. Comparison of undirected phylogenetic trees based on subtrees of four evolutionary units. Syst Zool. 1985; 34(2):193–200.

    Article  Google Scholar 

  7. Critchlow DE, Pearl DK, Qian C. The triples distance for rooted bifurcating phylogenetic trees. Syst Zool. 1996; 45(3):323–34.

    Google Scholar 

  8. Jiang BDXHT, Li M, Tromp J, Zhang L. On computing the nearest neighbor interchange distance. In: Discrete Mathematical Problems with Medical Applications: DIMACS Workshop Discrete Mathematical Problems with Medical Applications, December 8-10, 1999, DIMACS Center, vol. 55. Providence: American Mathematical Soc.: 2000. p. 125.

    Google Scholar 

  9. Hickey G, Dehne F, Rau-Chaplin A, Blouin C. Spr distance computation for unrooted trees. Evol Bioinforma. 2008; 4:419.

    Article  Google Scholar 

  10. Allen BL, Steel M. Subtree transfer operations and their induced metrics on evolutionary trees. Ann Comb. 2001; 5(1):1–15.

    Article  Google Scholar 

  11. Lin Y, Rajan V, Moret BM. A metric for phylogenetic trees based on matching. IEEE/ACM Trans Comput Biol Bioinforma (TCBB). 2012; 9(4):1014–22.

    Article  Google Scholar 

  12. Mittal S, Munjal G. Tree mining and tree validation metrics: A review. IOSR: J Comput Eng. 2015; 2:31–36.

    Google Scholar 

  13. Day WH. Optimal algorithms for comparing trees with labeled leaves. J Classif. 1985; 2(1):7–28.

    Article  Google Scholar 

  14. Pattengale ND, Gottlieb EJ, Moret BM. Efficiently computing the robinson-foulds metric. J Comput Biol. 2007; 14(6):724–35.

    Article  CAS  PubMed  Google Scholar 

  15. Steel MA, Penny D. Distributions of tree comparison metric–some new results. Syst Biol. 1993; 42(2):126–41.

    Google Scholar 

  16. Bryant D, Steel M. Computing the distribution of a tree metric. IEEE/ACM Trans Comput Biol Bioinforma. 2009; 6(3):420–6.

    Article  Google Scholar 

  17. Chaudhary R, Burleigh JG, Fernandez-Baca D. Fast local search for unrooted robinson-foulds supertrees. IEEE/ACM Trans Comput Biol Bioinforma (TCBB). 2012; 9(4):1004–13.

    Article  Google Scholar 

  18. Moon J, Eulenstein O. Cluster matching distance for rooted phylogenetic trees. In: International Symposium on Bioinformatics Research and Applications. Springer: 2018. p. 321–32. https://doi.org/10.1007/978-3-319-94968-0_31.

  19. Robinson DF, Foulds LR. Comparison of phylogenetic trees. Math Biosci. 1981; 53(1-2):131–47.

    Article  Google Scholar 

  20. Zhang K, Statman R, Shasha D. On the editing distance between unordered labeled trees. Inf Process Lett. 1992; 42(3):133–9.

    Article  Google Scholar 

  21. Zhang K. A new editing based distance between unordered labeled trees. In: Annual Symposium on Combinatorial Pattern Matching. Berlin: Springer: 1993. p. 254–65.

    Google Scholar 

  22. Zhang K. A constrained edit distance between unordered labeled trees. Algorithmica. 1996; 15(3):205–22.

    Article  CAS  Google Scholar 

  23. Schwarz S, Pawlik M, Augsten N. A new perspective on the tree edit distance. In: International Conference on Similarity Search and Applications. Springer: 2017. p. 156–70. https://doi.org/10.1007/978-3-319-68474-1_11.

  24. Vilella AJ, Severin J, Ureta-Vidal A, Heng L, Durbin R, Birney E. EnsemblCompara gene trees: Complete, duplication-aware phylogenetic trees in vertebrates. Genome Res. 2009; 19:327–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Schreiber F, Patricio M, Muffato M, Pignatelli M, Bateman A. Treefam v9: a new website, more species and orthology-on-the-fly. Nucleic Acids Res. 2013; 42(D1):D922–D925. https://doi.org/10.1093/nar/gkt1055.

    Article  PubMed  PubMed Central  Google Scholar 

  26. Dress A. Towards a theory of holistic clustering. DIMACS Ser Discrete Math Theoret Comput Sci. 1997; 37:271–89.

    Article  Google Scholar 

  27. Hernandez-Rosales M, Hellmuth M, Wieseke N, Huber KT, Moulton V, Stadler PF. From event-labeled gene trees to species trees. BMC Bioinformatics. 2012; 13:6. BioMed Central.

    Article  Google Scholar 

  28. Lafond M, El-Mabrouk N. Orthology and paralogy constraints: satisfiability and consistency. BMC Genomics. 2014; 15(6):12.

    Article  Google Scholar 

  29. Altenhoff AM, Boeckmann B, Capella-Gutierrez S, Dalquen DA, DeLuca T, Forslund K, Huerta-Cepas J, Linard B, Pereira C, Pryszcz LP, Schreiber F, da Silva AS, Szklarczyk D, Train C-M, Bork P, Lecompte O, von Mering C, Xenarios I, Sjölander K, Jensen LJ, Martin MJ, Muffato M, Quest for Orthologs consortium, Gabaldón T, Lewis SE, Thomas PD, Sonnhammer E, Dessimoz C. Standardized benchmarking in the quest for orthologs. Nature methods. 2016; 13(5):425–30. https://doi.org/10.1038/nmeth.3830.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The authors thank the Program Committee of APBC2020 for the three constructive peer-reviews.

About this supplement

This article has been published as part of BMC Genomics Volume 21 Supplement 10, 2020: Selected articles from the 18th Asia Pacific Bioinformatics Conference (APBC 2020): genomics. The full contents of the supplement are available online at https://bmcgenomics.biomedcentral.com/articles/supplements/volume-21-supplement-10.

Funding

Christophe Dessimoz is supported by the Swiss National Science Foundation (SNSF) Professorship grant 183723. Nadia El-Mabrouk is supported by Natural Sciences and Engineering Research Council of Canada (NSERC), and Fonds de Recherche Nature et Technologies of Quebec (FRQNT). Manuel Lafond is supported by Natural Sciences and Engineering Research Council of Canada (NSERC). Publication costs were funded by the Swiss National Science Foundation. The funding sources had no role in the design and conduct of the study, nor in the writing, review or approval of the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

SB, CD, NEM, and ML devised the proofs and algorithms and wrote the paper. CD and GL implemented the distance software introduced in the manuscript. All author(s) read and approved the final manuscript.

Corresponding authors

Correspondence to Christophe Dessimoz or Nadia El-Mabrouk.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Briand, S., Dessimoz, C., El-Mabrouk, N. et al. A generalized Robinson-Foulds distance for labeled trees. BMC Genomics 21 (Suppl 10), 779 (2020). https://doi.org/10.1186/s12864-020-07011-0

Download citation

  • Published:

  • DOI: https://doi.org/10.1186/s12864-020-07011-0

Keywords