This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Minimum 0-Extension Problems on Directed Metrics

Hiroshi HIRAI and Ryuhei MIZUTANI
Department of Mathematical Informatics,
Graduate School of Information Science and Technology,
The University of Tokyo, Tokyo, 113-8656, Japan.
[email protected]
[email protected]
(February 7, 2021)
Abstract

For a metric μ\mu on a finite set TT, the minimum 0-extension problem 0-Ext[μ][\mu] is defined as follows: Given VTV\supseteq T and c:(V2)+\ c:{\scriptsize\begin{pmatrix}V\\ 2\\ \end{pmatrix}}\rightarrow\mathbb{Q}_{+}, minimize c(xy)μ(γ(x),γ(y))\sum c(xy)\mu(\gamma(x),\gamma(y)) subject to γ:VT,γ(t)=t(tT)\gamma:V\rightarrow T,\ \gamma(t)=t\ (\forall t\in T), where the sum is taken over all unordered pairs in VV. This problem generalizes several classical combinatorial optimization problems such as the minimum cut problem or the multiterminal cut problem. Karzanov and Hirai established a complete classification of metrics μ\mu for which 0-Ext[μ][\mu] is polynomial time solvable or NP-hard. This result can also be viewed as a sharpening of the general dichotomy theorem for finite-valued CSPs (Thapper and Živný 2016) specialized to 0-Ext[μ][\mu].

In this paper, we consider a directed version 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] of the minimum 0-extension problem, where μ\mu and cc are not assumed to be symmetric. We extend the NP-hardness condition of 0-Ext[μ][\mu] to 0\overrightarrow{\textbf{0}}-Ext[μ][\mu]: If μ\mu cannot be represented as the shortest path metric of an orientable modular graph with an orbit-invariant “directed” edge-length, then 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] is NP-hard. We also show a partial converse: If μ\mu is a directed metric of a modular lattice with an orbit-invariant directed edge-length, then 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] is tractable. We further provide a new NP-hardness condition characteristic of 0\overrightarrow{\textbf{0}}-Ext[μ][\mu], and establish a dichotomy for the case where μ\mu is a directed metric of a star.

1 Introduction

A metric on a finite set TT is a function μ:T×T+\mu:T\times T\rightarrow\mathbb{R}_{+} that satisfies μ(x,x)=0,μ(x,y)=μ(y,x),\mu(x,x)=0,\ \mu(x,y)=\mu(y,x), and μ(x,y)+μ(y,z)μ(x,z)\mu(x,y)+\mu(y,z)\geq\mu(x,z) for every x,y,zTx,y,z\in T, and μ(x,y)>0\mu(x,y)>0 for every xyTx\neq y\in T. For a rational-valued metric μ\mu on TT, the minimum 0-extension problem 0-Ext[μ][\mu] on μ\mu is defined as follows:

0-Ext[μ]:Instance:\displaystyle\textbf{0-Ext}[\mu]\mathrm{:\ \ \ \ }\mathrm{Instance:} VT,c:(V2)+\displaystyle\mathrm{\ \ }V\supseteq T,\ c:\begin{pmatrix}V\\ 2\\ \end{pmatrix}\rightarrow\mathbb{Q}_{+}
Min.\displaystyle\mathrm{Min.} xy(V2)c(xy)μ(γ(x),γ(y))\displaystyle\displaystyle\sum_{\tiny xy\in\begin{pmatrix}V\\ 2\\ \end{pmatrix}\normalsize}c(xy)\mu(\gamma(x),\gamma(y))
s.t.\displaystyle\mathrm{s.t.} γ:VTwithγ(t)=tforalltT,\displaystyle\mathrm{\ \ }\gamma:V\rightarrow T\ \mathrm{with\ }\gamma(t)=t\mathrm{\ for\ all\ }t\in T,\mathrm{\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ } (1.1)

where (V2){\scriptsize\begin{pmatrix}V\\ 2\\ \end{pmatrix}} denotes the set of all unordered pairs of VV, and xyxy denotes the unordered pair consisting of x,yVx,y\in V. The minimum 0-extension problem was introduced by Karzanov [13], and also known as the multifacility location problem in facility location theory [17]. Note that the formulation (1) of 0-Ext[μ][\mu] is different from but equivalent to that of [13].

The minimum 0-extension problem generalizes several classical combinatorial optimization problems: If T={s,t}T=\{s,t\}, then 0-Ext[μ][\mu] is nothing but the minimum ss-tt cut problem in an undirected network. If T={x,y,z}T=\{x,y,z\} and μ(x,y)=μ(y,z)=μ(z,x)=1\mu(x,y)=\mu(y,z)=\mu(z,x)=1, then 0-Ext[μ][\mu] is the 33-terminal cut problem. Similarly, 0-Ext[μ][\mu] can formulate the kk-terminal cut problem. Moreover, 0-Ext[μ][\mu] appears as a discretized LP-dual problem for a class of maximum multiflow problems [12, 13] (also see [8, 9]).

The computational complexity of 0-Ext[μ][\mu] depends on metric μ\mu. In the above examples, the minimum ss-tt cut problem is in P and the 33-terminal cut problem is NP-hard. In [13], Karzanov addressed the classification problem of the computational complexity of 0-Ext[μ][\mu] with respect to μ\mu. After [5, 15], the complexity dichotomy of 0-Ext[μ][\mu] was fully established by Karzanov [14] and Hirai [10], which we explain below.

A metric μ\mu on TT is called modular if for every s0,s1,s2Ts_{0},s_{1},s_{2}\in T, there exists an element mTm\in T, called a median, such that μ(si,sj)=μ(si,m)+μ(m,sj)\mu(s_{i},s_{j})=\mu(s_{i},m)+\mu(m,s_{j}) holds for every 0i<j20\leq i<j\leq 2. The underlying graph of μ\mu is defined as the undirected graph Hμ=(T,U)H_{\mu}=(T,U), where U={{x,y}x,yT(xy),zT{x,y},μ(x,y)<μ(x,z)+μ(z,y)}U=\{\{x,y\}\mid x,y\in T\ (x\neq y),\ \forall z\in T\setminus\{x,y\},\ \mu(x,y)<\mu(x,z)+\mu(z,y)\}. We say that an undirected graph is orientable if it has an edge-orientation such that for every 4-cycle (u,v,w,z,u)(u,v,w,z,u), the edge {u,v}\{u,v\} is oriented from uu to vv if and only if the edge {w,z}\{w,z\} is oriented from zz to ww.

The dichotomy theorem of the minimum 0-extension problem is the following:

Theorem 1.1 ([14]).

Let μ\mu be a rational-valued metric. 0-Ext[μ][\mu] is strongly NP-hard111A problem 𝒫\mathcal{P} is called strongly NP-hard if 𝒫\mathcal{P} is still NP-hard when all numbers of the instance are bounded by some polynomial in the length of the instance. if

  1. (i)

    μ\mu is not modular, or

  2. (ii)

    HμH_{\mu} is not orientable.

Theorem 1.2 ([10]).

Let μ\mu be a rational-valued metric. If μ\mu is modular and HμH_{\mu} is orientable, then 0-Ext[μ][\mu] is solvable in polynomial time.

The minimum 0-extension problem constitutes a fundamental class of valued CSPs (valued constraint satisfaction problem) [10] — a minimization problem of a sum of functions having a constant number of variables. More concretely, 0-Ext[μ][\mu] is precisely the finite-valued CSP generated by a single binary function μ:T×T+\mu:T\times T\rightarrow\mathbb{Q}_{+} that is a metric. Thapper and Živný [19] established a P-or-NP-hard dichotomy theorem for finite-valued CSPs in terms of a certain fractional polymorphism, and moreover, gave a polynomial time algorithm for the meta problem of deciding whether a given template (meaning μ\mu in the case of 0-Ext[μ][\mu]) defines tractable valued CSPs. Their framework is so general and powerful as to be applicable to 0-Ext[μ][\mu], which also provides the complexity dichotomy property of 0-Ext[μ][\mu] and a polynomial time checking of a tractable metric μ\mu for 0-Ext[μ][\mu]. On the other hand, it does not unravel which combinatorial or geometric properties of metric μ\mu are connected to the tractability of 0-Ext[μ][\mu]. Although the proof of Theorem 1.2 also utilized a related polymorphism condition obtained by Kolmogorov, Thapper, and Živný [16], it required a deep study on modular metric spaces to verify this condition. In addition to geometric insights, such an explicit tractability characterization as in Theorems 1.1 and 1.2 brings an efficient combinatorial algorithm checking the 0-Ext-tractability of metric μ\mu, i.e., the meta problem for 0-Ext[μ][\mu]. Indeed, we can verify modularity of μ\mu by checking whether mm is a median of triple t1,t2,t3t_{1},t_{2},t_{3} for every m,t1,t2,t3Tm,t_{1},t_{2},t_{3}\in T. We can also verify the orientability of HμH_{\mu} by orienting each edge in depth first order with respect to an adjacency relation such that edges {u,v}\{u,v\} and {z,w}\{z,w\} in each 4-cycle (u,v,w,z,u)(u,v,w,z,u) are said to be adjacent. It is an O(|T|4)O(|T|^{4})-time algorithm. On the other hand, a polynomial time algorithm obtained by the framework of [19] requires repeated uses of the ellipsoid method. So it is a natural direction to seek such an efficient combinatorial characterization for a more general binary function μ:T×T+\mu:T\times T\rightarrow\mathbb{Q}_{+} for which the corresponding valued CSP is tractable.

Motivated by these facts, in this paper, we consider a directed version of the minimum 0-extension problem, aiming to extend the above results. Here, by “directed” we mean that symmetry of μ\mu and cc is not assumed. A directed metric on a finite set TT is a function μ:T×T+\mu:T\times T\rightarrow\mathbb{R}_{+} that satisfies μ(x,x)=0\mu(x,x)=0 and μ(x,y)+μ(y,z)μ(x,z)\mu(x,y)+\mu(y,z)\geq\mu(x,z) for every x,y,zTx,y,z\in T, and μ(x,y)+μ(y,x)>0\mu(x,y)+\mu(y,x)>0 for every xyTx\neq y\in T. For a rational-valued directed metric μ\mu on TT, the directed minimum 0-extension problem 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] on μ\mu is defined as follows:

0-Ext[μ]:Instance:\displaystyle\overrightarrow{\textbf{0}}\textbf{-Ext}[\mu]\mathrm{:\ \ \ \ }\mathrm{Instance}: VT,c:V×V+\displaystyle\mathrm{\ \ }V\supseteq T,\mathrm{\ }c:V\times V\rightarrow\mathbb{Q}_{+}
Min.\displaystyle\mathrm{Min.} (x,y)V×Vc(x,y)μ(γ(x),γ(y))\displaystyle\displaystyle\sum_{\tiny(x,y)\in V\times V\normalsize}c(x,y)\mu(\gamma(x),\gamma(y))
s.t.\displaystyle\mathrm{s.t.} γ:VTwithγ(t)=tforalltT.\displaystyle\mathrm{\ \ }\gamma:V\rightarrow T\mathrm{\ with\ }\gamma(t)=t\mathrm{\ for\ all\ }t\in T.\mathrm{\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ }

The minimum ss-tt cut problem on a directed network is a typical example of 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] in the case of T={s,t},μ(s,t)=1,T=\{s,t\},\ \mu(s,t)=1, and μ(t,s)=0\mu(t,s)=0. Also, the directed minimum 0-extension problem contains the undirected version. Hence, the complexity classification of the directed version is an extension of that of the undirected version.

In this paper, we explore sufficient conditions for which 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] is tractable, and for which 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] is NP-hard. Our first contribution is an extension of Theorem 1.1 to the directed version:

Theorem 1.3.

Let μ\mu be a rational-valued directed metric. 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] is strongly NP-hard if one of the following holds:

  1. (i)

    μ\mu is not modular.

  2. (ii)

    HμH_{\mu} is not orientable.

  3. (iii)

    μ\mu is not directed orbit-invariant.

The modularity and the underlying graph HμH_{\mu} of a directed metric μ\mu are natural extensions of those of a metric. In 0-Ext[μ][\mu], the condition (i) in Theorem 1.1 contains the condition (iii) in Theorem 1.3. See Section 3 for the precise definitions of the terminologies.

We next consider the converse of Theorem 1.3. It is known [1] that a canonical example of a modular metric is the graph metric of the covering graph of a modular lattice with respect to an orbit-invariant edge-length. Moreover, a tractable metric μ\mu in Theorem 1.2 is obtained by gluing such metrics of modular lattices [10]. It turns out in Section 3 that a directed metric excluded by (i), (ii), and (iii) in Theorem 1.3 also admits an amalgamated structure of modular lattices. Our second contribution is the tractability for the building block of such a directed metric.

Theorem 1.4.

Let μ\mu be a rational-valued directed metric. Suppose that HμH_{\mu} is the covering graph of a modular lattice and μ\mu is directed orbit-invariant. Then 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] is solvable in polynomial time.

See Sections 2 and 3 for the undefined terminologies.

The converse of Theorem 1.3 is not true: Even if HμH_{\mu} is a tree (that is excluded by (i), (ii), and (iii) in Theorem 1.3), 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] can be NP-hard. On the other hand, 0-Ext[μ][\mu] for which HμH_{\mu} is a tree is always tractable (see [17]). This is a notable difference between 0-Ext[μ][\mu] and 0\overrightarrow{\textbf{0}}-Ext[μ][\mu]. Our third contribution is a new hardness condition capturing this difference. For x,yTx,y\in T, let Iμ(x,y):={zTμ(x,y)=μ(x,z)+μ(z,y)}I_{\mu}(x,y):=\{z\in T\mid\mu(x,y)=\mu(x,z)+\mu(z,y)\}, which is called the interval from xx to yy. We denote I:=IμI:=I_{\mu} if μ\mu is clear in the context. For x,yTx,y\in T, the ratio Rμ(x,y)R_{\mu}(x,y) from xx to yy is defined by Rμ(x,y):=μ(x,y)/μ(y,x)R_{\mu}(x,y):=\mu(x,y)/\mu(y,x) (if μ(y,x)=0\mu(y,x)=0, then Rμ(x,y):=R_{\mu}(x,y):=\infty). A pair (x,y)(T2)(x,y)\in{\scriptsize\begin{pmatrix}T\\ 2\\ \end{pmatrix}} is called a biased pair if Rμ(x,z)>Rμ(z,y)R_{\mu}(x,z)>R_{\mu}(z,y) holds for every zI(x,y)I(y,x){x,y}z\in I(x,y)\cap I(y,x)\setminus\{x,y\}, or Rμ(x,z)<Rμ(z,y)R_{\mu}(x,z)<R_{\mu}(z,y) holds for every zI(x,y)I(y,x){x,y}z\in I(x,y)\cap I(y,x)\setminus\{x,y\}. A triple (s0,s1,s2)(s_{0},s_{1},s_{2}) is called a non-collinear triple if siI(si1,si+1)I(si+1,si1)s_{i}\notin I(s_{i-1},s_{i+1})\cap I(s_{i+1},s_{i-1}) holds for every i{0,1,2}i\in\{0,1,2\} (the indices of sis_{i} are taken modulo 3). A non-collinear triple (s0,s1,s2)(s_{0},s_{1},s_{2}) is also called a biased non-collinear triple if (si,sj)(s_{i},s_{j}) is a biased pair for every iji\neq j. We now state an additional NP-hardness condition of 0\overrightarrow{\textbf{0}}-Ext[μ][\mu]:

Theorem 1.5.

Let μ\mu be a rational-valued directed metric on TT. If there exists a biased non-collinear triple for μ\mu, then 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] is strongly NP-hard.

\begin{overpic}[width=398.33858pt]{biased.eps} \put(10.0,8.0){1} \put(14.5,4.0){2} \put(20.0,3.7){1} \put(23.5,9.0){2} \put(14.0,15.5){2} \put(20.5,15.5){1} \put(15.6,11.3){$r$} \put(5.5,1.5){$x$} \put(17.0,24.5){$y$} \put(28.8,1.5){$z$} \put(42.0,8.0){1} \put(47.0,4.0){2} \put(56.0,8.5){2} \put(52.0,3.5){4} \put(43.0,13.0){1} \put(46.7,18.5){3} \put(56.5,13.7){2} \put(51.7,18.0){6} \put(47.6,11.0){$r$} \put(37.5,1.5){$x$} \put(38.0,22.0){$y$} \put(60.8,21.5){$z$} \put(60.8,1.5){$w$} \put(76.0,8.0){1} \put(80.7,4.0){2} \put(90.2,8.5){2} \put(86.1,3.5){4} \put(75.5,11.5){1} \put(77.0,17.0){1} \put(80.5,17.0){3} \put(86.0,17.0){3} \put(92.0,11.8){2} \put(89.0,16.8){2} \end{overpic}
Figure 1: (a) NP-hard case        (b) tractable case            (c) tractable case     

Figure 1 (a) is an example of a directed metric satisfying the condition in Theorem 1.5, as described below. Suppose that μ\mu is the directed metric such that the underlying graph HμH_{\mu} of μ\mu is the undirected graph described in Figure 1 (a). Then, we have μ(v,r)=1,μ(r,v)=2\mu(v,r)=1,\ \mu(r,v)=2 for every v{x,y,z}v\in\{x,y,z\}. Since Rμ(x,r)=1/2R_{\mu}(x,r)=1/2 and Rμ(x,y)=3/3=1R_{\mu}(x,y)=3/3=1, (x,y)(x,y) is a biased pair. Similarly, (y,z)(y,z) and (z,x)(z,x) are biased pairs. Hence, (x,y,z)(x,y,z) is a biased non-collinear triple.

Our fourth contribution says that the non-existence of a biased non-collinear triple implies tractability, provided the underlying graph is a star.

Theorem 1.6.

Let μ\mu be a rational-valued directed metric on TT. If HμH_{\mu} is a star and there exists no biased non-collinear triple for μ\mu, then 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] is solvable in polynomial time.

Figure 1 (b) is an example of a directed metric satisfying the condition in Theorem 1.6, as (x,w)(x,w) and (y,z)(y,z) are not biased pairs in the directed metric of Figure 1 (b) (the directed metric of Figure 1 (b) is defined in the same way as that of Figure 1 (a)). The directed metric of Figure 1 (c) also satisfies the condition in Theorem 1.6.

The organization of this paper is as follows. Section 2 provides preliminary arguments which are necessary for the proofs. Section 3 introduces some notions and shows several properties in directed metric spaces. Section 4 proves the tractability results of 0\overrightarrow{\textbf{0}}-Ext[μ][\mu]. To show Theorem 1.4 and 1.6, we utilize the tractability condition of valued CSPs by Kolmogorov, Thapper, and Živný [16], as was utilized in [10] to prove Theorem 1.2. Section 5 shows the hardness results of 0\overrightarrow{\textbf{0}}-Ext[μ][\mu]. We prove Theorem 1.3 and 1.5 by use of the MC condition in [6], which is similar to the hardness proof of the multiterminal cut problem [7], and the proof of Theorem 1.1 in [13, 14].

A preliminary version [11] of this paper has appeared in Proceedings of the 45th International Symposium on Mathematical Foundations of Computer Science (MFCS 2020).

Notation.

Let ,,\mathbb{R},\ \mathbb{Q}, \mathbb{Z}, and \mathbb{N} denote the sets of reals, rationals, integers, and positive integers, respectively. We also denote the sets of nonnegative reals and rationals by +\mathbb{R}_{+} and +\mathbb{Q}_{+}.

2 Preliminaries

2.1 Valued CSP

Let DD be a finite set. For a positive integer kk, a function f:Dkf:D^{k}\rightarrow\mathbb{Q} is called a kk-ary cost function on DD. Let ar(f):=kar(f):=k denote the arity of ff. The valued constraint satisfaction problem (VCSP) on DD is defined as follows [21]:

VCSP:Instance:\displaystyle\textbf{VCSP:}\mathrm{\ \ \ \ }\mathrm{Instance}: asetV={x1,x2,,xn}ofvariables,\displaystyle\mathrm{\ \ }\mathrm{a\ set\ }V=\{x_{1},x_{2},\ldots,x_{n}\}\mathrm{\ of\ variables},
rationalvaluedcostfunctionsf1,f2,,fqonD,\displaystyle\mathrm{\ \ rational\mathchar 45\relax valued\ cost\ functions\ }f_{1},f_{2},\ldots,f_{q}\ \mathrm{on\ }D,\mathrm{\ }
weightsw1,w2,,wq+ofcostfunctions,\displaystyle\mathrm{\ \ weights\ }w_{1},w_{2},\ldots,w_{q}\in\mathbb{Q}_{+}\ \mathrm{of\ cost\ functions},
assignmentsσi:{1,2,,ar(fi)}{1,2,,n}(i=1,2,,q)\displaystyle\mathrm{\ \ assignments\ }\sigma_{i}:\{1,2,\ldots,ar(f_{i})\}\rightarrow\{1,2,\ldots,n\}\ (i=1,2,\ldots,q)
Min.\displaystyle\mathrm{Min.} i=1qwifi(xσi(1),xσi(2),,xσi(ar(fi)))\displaystyle\ \displaystyle\sum_{i=1}^{q}w_{i}\cdot f_{i}\left(x_{\sigma_{i}(1)},x_{\sigma_{i}(2)},\ldots,x_{\sigma_{i}(ar(f_{i}))}\right)
s.t.\displaystyle\mathrm{s.t.} (x1,x2,,xn)Dn.\displaystyle\mathrm{\ \ }(x_{1},x_{2},\ldots,x_{n})\in D^{n}.

In an instance of VCSP, cost functions are extensionally represented; we are given all possible values of all cost functions. Hence, the size of an instance is exponential in the arity of each cost function.

A set of cost functions on DD is called a constraint language on DD. Unless specifically said otherwise, we assume that all constraint languages are finite. Let Γ\Gamma be a constraint language on DD. The instance of VCSP is called a Γ\Gamma-instance if all cost functions in the instance belong to Γ\Gamma. Let VCSP[Γ][\Gamma] denote the class of the optimization problems whose instances are restricted to Γ\Gamma-instances.

Let μ\mu be a directed metric on TT. The directed minimum 0-extension problem 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] is viewed as a subclass of VCSP. Indeed, let

D\displaystyle D :=T,\displaystyle:=T,
f(x,y)\displaystyle f(x,y) :=μ(x,y),\displaystyle:=\mu(x,y),
gt(x)\displaystyle g_{t}(x) :=μ(x,t),\displaystyle:=\mu(x,t),
ht(x)\displaystyle h_{t}(x) :=μ(t,x),\displaystyle:=\mu(t,x),

and let

Γ:={f}{gttT}{httT}.\displaystyle\Gamma:=\{f\}\cup\{g_{t}\mid t\in T\}\cup\{h_{t}\mid t\in T\}. (2.1)

Then we can conclude that 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] is exactly VCSP[Γ][\Gamma] on DD.

Kolmogorov, Thapper, and Živný [16] discovered a criterion for a constraint language Γ\Gamma such that VCSP[Γ][\Gamma] is tractable. To describe this, we need some definitions. A function φ:DmD\varphi:D^{m}\rightarrow D is called an mm-ary operation on DD. We denote by 𝒪D(m)\mathcal{O}_{D}^{(m)} the set of mm-ary operations on DD. A vector ω:𝒪D(m)+\omega:\mathcal{O}^{(m)}_{D}\rightarrow\mathbb{R}_{+} is called a fractional operation of arity mm on DD if

φ𝒪D(m)ω(φ)=1\displaystyle\sum_{\varphi\in\mathcal{O}^{(m)}_{D}}\omega(\varphi)=1 (2.2)

is satisfied. We denote the support of ω\omega by

supp(ω):={φ𝒪D(m)ω(φ)>0}.\displaystyle\mathrm{supp}(\omega):=\{\varphi\in\mathcal{O}^{(m)}_{D}\mid\omega(\varphi)>0\}. (2.3)

Let Γ\Gamma be a constraint language on DD. A fractional operation ω:𝒪D(m)+\omega:\mathcal{O}^{(m)}_{D}\rightarrow\mathbb{R}_{+} is called a fractional polymorphism of Γ\Gamma if for every cost function fΓ(f:Dn)f\in\Gamma\ (f:D^{n}\rightarrow\mathbb{Q}) and vectors x1,,xmDnx^{1},\ldots,x^{m}\in D^{n},

φ𝒪D(m)ω(φ)f(φ(x1,,xm))1mi=1mf(xi),\displaystyle\sum_{\varphi\in\mathcal{O}^{(m)}_{D}}\omega(\varphi)f(\varphi(x^{1},\ldots,x^{m}))\leq\frac{1}{m}\sum_{i=1}^{m}f(x^{i}), (2.4)

where φ(x1,,xm)\varphi(x^{1},\ldots,x^{m}) indicates

φ(x1,,xm):=(φ(x11,,x1m)φ(xn1,,xnm)).\displaystyle\varphi(x^{1},\ldots,x^{m}):=\left(\begin{array}[]{c}\varphi(x_{1}^{1},\ldots,x_{1}^{m})\\ \vdots\\ \varphi(x_{n}^{1},\ldots,x_{n}^{m})\end{array}\right). (2.8)

We now state a powerful theorem on the relation between fractional polymorphisms and tractability of VCSP[Γ][\Gamma] by Kolmogorov, Thapper, and Živný [16]:

Theorem 2.1 ([16]).

Let Γ\Gamma be a constraint language. If Γ\Gamma admits a fractional polymorphism with a semilattice operation in its support, then VCSP[Γ][\Gamma] can be solved in polynomial time.

Here a semilattice operation φ\varphi on DD is a binary operation which satisfies φ(x,x)=x,φ(x,y)=φ(y,x)\varphi(x,x)=x,\ \varphi(x,y)=\varphi(y,x), and φ(x,φ(y,z))=φ(φ(x,y),z)\varphi(x,\varphi(y,z))=\varphi(\varphi(x,y),z) for every x,y,zDx,y,z\in D.

Remark 2.2.

Thapper and Živný [18] showed that the tractability of finite-valued CSP VCSP[Γ][\Gamma] is completely characterized by the existence of a certain fractional polymorphism for Γ\Gamma. The existence of this fractional polymorphism reduces to the feasibility of a linear program over the space of binary fractional polymorphisms. Via a version of Farkas’ lemma, the infeasibility of the LP leads to instances that can solve maximum cut problems. They showed that by using the ellipsoid method (i.e., the machinery of separation-optimization equivalence), the feasibility of this LP can be checked in polynomial time provided the constraint language is a ‘core’. Here a constraint language Γ\Gamma is called a ‘core’ if every unary fractional polymorphism ω\omega of Γ\Gamma consists of injective operations in its support.

Our constraint language Γ\Gamma defined in (2.1) is a ‘core’. Indeed, Γ\Gamma contains gtg_{t} and hth_{t}, and hence every operation φ\varphi in a unary fractional polymorphism must be an identity map (since gt(t)+ht(t)gt(φ(t))+ht(φ(t))g_{t}(t)+h_{t}(t)\geq g_{t}(\varphi(t))+h_{t}(\varphi(t)) implies φ(t)=t\varphi(t)=t).

Cohen et al. [6] discovered a sufficient condition for a constraint language Γ\Gamma such that VCSP[Γ][\Gamma] is NP-hard. We describe this condition with some definitions. For a constraint language Γ\Gamma, let Γ\langle\Gamma\rangle denote the set of all functions f(x1,,xm)f(x_{1},\ldots,x_{m}) such that for some instance II of VCSP[Γ][\Gamma] with objective function fI(x1,,xm,xm+1,,xn)f_{I}(x_{1},\ldots,x_{m},x_{m+1},\ldots,x_{n}), we have

f(x1,,xm)=minxm+1,,xnfI(x1,,xm,xm+1,,xn).\displaystyle f(x_{1},\ldots,x_{m})=\min_{x_{m+1},\ldots,x_{n}}f_{I}(x_{1},\ldots,x_{m},x_{m+1},\ldots,x_{n}). (2.9)

For a constraint language Γ\Gamma on DD, we define the following condition (MC):
(MC) There exist distinct a,bDa,b\in D such that Γ\langle\Gamma\rangle contains a binary cost function ff satisfying argminf={(a,b),(b,a)}\mathrm{argmin}\ f=\{(a,b),(b,a)\}.
Cohen et al. [6] revealed the relation between the condition (MC) and NP-hardness of VCSP[Γ][\Gamma] (also see [19]).

Proposition 2.3 ([6, Proposition 5.1]; see [19, Lemma 2]).

If a constraint language Γ\Gamma on DD satisfies the condition (MC), then VCSP[Γ][\Gamma] is strongly NP-hard.

Let Γ\Gamma be a constraint language on DD satisfying (MC). In [6], only the NP-hardness of VCSP[Γ][\Gamma] was shown, whereas the “strongly” NP-hardness was not explicitly shown. However, the strongly NP-hardness easily follows from the proofs of Proposition 5.1 and Theorem 3.4 in [6].

Remark 2.4.

Let Γ\Gamma be a constraint language satisfying (MC). The proof of NP-hardness of VCSP[Γ][\Gamma] in [6] was based on the reduction from the maximum cut problem (MAX CUT). This reduction is an extension of the hardness proofs of the multiterminal cut problem [7] and 0-Ext[μ][\mu] [13, 14] (Theorem 1.1). We prove NP-hardness of 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] with the aid of NP-hardness of VCSP[Γ][\Gamma] in Section 5. Thapper and Živný [19] also showed the complete characterization of constraint languages for which finite-valued CSPs are NP-hard, see [19] for details.

2.2 Modular graphs

Let G=(V,E)G=(V,E) be a connected undirected graph. The graph metric dG:V×Vd_{G}:V\times V\rightarrow\mathbb{Z} is defined as follows:

dG(x,y):=thenumberofedgesinashortestpathfromxtoyinG(x,yV).\displaystyle d_{G}(x,y):=\mathrm{the\ number\ of\ edges\ in\ a\ shortest\ path\ from}\ x\ \mathrm{to}\ y\ \mathrm{in}\ G\ (x,y\in V). (2.10)

We denote dGd_{G} simply by dd if GG is clear in the context. We say that GG is modular if its graph metric dGd_{G} is modular (remember that a metric μ\mu on TT is called modular if for every s0,s1,s2Ts_{0},s_{1},s_{2}\in T, there exists a median mTm\in T, which satisfies μ(si,sj)=μ(si,m)+μ(m,sj)\mu(s_{i},s_{j})=\mu(s_{i},m)+\mu(m,s_{j}) for every 0i<j20\leq i<j\leq 2). We show examples of a modular graph and a nonmodular graph in Figure 2.

\begin{overpic}[width=483.69684pt]{mfigur3.eps} \put(12.0,15.0){$p$} \put(11.5,1.2){$p_{1}$} \put(18.3,22.0){$p_{2}$} \put(37.8,7.3){$q$} \put(63.6,15.0){$p$} \put(63.1,1.2){$p_{1}$} \put(69.9,22.0){$p_{2}$} \put(89.4,7.3){$q$} \put(19.2,7.6){$p^{\prime}$} \end{overpic}
Figure 2: (a) a modular graph                                (b) a nonmodular graph
Lemma 2.5 ([2]).

A connected undirected graph G=(V,E)G=(V,E) is modular if and only if the following two conditions hold:

  1. (i)

    GG is a bipartite graph.

  2. (ii)

    For every pair of vertices p,qVp,q\in V and neighbors p1,p2p_{1},p_{2} of pp with d(p,q)=1+d(p1,q)=1+d(p2,q)d(p,q)=1+d(p_{1},q)=1+d(p_{2},q), there exists a common neighbor pp^{\prime} of p1,p2p_{1},p_{2} with d(p,q)=2+d(p,q)d(p,q)=2+d(p^{\prime},q).

By Lemma 2.5, we can easily check (non)modularity of the graphs in Figure 2. The graph in Figure 2 (b) satisfies d(p,q)=1+d(p1,q)=1+d(p2,q)d(p,q)=1+d(p_{1},q)=1+d(p_{2},q). However, there exists no common neighbor pp^{\prime} of p1,p2p_{1},p_{2} with d(p,q)=2+d(p,q)d(p,q)=2+d(p^{\prime},q), which implies that this graph is not modular. On the other hand, the graph in Figure 2 (a) has common neighbor pp^{\prime} of p1,p2p_{1},p_{2} with d(p,q)=2+d(p,q)d(p,q)=2+d(p^{\prime},q). By similar arguments, we can easily verify that the graph in Figure 2 (a) satisfies the condition (ii) of Lemma 2.5. It is also easy to verify that this graph satisfies the condition (i) of Lemma 2.5. Hence, the graph in Figure 2 (a) is modular.

Let (T,μ)(T,\mu) be a metric space. For x,yTx,y\in T, we denote the interval of x,yx,y by Iμ(x,y):={zTμ(x,y)=μ(x,z)+μ(z,y)}I_{\mu}(x,y):=\{z\in T\mid\mu(x,y)=\mu(x,z)+\mu(z,y)\}. We denote I:=IμI:=I_{\mu} if μ\mu is clear in the context. A subset XTX\subseteq T is called a convex set if I(p,q)XI(p,q)\subseteq X for every p,qXp,q\in X. A subset XTX\subseteq T is called a gated set if for every pTp\in T, there exists pXp^{\prime}\in X, called the gate of pp at XX, such that μ(p,q)=μ(p,p)+μ(p,q)\mu(p,q)=\mu(p,p^{\prime})+\mu(p^{\prime},q) for every qXq\in X. The gate of pp at XX is unique. Chepoi [4] showed the following relation between convex sets and gated sets:

Lemma 2.6 ([4]).

Let G=(V,E)G=(V,E) be a modular graph. For the metric space (V,d)(V,d) and a subset XVX\subseteq V, the following conditions are equivalent:

  1. (i)

    XX is convex.

  2. (ii)

    XX is gated.

2.3 Modular lattices

Let \mathcal{L} be a partially ordered finite set with a partial order \preceq. By aba\prec b we mean aba\preceq b and aba\neq b. For a,ba,b\in\mathcal{L}, we denote by aba\vee b the minimum element of the set {cca\{c\in\mathcal{L}\mid c\succeq a and cb}c\succeq b\}, and denote by aba\wedge b the maximum element of the set {cca\{c\in\mathcal{L}\mid c\preceq a and cb}c\preceq b\}. If for every a,ba,b\in\mathcal{L} there exist aba\vee b and aba\wedge b, then \mathcal{L} is called a lattice. A lattice \mathcal{L} is called modular if for every a,b,ca,b,c\in\mathcal{L} with aca\preceq c it holds that a(bc)=(ab)ca\vee(b\wedge c)=(a\vee b)\wedge c. For aba\preceq b\in\mathcal{L}, we let [a,b][a,b] denote the interval {cacb}\{c\in\mathcal{L}\mid a\preceq c\preceq b\}. For aba\prec b\in\mathcal{L}, a sequence (a=u0,u1,,un=b)(a=u_{0},u_{1},\ldots,u_{n}=b) is called a chain from aa to bb if ui1uiu_{i-1}\prec u_{i} holds for all i{1,2,,n}i\in\{1,2,\ldots,n\}. Here the length of a chain (u0,u1,,un)(u_{0},u_{1},\ldots,u_{n}) is nn. We denote by r[a,b]r[a,b] the length of the longest chain from aa to bb. For a lattice \mathcal{L}, let 0 denote the minimum element of \mathcal{L}, and let 1 denote the maximum element of \mathcal{L}. The rank r(a)r(a) of an element aa is defined by r(a):=r[0,a]r(a):=r[\textbf{0},a].

Lemma 2.7 (see [3, Chapter II]).

Let \mathcal{L} be a modular lattice. For aba\preceq b\in\mathcal{L}, the following condition (called Jordan–Dedekind chain condition) holds:

All maximal chains from atobhave the same length.\displaystyle\mathrm{\textit{All\ maximal\ chains\ from\ }}a\ \mathrm{\textit{to}}\ b\ \mathrm{\textit{have\ the\ same\ length.}} (2.11)

By Lemma 2.7, we can see that for a modular lattice \mathcal{L} and aa\in\mathcal{L}, r(a)r(a) is equal to the length of a maximal chain from 0 to aa. A modular lattice is also characterized by rank as follows:

Lemma 2.8 (see [3, Chapter II]).

A lattice \mathcal{L} is modular if and only if for every a,ba,b\in\mathcal{L}, r(a)+r(b)=r(ab)+r(ab)r(a)+r(b)=r(a\wedge b)+r(a\vee b) holds.

For a poset \mathcal{L} and a,ba,b\in\mathcal{L}, we say that bb covers aa if aba\prec b holds and there is no cc\in\mathcal{L} with acba\prec c\prec b. The covering graph of \mathcal{L} is the undirected graph obtained by linking all pairs a,ba,b of \mathcal{L} such that aa covers bb, or bb covers aa. Here we have the following relation between modular lattices and modular graphs:

Lemma 2.9 ([20]).

A lattice \mathcal{L} is modular if and only if the covering graph of \mathcal{L} is modular.

Let \mathcal{L} be a lattice. A function f:f:\mathcal{L}\rightarrow\mathbb{R} is called submodular if f(p)+f(q)f(pq)+f(pq)f(p)+f(q)\geq f(p\vee q)+f(p\wedge q) holds for every p,qp,q\in\mathcal{L}. If a,ba,b\in\mathcal{L} are covered by aba\vee b, then the pair (a,b)(a,b) is called a 2-covered pair. We have the following characterization of submodular functions on modular lattices:

Lemma 2.10.

Let \mathcal{L} be a modular lattice. A function f:f:\mathcal{L}\rightarrow\mathbb{R} is submodular if and only if f(a)+f(b)f(ab)+f(ab)f(a)+f(b)\geq f(a\vee b)+f(a\wedge b) holds for every 2-covered pair (a,b)(a,b).

Proof.

The only if part is obvious. We prove the if part. Take p,qp,q\in\mathcal{L} and maximal chains (pq=p0,p1,,pk=p)(p\wedge q=p_{0},p_{1},\ldots,p_{k}=p) and (pq=q0,q1,,ql=q)(p\wedge q=q_{0},q_{1},\ldots,q_{l}=q). First we show that pi+1qjp_{i+1}\vee q_{j} covers piqjp_{i}\vee q_{j}. Note that pi+1p_{i+1} covers pip_{i}. Since \mathcal{L} is modular, we have pi+1(piqj)=pi(qjpi+1)=pip_{i+1}\wedge(p_{i}\vee q_{j})=p_{i}\vee(q_{j}\wedge p_{i+1})=p_{i} andpi+1(piqj)=pi+1qj\ p_{i+1}\vee(p_{i}\vee q_{j})=p_{i+1}\vee q_{j}. Hence, we can conclude that pi+1qjp_{i+1}\vee q_{j} covers piqjp_{i}\vee q_{j} by Lemma 2.4. Similarly, piqj+1p_{i}\vee q_{j+1} covers piqjp_{i}\vee q_{j}. Also, by modularity we have (pi+1qj)(piqj+1)=pi+1qj+1(p_{i+1}\vee q_{j})\vee(p_{i}\vee q_{j+1})=p_{i+1}\vee q_{j+1} and (pi+1qj)(piqj+1)=pi(qj+1(pi+1qj))=piqj(pi+1qj+1)=piqj(p_{i+1}\vee q_{j})\wedge(p_{i}\vee q_{j+1})=p_{i}\vee(q_{j+1}\wedge(p_{i+1}\vee q_{j}))=p_{i}\vee q_{j}\vee(p_{i+1}\wedge q_{j+1})=p_{i}\vee q_{j}. Then we conclude that (pi+1qj,piqj+1)(p_{i+1}\vee q_{j},p_{i}\vee q_{j+1}) is a 2-covered pair. Let ai,j:=piqja_{i,j}:=p_{i}\vee q_{j}. Then we have f(p)+f(q)f(pq)f(pq)=i,j(f(ai+1,j)+f(ai,j+1)f(ai+1,j+1)f(ai,j))0f(p)+f(q)-f(p\vee q)-f(p\wedge q)=\sum_{i,j}(f(a_{i+1,j})+f(a_{i,j+1})-f(a_{i+1,j+1})-f(a_{i,j}))\geq 0. Thus, we conclude that ff is submodular. ∎

3 Directed metric spaces

3.1 Modular directed metrics

We first extend the notions of modularity, medians, and underlying graphs to directed metric spaces. Let μ\mu be a directed metric on TT. We say that μ\mu is modular if and only if for every s0,s1,s2Ts_{0},s_{1},s_{2}\in T, there exists an element mTm\in T, called a median, such that μ(si,sj)=μ(si,m)+μ(m,sj)\mu(s_{i},s_{j})=\mu(s_{i},m)+\mu(m,s_{j}) for every 0i,j2(ij)0\leq i,j\leq 2\ (i\neq j). See Figure 3 (a) for an example of modular directed metrics. We define the underlying graph of μ\mu as the undirected graph Hμ=(T,U)H_{\mu}=(T,U), where

U:={{x,y}x,yT(xy),zT{x,y},μ(x,y)<μ(x,z)+μ(z,y)\displaystyle U:=\{\{x,y\}\mid x,y\in T\ (x\neq y),\ \forall z\in T\setminus\{x,y\},\ \mu(x,y)<\mu(x,z)+\mu(z,y)
orzT{x,y},μ(y,x)<μ(y,z)+μ(z,x)}.\displaystyle\mathrm{or}\ \forall z\in T\setminus\{x,y\},\ \mu(y,x)<\mu(y,z)+\mu(z,x)\}. (3.1)
\begin{overpic}[width=398.33858pt]{mfigure4.eps} \put(7.2,5.0){$p$} \put(25.5,11.5){$q$} \put(43.0,5.0){$r$} \put(25.5,31.0){$s$} \put(55.5,5.0){$p$} \put(74.0,11.5){$q$} \put(91.0,5.0){$r$} \put(74.0,31.0){$s$} \put(25.2,1.7){2} \put(25.2,6.8){3} \put(19.5,8.0){1} \put(17.5,12.0){2} \put(31.4,8.0){1} \put(33.0,11.9){1} \put(23.8,20.0){1} \put(27.3,20.0){4} \put(16.0,20.0){0} \put(19.0,16.0){2} \put(31.2,16.2){0} \put(34.5,20.0){3} \end{overpic}
Figure 3: (a) a modular directed metric            (b) the underlying graph of (a)

For a directed metric μ\mu on TT and v0,v1,,vnTv_{0},v_{1},\ldots,v_{n}\in T, we say that a sequence (v0,v1,,vn)(v_{0},v_{1},\ldots,v_{n}) is μ\mu-shortest if μ(v0,vn)=i=0n1μ(vi,vi+1)\mu(v_{0},v_{n})=\sum_{i=0}^{n-1}\mu(v_{i},v_{i+1}). Bandelt [1] showed that for a modular (undirected) metric μ\mu, a μ\mu-shortest sequence is also dHμd_{H_{\mu}}-shortest. We have the following directed version of this property:

Lemma 3.1.

Let μ\mu be a modular directed metric on TT, and let v0,v1,,vnTv_{0},v_{1},\ldots,v_{n}\in T.

  1. (1)

    If a sequence (v0,v1,,vn)(v_{0},v_{1},\ldots,v_{n}) is μ\mu-shortest, then the inverted sequence (vn,vn1,,v0)(v_{n},v_{n-1},\ldots,v_{0}) is also μ\mu-shortest.

  2. (2)

    If a sequence (v0,v1,,vn)(v_{0},v_{1},\ldots,v_{n}) is μ\mu-shortest, then this sequence is also dHμd_{H_{\mu}}-shortest.

Proof.

In this proof, we denote I:=IμI:=I_{\mu}. We first show (1) by induction on nn. Suppose that (v0,v1,,vn)(v_{0},v_{1},\ldots,v_{n}) is μ\mu-shortest. If n=1n=1 then (v1,v0)(v_{1},v_{0}) is μ\mu-shortest and (1) holds. Suppose that n2n\geq 2. Since μ\mu is modular, there exists a median mTm\in T of v0,v1,vnv_{0},v_{1},v_{n}. There exists no element zI(v0,v1)I(v1,vn){v1}z\in I(v_{0},v_{1})\cap I(v_{1},v_{n})\setminus\{v_{1}\}, since otherwise a sequence (v0,z,v1,z,vn)(v_{0},z,v_{1},z,v_{n}) is μ\mu-shortest, which is impossible by μ(z,v1)+μ(v1,z)>0\mu(z,v_{1})+\mu(v_{1},z)>0. Hence, we have m=v1m=v_{1}. Then a sequence (vn,v1,v0)(v_{n},v_{1},v_{0}) is μ\mu-shortest. Therefore, it suffices to show that a sequence (vn,vn1,,v1)(v_{n},v_{n-1},\ldots,v_{1}) is μ\mu-shortest, which is implied by induction. Hence, we can conclude that (vn,vn1,,v0)(v_{n},v_{n-1},\ldots,v_{0}) is μ\mu-shortest.

Next we show (2). We denote d:=dHμd:=d_{H_{\mu}} for simplicity. Suppose that (v0,v1,,vn)(v_{0},v_{1},\ldots,v_{n}) is μ\mu-shortest. Without loss of generality, we may assume that vivjv_{i}\neq v_{j} for any iji\neq j. If n=1n=1, then (v0,v1)(v_{0},v_{1}) is obviously dd-shortest. Suppose that n2n\geq 2. If d(v0,vn)=1d(v_{0},v_{n})=1, then HμH_{\mu} has the edge {v0,vn}\{v_{0},v_{n}\}. However, this contradicts μ(v0,vn)=μ(v0,v1)+μ(v1,vn)\mu(v_{0},v_{n})=\mu(v_{0},v_{1})+\mu(v_{1},v_{n}) and the definition of HμH_{\mu} because of (1). Hence, it suffices to consider the case when d(v0,vn)2d(v_{0},v_{n})\geq 2. In the case of n=2n=2 with I(v0,v1)={v0,v1}I(v_{0},v_{1})=\{v_{0},v_{1}\} and I(v1,v2)={v1,v2}I(v_{1},v_{2})=\{v_{1},v_{2}\}, the underlying graph HμH_{\mu} has the edges {v0,v1}\{v_{0},v_{1}\} and {v1,v2}\{v_{1},v_{2}\}. Then the edge v0v2v_{0}v_{2} is not contained in HμH_{\mu}, and hence (v0,v1,v2)(v_{0},v_{1},v_{2}) is dd-shortest. In the case of n=2n=2 with I(v0,v1){v0,v1}I(v_{0},v_{1})\neq\{v_{0},v_{1}\}, there exists an element zI(v0,v1){v0,v1}z\in I(v_{0},v_{1})\setminus\{v_{0},v_{1}\}. Then the case of (v0,v1,v2)(v_{0},v_{1},v_{2}) is reduced to the case of (v0,z,v1,v2)(v_{0},z,v_{1},v_{2}). Similarly, the case of n=2n=2 with I(v1,v2){v1,v2}I(v_{1},v_{2})\neq\{v_{1},v_{2}\} is reduced to the case of n=3n=3.

Thus, it suffices to consider the case of n3n\geq 3. We show by induction on d(v0,vn)d(v_{0},v_{n}) that (v0,v1,,vn)(v_{0},v_{1},\ldots,v_{n}) is dd-shortest. Let (u0,u1,,ul)(u_{0},u_{1},\ldots,u_{l}) be a shortest path in HμH_{\mu} from v0v_{0} to vnv_{n} (u0=v0,ul=vnu_{0}=v_{0},u_{l}=v_{n}). Since μ\mu is modular, there exists a median mTm\in T of u0,ul1,vn1u_{0},u_{l-1},v_{n-1}. Then we have mI(u0,ul1)m\in I(u_{0},u_{l-1}). We may assume that (u0,m,ul1)(u_{0},m,u_{l-1}) is dd-shortest by induction hypothesis. Then (u0,m,ul)(u_{0},m,u_{l}) is also dd-shortest. Since (u0,m,vn1)(u_{0},m,v_{n-1}) is μ\mu-shortest, (m,vn1,vn)(m,v_{n-1},v_{n}) is also μ\mu-shortest. Suppose that mu0m\neq u_{0}. In this case, (m,vn1,vn)(m,v_{n-1},v_{n}) is dd-shortest by induction hypothesis. This implies that (u0,m,vn1,vn)(u_{0},m,v_{n-1},v_{n}) is dd-shortest, because (u0,m,ul)(u_{0},m,u_{l}) is dd-shortest. We may also assume that (v0,v1,,vn1)(v_{0},v_{1},\ldots,v_{n-1}) is dd-shortest by induction hypothesis. Hence, we conclude that (v0,v1,,vn)(v_{0},v_{1},\ldots,v_{n}) is dd-shortest.

Consider the case when m=u0m=u_{0}. In this case, (ul1,u0,vn1)(u_{l-1},u_{0},v_{n-1}) is μ\mu-shortest. Then we have μ(ul1,vn1)=μ(ul1,v1)+μ(v1,vn1)\mu(u_{l-1},v_{n-1})=\mu(u_{l-1},v_{1})+\mu(v_{1},v_{n-1}). We next apply the similar argument to ul,ul1,v1u_{l},u_{l-1},v_{1}, which we apply to u0,ul1,vn1u_{0},u_{l-1},v_{n-1} above. Since μ\mu is modular, there exists a median mTm^{\prime}\in T of ul,ul1,v1u_{l},u_{l-1},v_{1}. Then we may assume that (ul1,m,ul)(u_{l-1},m^{\prime},u_{l}) is dd-shortest by induction hypothesis. Hence, (u0,m,ul)(u_{0},m^{\prime},u_{l}) is also dd-shortest. Since (v1,m,ul)(v_{1},m^{\prime},u_{l}) is μ\mu-shortest, (v0,v1,m)(v_{0},v_{1},m^{\prime}) is also μ\mu-shortest. Suppose that mulm^{\prime}\neq u_{l}. In this case, (v0,v1,m)(v_{0},v_{1},m^{\prime}) is dd-shortest by induction hypothesis, which implies that (v0,v1,m,ul)(v_{0},v_{1},m^{\prime},u_{l}) is dd-shortest. We may also assume that (v1,v2,,vn)(v_{1},v_{2},\ldots,v_{n}) is dd-shortest by induction hypothesis. Hence, we conclude that (v0,v1,,vn)(v_{0},v_{1},\ldots,v_{n}) is dd-shortest. Thus, it suffices to consider the case when m=ulm^{\prime}=u_{l}. In this case, since (ul1,ul,v1)(u_{l-1},u_{l},v_{1}) is μ\mu-shortest, we have μ(ul1,v1)=μ(ul1,vn1)+μ(vn1,v1)\mu(u_{l-1},v_{1})=\mu(u_{l-1},v_{n-1})+\mu(v_{n-1},v_{1}). Thus, we have μ(v1,vn1)+μ(vn1,v1)=0\mu(v_{1},v_{n-1})+\mu(v_{n-1},v_{1})=0 (recall that μ(ul1,vn1)=μ(ul1,v1)+μ(v1,vn1)\mu(u_{l-1},v_{n-1})=\mu(u_{l-1},v_{1})+\mu(v_{1},v_{n-1})). This is a contradiction. ∎

For a modular directed metric μ\mu on TT, let mm be a median of x,y,zTx,y,z\in T in μ\mu. Then, by Lemma 3.1 mm is also a median of x,y,zx,y,z in HμH_{\mu}. Hence, we have the following lemma:

Lemma 3.2.

If a directed metric μ\mu is modular, then HμH_{\mu} is also modular.

3.2 Directed orbits and directed orbit invariance

Let G=(V,E)G=(V,E) be an undirected graph. Let E:={(u,v),(v,u){u,v}E}V×V\overleftrightarrow{E}:=\{(u,v),\ (v,u)\mid\{u,v\}\in E\}\subseteq V\times V, and G:=(V,E)\overleftrightarrow{G}:=(V,\overleftrightarrow{E}). An element of E\overleftrightarrow{E} is called an oriented edge of EE. For a path PP from ss to tt in GG, we orient each edge of PP along the direction of PP, and we denote by P\overrightarrow{P} the corresponding path in G\overleftrightarrow{G}. Let P\overrightarrow{P} and W\overrightarrow{W} be paths in G\overleftrightarrow{G} such that the end point of P\overrightarrow{P} and the start point of W\overrightarrow{W} are identified. Then we denote by PW\overrightarrow{P}\cup\overrightarrow{W} the path obtained by concatenating P\overrightarrow{P} and W\overrightarrow{W} in this order. In particular, if W\overrightarrow{W} consists of one oriented edge (p,q)(p,q), then we simply denote W:=(p,q)\overrightarrow{W}:=(p,q) and PW:=P(p,q)\overrightarrow{P}\cup\overrightarrow{W}:=\overrightarrow{P}\cup(p,q). For e,eE\overrightarrow{e},\overrightarrow{e^{\prime}}\in\overleftrightarrow{E}, we say that e\overrightarrow{e} and e\overrightarrow{e^{\prime}} are projective if there exists a sequence (e=e0,e1,,em=e)(ei=(pi,qi)Eforeachi)(\overrightarrow{e}=\overrightarrow{e_{0}},\overrightarrow{e_{1}},...,\overrightarrow{e_{m}}=\overrightarrow{e^{\prime}})\ (\overrightarrow{e_{i}}=(p_{i},q_{i})\in\overleftrightarrow{E}\ \mathrm{for\ each\ }i) such that (pi,qi,qi+1,pi+1,pi)(p_{i},q_{i},q_{i+1},p_{i+1},p_{i}) is a 4-cycle in GG for each ii. An equivalence class of the projectivity relation is called a directed orbit. Then we have the following lemma about the number of oriented edges of each directed orbit included in a shortest path. This is a sharpening of the result for undirected graphs due to Bandelt [1], and is similarly shown by the proof of the undirected version.

Lemma 3.3 ([1]).

Let G=(V,E)G=(V,E) be a modular graph, and let Q\overrightarrow{Q} be a directed orbit. For x,yVx,y\in V, let PP be a path from xx to yy, and let PP^{*} be a shortest path from xx to yy. Then we have

|PQ||PQ|.\displaystyle|\overrightarrow{P^{*}}\cap\overrightarrow{Q}|\leq|\overrightarrow{P}\cap\overrightarrow{Q}|. (3.2)

Let G=(V,E)G=(V,E) be an undirected graph. If a function h:E+h:\overleftrightarrow{E}\rightarrow\mathbb{R}_{+} satisfies h(e)=h(e)h(\overrightarrow{e})=h(\overrightarrow{e^{\prime}}) for every e,eE\overrightarrow{e},\overrightarrow{e^{\prime}}\in\overleftrightarrow{E} belonging to the same directed orbit, then we say that hh is directed orbit-invariant. Let μ\mu be a directed metric on TT with the underlying graph Hμ=(T,U)H_{\mu}=(T,U). We say that μ\mu is directed orbit-invariant if μ(u1,u2)=μ(u1,u2)\mu(u_{1},u_{2})=\mu(u_{1}^{\prime},u_{2}^{\prime}) holds for every u=(u1,u2),u=(u1,u2)U\overrightarrow{u}=(u_{1},u_{2}),\overrightarrow{u^{\prime}}=(u_{1}^{\prime},u_{2}^{\prime})\in\overleftrightarrow{U} belonging to the same directed orbit in HμH_{\mu}. A 4-cycle (p,q,r,s,p)(p,q,r,s,p) in HμH_{\mu} is called a directed orbit-varying modular cycle if μ(p,q)μ(s,r)=μ(r,s)μ(q,p)=μ(p,s)μ(q,r)=μ(r,q)μ(s,p)0\mu(p,q)-\mu(s,r)=\mu(r,s)-\mu(q,p)=\mu(p,s)-\mu(q,r)=\mu(r,q)-\mu(s,p)\neq 0. The cycle (p,q,r,s,p)(p,q,r,s,p) in Figure 3 (b) is an example of a directed orbit-varying modular cycle.

Bandelt [1] showed that a metric μ\mu is orbit-invariant if μ\mu is modular. A directed metric μ\mu is not necessarily directed orbit-invariant even if μ\mu is modular. For example, if HμH_{\mu} is a directed orbit-varying modular cycle, then μ\mu is modular but not directed orbit-invariant. The name “directed orbit-varying modular cycle” is motivated by this fact. We now have the following sufficient condition of a directed metric to be directed orbit-invariant.

Lemma 3.4.

Let μ\mu be a modular directed metric. Suppose that HμH_{\mu} has no directed orbit-varying modular cycle. Then, μ\mu is directed orbit-invariant.

Proof.

It suffices to show that μ(p,q)=μ(s,r)\mu(p,q)=\mu(s,r) for any 4-cycle (p,q,r,s,p)(p,q,r,s,p) in HμH_{\mu}. Suppose to the contrary that a 4-cycle (p,q,r,s,p)(p,q,r,s,p) in HμH_{\mu} satisfies μ(p,q)μ(s,r)\mu(p,q)\neq\mu(s,r). Let k:=μ(p,q)μ(s,r)0k:=\mu(p,q)-\mu(s,r)\neq 0. Since μ\mu is modular, the triple p,q,rp,q,r has a median mm. The underlying graph HμH_{\mu} has edges {p,q}\{p,q\} and {q,r}\{q,r\}, which implies m=qm=q. Hence, the sequence (p,q,r)(p,q,r) is μ\mu-shortest. Similarly, (p,s,r)(p,s,r) is μ\mu-shortest. Then we have μ(p,q)+μ(q,r)=μ(p,s)+μ(s,r)=μ(p,r)\mu(p,q)+\mu(q,r)=\mu(p,s)+\mu(s,r)=\mu(p,r). Hence, we have μ(p,s)μ(q,r)=μ(p,q)μ(s,r)=k\mu(p,s)-\mu(q,r)=\mu(p,q)-\mu(s,r)=k. Similarly, sequences (s,p,q)(s,p,q) and (s,r,q)(s,r,q) are μ\mu-shortest, then we have μ(s,p)+μ(p,q)=μ(s,r)+μ(r,q)\mu(s,p)+\mu(p,q)=\mu(s,r)+\mu(r,q). Hence, we have μ(r,q)μ(s,p)=μ(p,q)μ(s,r)=k\mu(r,q)-\mu(s,p)=\mu(p,q)-\mu(s,r)=k. Similarly, sequences (q,p,s)(q,p,s) and (q,r,s)(q,r,s) are μ\mu-shortest, then we have μ(q,p)+μ(p,s)=μ(q,r)+μ(r,s)\mu(q,p)+\mu(p,s)=\mu(q,r)+\mu(r,s). Hence, we have μ(r,s)μ(q,p)=μ(p,s)μ(q,r)=k\mu(r,s)-\mu(q,p)=\mu(p,s)-\mu(q,r)=k. Therefore, we have μ(p,q)μ(s,r)=μ(p,s)μ(q,r)=μ(r,q)μ(s,p)=μ(r,s)μ(q,p)=k0\mu(p,q)-\mu(s,r)=\mu(p,s)-\mu(q,r)=\mu(r,q)-\mu(s,p)=\mu(r,s)-\mu(q,p)=k\neq 0, then we conclude that the cycle (p,q,r,s,p)(p,q,r,s,p) is a directed orbit-varying modular cycle. This is a contradiction. ∎

We now consider a sufficient condition for the converse of Lemma 3.1 (2) to hold. For an undirected metric μ\mu, Bandelt [1] showed that if μ\mu is orbit-invariant and HμH_{\mu} is modular, then a dHμd_{H_{\mu}}-shortest sequence is also μ\mu-shortest. The similar property also holds for a directed metric as follows:

Lemma 3.5.

Let μ\mu be a directed metric on TT, and let v0,v1,,vnTv_{0},v_{1},\ldots,v_{n}\in T. If μ\mu is directed orbit-invariant and HμH_{\mu} is modular, then the following condition holds:

If (v0,v1,,vn)isdHμ-shortest,then it is alsoμ-shortest.\displaystyle\textit{If\ }(v_{0},v_{1},\ldots,v_{n})\ \textit{is}\ d_{H_{\mu}}\textit{-shortest,}\ \textit{then\ it\ is\ also}\ \mu\textit{-shortest.} (3.3)
Proof.

Without loss of generality, we may assume that (v0,v1,,vn)(v_{0},v_{1},\ldots,v_{n}) is a shortest path from v0v_{0} to vnv_{n} in HμH_{\mu}. Also, by the definition of HμH_{\mu} we see that there is a path (v0=u0,u1,,um=vn)(v_{0}=u_{0},u_{1},\ldots,u_{m}=v_{n}) in HμH_{\mu} with μ(u0,um)=μ(u0,u1)+μ(u1,u2)++μ(um1,um)\mu(u_{0},u_{m})=\mu(u_{0},u_{1})+\mu(u_{1},u_{2})+\cdots+\mu(u_{m-1},u_{m}). Hence, by Lemma 3.3 and directed orbit-invariance of μ\mu, we have

μ(v0,v1)+μ(v1,v2)++μ(vn1,vn)\displaystyle\mu(v_{0},v_{1})+\mu(v_{1},v_{2})+\cdots+\mu(v_{n-1},v_{n}) μ(u0,u1)+μ(u1,u2)++μ(um1,um)\displaystyle\leq\mu(u_{0},u_{1})+\mu(u_{1},u_{2})+\cdots+\mu(u_{m-1},u_{m})
=μ(v0,vn).\displaystyle=\mu(v_{0},v_{n}). (3.4)

Thus, we have μ(v0,v1)+μ(v1,v2)++μ(vn1,vn)=μ(v0,vn)\mu(v_{0},v_{1})+\mu(v_{1},v_{2})+\cdots+\mu(v_{n-1},v_{n})=\mu(v_{0},v_{n}). ∎

As we obtain Lemma 3.2 from Lemma 3.1, we also obtain the following property from Lemma 3.5 by applying a similar argument:

Lemma 3.6.

Let μ\mu be a directed metric. If μ\mu is directed orbit-invariant and HμH_{\mu} is modular, then μ\mu is also modular.

4 Proof of tractability

In this section, we give proofs of Theorem 1.4 and Theorem 1.6. Let μ\mu be a directed metric on TT, and Γ\Gamma be the constraint language defined in (2.1). Then, as we see in Section 2.1, 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] is exactly VCSP[Γ][\Gamma]. Hence, by Theorem 2.1 we can prove the tractability of 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] by constructing a fractional polymorphism with a semilattice operation in its support. In the proof of Theorem 1.4, we construct a fractional polymorphism that characterizes submodularity of μ\mu. To show submodularity, we imitate the proof of submodularity of metric functions on modular semilattices in the undirected version [10]. In the proof of Theorem 1.6, we also construct fractional polymorphisms that are similar to the fractional polymorphisms characterizing submodularity.

4.1 Proof of Theorem 1.4

Note that the underlying graph HμH_{\mu} of μ\mu is the covering graph of a modular lattice \mathcal{L} with a partial order \preceq. We define a partial order \preceq on ×\mathcal{L}\times\mathcal{L} by (a,b)(c,d)ac(a,b)\preceq(c,d)\Longleftrightarrow a\preceq c and bd(a,b,c,d)b\preceq d\ (a,b,c,d\in\mathcal{L}). Then ×\mathcal{L}\times\mathcal{L} is also a modular lattice. If μ\mu is a submodular function on ×\mathcal{L}\times\mathcal{L}, then by Theorem 2.1 we can conclude that 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] is solvable in polynomial time. Hence, the following property completes the proof:

Theorem 4.1.

Let μ\mu be a directed metric. Suppose that HμH_{\mu} is the covering graph of a modular lattice \mathcal{L} and μ\mu is directed orbit-invariant. Then the function μ:×+\mu:\mathcal{L}\times\mathcal{L}\rightarrow\mathbb{R}_{+} is submodular.

Proof.

Note that ×\mathcal{L}\times\mathcal{L} is a modular lattice. By Lemma 2.10, μ\mu is a submodular function on ×\mathcal{L}\times\mathcal{L} if and only if μ(a)+μ(b)μ(ab)+μ(ab)\mu(a)+\mu(b)\geq\mu(a\vee b)+\mu(a\wedge b) holds for every 2-covered pair (a,b)(a,b×)(a,b)\ (a,b\in\mathcal{L}\times\mathcal{L}). Thus, it suffices to show that μ(a)+μ(b)μ(ab)+μ(ab)\mu(a)+\mu(b)\geq\mu(a\vee b)+\mu(a\wedge b) holds for any 2-covered pair (a,b)(a,b). Let a=(a1,a2),b=(b1,b2)(a1,a2,b1,b2)a=(a_{1},a_{2}),b=(b_{1},b_{2})\ (a_{1},a_{2},b_{1},b_{2}\in\mathcal{L}). Then, it suffices to consider the following two cases:

  1. (i)

    a1=b1a_{1}=b_{1}, and a2b2a_{2}\vee b_{2} covers a2,b2a_{2},b_{2}.

  2. (ii)

    a1a_{1} covers b1b_{1}, and b2b_{2} covers a2a_{2}.

We first consider the case (i). It suffices to show that μ(a1,a2)+μ(a1,b2)μ(a1,a2b2)+μ(a1,a2b2)\mu(a_{1},a_{2})+\mu(a_{1},b_{2})\geq\mu(a_{1},a_{2}\vee b_{2})+\mu(a_{1},a_{2}\wedge b_{2}). Let Y:=[a2b2,a2b2]Y:=[a_{2}\wedge b_{2},a_{2}\vee b_{2}]. Then, for every yY{a2b2,a2b2}y\in Y\setminus\{a_{2}\wedge b_{2},a_{2}\vee b_{2}\}, it holds that a2b2a_{2}\vee b_{2} covers yy and yy covers a2b2a_{2}\wedge b_{2}, because of Lemma 2.7 and Lemma 2.8. Hence, YY is a convex set in the metric space (T,d)(T,d) (in this proof, we denote d:=dHμd:=d_{H_{\mu}} for simplicity). Since \mathcal{L} is modular, by Lemma 2.9 HμH_{\mu} is modular. Then, by Lemma 2.6 YY is a gated set. Hence, there exists yYy^{*}\in Y such that d(a1,y)=d(a1,y)+d(y,y)d(a_{1},y)=d(a_{1},y^{*})+d(y^{*},y) holds for every yYy\in Y. Therefore, by Lemma 3.5, (a1,y,y)(a_{1},y^{*},y) is μ\mu-shortest for every yYy\in Y. If y=a2y^{*}=a_{2}, then we have

μ(a1,a2b2)\displaystyle\mu(a_{1},a_{2}\vee b_{2}) =μ(a1,a2)+μ(a2,a2b2),\displaystyle=\mu(a_{1},a_{2})+\mu(a_{2},a_{2}\vee b_{2}),
μ(a1,a2b2)\displaystyle\mu(a_{1},a_{2}\wedge b_{2}) =μ(a1,a2)+μ(a2,a2b2),\displaystyle=\mu(a_{1},a_{2})+\mu(a_{2},a_{2}\wedge b_{2}),
μ(a1,b2)\displaystyle\mu(a_{1},b_{2}) =μ(a1,a2)+μ(a2,b2).\displaystyle=\mu(a_{1},a_{2})+\mu(a_{2},b_{2}). (4.1)

Furthermore, since μ\mu is directed orbit-invariant, we have μ(a2b2,b2)=μ(a2,a2b2)\mu(a_{2}\vee b_{2},b_{2})=\mu(a_{2},a_{2}\wedge b_{2}). In addition, by Lemma 3.5 we have μ(a2,b2)=μ(a2,a2b2)+μ(a2b2,b2)\mu(a_{2},b_{2})=\mu(a_{2},a_{2}\vee b_{2})+\mu(a_{2}\vee b_{2},b_{2}). Hence, we have μ(a1,b2)=μ(a1,a2)+μ(a2,a2b2)+μ(a2,a2b2)\mu(a_{1},b_{2})=\mu(a_{1},a_{2})+\mu(a_{2},a_{2}\vee b_{2})+\mu(a_{2},a_{2}\wedge b_{2}). Therefore, we obtain μ(a1,a2)+μ(a1,b2)=μ(a1,a2b2)+μ(a1,a2b2)\mu(a_{1},a_{2})+\mu(a_{1},b_{2})=\mu(a_{1},a_{2}\vee b_{2})+\mu(a_{1},a_{2}\wedge b_{2}). Similarly, if y=b2y^{*}=b_{2}, we obtain μ(a1,a2)+μ(a1,b2)=μ(a1,a2b2)+μ(a1,a2b2)\mu(a_{1},a_{2})+\mu(a_{1},b_{2})=\mu(a_{1},a_{2}\vee b_{2})+\mu(a_{1},a_{2}\wedge b_{2}). If y=a2b2y^{*}=a_{2}\vee b_{2}, then we have

μ(a1,a2)\displaystyle\mu(a_{1},a_{2}) =μ(a1,a2b2)+μ(a2b2,a2),\displaystyle=\mu(a_{1},a_{2}\vee b_{2})+\mu(a_{2}\vee b_{2},a_{2}),
μ(a1,b2)\displaystyle\mu(a_{1},b_{2}) =μ(a1,a2b2)+μ(a2b2,b2),\displaystyle=\mu(a_{1},a_{2}\vee b_{2})+\mu(a_{2}\vee b_{2},b_{2}),
μ(a1,a2b2)\displaystyle\mu(a_{1},a_{2}\wedge b_{2}) =μ(a1,a2b2)+μ(a2b2,a2)+μ(a2,a2b2).\displaystyle=\mu(a_{1},a_{2}\vee b_{2})+\mu(a_{2}\vee b_{2},a_{2})+\mu(a_{2},a_{2}\wedge b_{2}). (4.2)

Since μ\mu is directed orbit-invariant, we have μ(a2b2,b2)=μ(a2,a2b2)\mu(a_{2}\vee b_{2},b_{2})=\mu(a_{2},a_{2}\wedge b_{2}). Hence, by (4.1) we obtain μ(a1,a2)+μ(a1,b2)=μ(a1,a2b2)+μ(a1,a2b2)\mu(a_{1},a_{2})+\mu(a_{1},b_{2})=\mu(a_{1},a_{2}\vee b_{2})+\mu(a_{1},a_{2}\wedge b_{2}). Similarly, if y=a2b2y^{*}=a_{2}\wedge b_{2}, then we obtain μ(a1,a2)+μ(a1,b2)=μ(a1,a2b2)+μ(a1,a2b2)\mu(a_{1},a_{2})+\mu(a_{1},b_{2})=\mu(a_{1},a_{2}\vee b_{2})+\mu(a_{1},a_{2}\wedge b_{2}). Thus, it suffices to consider the case when ya2,b2,a2b2,a2b2y^{*}\neq a_{2},b_{2},a_{2}\vee b_{2},a_{2}\wedge b_{2}. In this case, we have

μ(a1,a2)\displaystyle\mu(a_{1},a_{2}) =μ(a1,y)+μ(y,a2b2)+μ(a2b2,a2)μ(a1,a2b2),\displaystyle=\mu(a_{1},y^{*})+\mu(y^{*},a_{2}\vee b_{2})+\mu(a_{2}\vee b_{2},a_{2})\geq\mu(a_{1},a_{2}\vee b_{2}),
μ(a1,b2)\displaystyle\mu(a_{1},b_{2}) =μ(a1,y)+μ(y,a2b2)+μ(a2b2,b2)μ(a1,a2b2).\displaystyle=\mu(a_{1},y^{*})+\mu(y^{*},a_{2}\wedge b_{2})+\mu(a_{2}\wedge b_{2},b_{2})\geq\mu(a_{1},a_{2}\wedge b_{2}). (4.3)

Hence, we have μ(a1,a2)+μ(a1,b2)μ(a1,a2b2)+μ(a1,a2b2)\mu(a_{1},a_{2})+\mu(a_{1},b_{2})\geq\mu(a_{1},a_{2}\vee b_{2})+\mu(a_{1},a_{2}\wedge b_{2}).

For the next, we consider the case (ii). The submodularity is μ(a1,a2)+μ(b1,b2)μ(a1,b2)+μ(b1,a2)\mu(a_{1},a_{2})+\mu(b_{1},b_{2})\geq\mu(a_{1},b_{2})+\mu(b_{1},a_{2}). Since HμH_{\mu} is bipartite, d(a1,b2)d(a_{1},b_{2}) is equal to either d(b1,a2)d(b_{1},a_{2}) or d(b1,a2)+2d(b_{1},a_{2})+2 or d(b1,a2)2d(b_{1},a_{2})-2. If d(a1,b2)d(a_{1},b_{2}) is equal to d(b1,a2)+2d(b_{1},a_{2})+2 or d(b1,a2)2d(b_{1},a_{2})-2, then by Lemma 3.5 we have μ(a1,a2)+μ(b1,b2)=μ(a1,b2)+μ(b1,a2)\mu(a_{1},a_{2})+\mu(b_{1},b_{2})=\mu(a_{1},b_{2})+\mu(b_{1},a_{2}). Thus, it suffices to consider the case when d(a1,b2)=d(b1,a2)d(a_{1},b_{2})=d(b_{1},a_{2}). In this case, d(a1,a2)d(a_{1},a_{2}) is equal to either d(a1,b2)1d(a_{1},b_{2})-1 or d(a1,b2)+1d(a_{1},b_{2})+1. Suppose that d(a1,a2)=d(a1,b2)+1d(a_{1},a_{2})=d(a_{1},b_{2})+1. Then, by Lemma 3.5 we have μ(a1,a2)=μ(a1,b2)+μ(b2,a2)\mu(a_{1},a_{2})=\mu(a_{1},b_{2})+\mu(b_{2},a_{2}). Hence, we obtain μ(a1,a2)+μ(b1,b2)=μ(a1,b2)+μ(b2,a2)+μ(b1,b2)μ(a1,b2)+μ(b1,a2)\mu(a_{1},a_{2})+\mu(b_{1},b_{2})=\mu(a_{1},b_{2})+\mu(b_{2},a_{2})+\mu(b_{1},b_{2})\geq\mu(a_{1},b_{2})+\mu(b_{1},a_{2}). Consider the case when d(a1,a2)=d(a1,b2)1d(a_{1},a_{2})=d(a_{1},b_{2})-1. Similarly, d(b1,b2)d(b_{1},b_{2}) is equal to either d(a1,b2)1d(a_{1},b_{2})-1 or d(a1,b2)+1d(a_{1},b_{2})+1, and by the similar argument, we may assume that d(b1,b2)=d(a1,b2)1d(b_{1},b_{2})=d(a_{1},b_{2})-1. Let PP be a shortest path in HμH_{\mu} from a1a_{1} to a2a_{2}. Let zz be the vertex in PP that is adjacent to a1a_{1}. Then, we have d(z,b2)=d(b1,b2)=d(a1,b2)1d(z,b_{2})=d(b_{1},b_{2})=d(a_{1},b_{2})-1. Hence, by Lemma 2.5, there exists a common neighbor ww of z,b1z,b_{1} with d(w,b2)=d(a1,b2)2d(w,b_{2})=d(a_{1},b_{2})-2. Then, we have d(z,b2)=d(w,a2)=d(a1,b2)1,d(z,a2)=d(z,b2)1,d(z,b_{2})=d(w,a_{2})=d(a_{1},b_{2})-1,\ d(z,a_{2})=d(z,b_{2})-1, and d(w,b2)=d(z,b2)1d(w,b_{2})=d(z,b_{2})-1. Furthermore, since a1a_{1} covers b1b_{1}, we see that zz covers ww. Hence, we can apply the same argument to z,w,a2,b2z,w,a_{2},b_{2} which we apply to a1,b1,a2,b2a_{1},b_{1},a_{2},b_{2} above. By repeating this argument, we can see that a2a_{2} covers b2b_{2}, but this is a contradiction. ∎

4.2 Proof of Theorem 1.6

To prove the tractability, we construct fractional polymorphisms which satisfy the property of Theorem 2.1. Let rr denote the internal node of a star HμH_{\mu}. A subset XT{r}X\subseteq T\setminus\{r\} is called unbiased if Rμ(p,r)=Rμ(r,q)R_{\mu}(p,r)=R_{\mu}(r,q) holds for every p,qX(pq)p,q\in X\ (p\neq q). Note that if XX is unbiased and |X|3|X|\geq 3, then μ(p,r)=μ(r,p)\mu(p,r)=\mu(r,p) holds for any pXp\in X, since Rμ(p,r)=Rμ(r,q)=Rμ(s,r)=Rμ(r,p)R_{\mu}(p,r)=R_{\mu}(r,q)=R_{\mu}(s,r)=R_{\mu}(r,p) holds for q,sX{p}(qs)q,s\in X\setminus\{p\}\ (q\neq s). We divide T{r}T\setminus\{r\} into the minimum number of disjoint unbiased sets, and denote by \mathcal{F} the family of them. From the assumption that there exists no biased non-collinear triple, the family \mathcal{F} consists of at most two sets. Hence, it suffices to consider the following three cases:

  1. (i)

    ={X,Y}\mathcal{F}=\{X,Y\}, and |X|,|Y|2|X|,|Y|\leq 2.

  2. (ii)

    ={X,Y}\mathcal{F}=\{X,Y\}, and |X|3,|Y|2|X|\geq 3,\ |Y|\leq 2.

  3. (iii)

    ={X}\mathcal{F}=\{X\}.

We first consider the case (i). It suffices to consider the case when |X|=|Y|=2|X|=|Y|=2, since fractional polymorphisms for this case work for the other cases. Let X={x1,x2}X=\{x_{1},x_{2}\} and Y={y1,y2}Y=\{y_{1},y_{2}\}. Let \preceq be a partial order on TT defined by yirxj(i=1,2,j=1,2)y_{i}\prec r\prec x_{j}\ (i=1,2,\ j=1,2) (see Figure 4 (a)). For i=1,2i=1,2, we extend \preceq to i\preceq_{i} by adding relation yiiyj(ji)y_{i}\prec_{i}y_{j}\ (j\neq i). Then each pair of two elements t1,t2t_{1},t_{2} in a partially ordered set (T,i)(T,\preceq_{i}) has a unique meet, denoted by t1it2t_{1}\wedge_{i}t_{2} (i=1,2i=1,2).

\begin{overpic}[width=398.33858pt]{star.eps} \put(13.0,23.0){$x_{1}$} \put(33.7,23.0){$x_{2}$} \put(26.0,13.0){$r$} \put(13.0,4.0){$y_{1}$} \put(33.7,4.0){$y_{2}$} \put(60.5,20.3){$x_{1}$} \put(85.7,20.3){$x_{4}$} \put(68.0,25.0){$x_{2}$} \put(77.7,24.7){$x_{3}$} \put(76.5,13.0){$r$} \put(63.0,4.0){$y_{1}$} \put(83.6,4.0){$y_{2}$} \end{overpic}
Figure 4: (a) the case (i)                      (b) the case (ii) (k=4k=4)

For i=1,2i=1,2, similar to i\preceq_{i}, we also extend \preceq to i\preceq_{i}^{\prime} by adding relation xjixi(ji)x_{j}\prec_{i}^{\prime}x_{i}\ (j\neq i). Then each pair of two elements t1,t2t_{1},t_{2} in (T,i)(T,\preceq_{i}^{\prime}) has a unique join, denoted by t1it2t_{1}\vee_{i}t_{2} (i=1,2i=1,2). Since XX and YY are unbiased, the lengths of the edges in HμH_{\mu} can be written as

μ(x1,r)=a,μ(r,x1)=b,μ(r,x2)=ka,μ(x2,r)=kb,\displaystyle\mu(x_{1},r)=a,\ \mu(r,x_{1})=b,\ \mu(r,x_{2})=ka,\ \mu(x_{2},r)=kb,
μ(r,y1)=c,μ(y1,r)=d,μ(y2,r)=lc,μ(r,y2)=ld.\displaystyle\mu(r,y_{1})=c,\ \mu(y_{1},r)=d,\ \mu(y_{2},r)=lc,\ \mu(r,y_{2})=ld. (4.4)

Since i,j\wedge_{i},\vee_{j} are semilattice operations, by Theorem 2.1 it suffices to show that for any t1,t2T×Tt^{1},t^{2}\in T\times T,

μ(t1)+μ(t2)1l+1μ(t11t2)+ll+1μ(t12t2)+1k+1μ(t11t2)+kk+1μ(t12t2).\displaystyle\mu(t^{1})+\mu(t^{2})\geq\frac{1}{l+1}\mu(t^{1}\wedge_{1}t^{2})+\frac{l}{l+1}\mu(t^{1}\wedge_{2}t^{2})+\frac{1}{k+1}\mu(t^{1}\vee_{1}t^{2})+\frac{k}{k+1}\mu(t^{1}\vee_{2}t^{2}). (4.5)

Let t1=(t11,t21)t^{1}=(t_{1}^{1},t_{2}^{1}) and t2=(t12,t22)t^{2}=(t_{1}^{2},t_{2}^{2}). If a pair ti1,ti2t_{i}^{1},t_{i}^{2} has both meet ti1ti2t_{i}^{1}\wedge t_{i}^{2} and join ti1ti2t_{i}^{1}\vee t_{i}^{2} in (T,)(T,\preceq) for all i=1,2i=1,2, then (4.5) reduces to Theorem 1.4. Suppose that a pair ti1,ti2t_{i}^{1},t_{i}^{2} has no meet or has no join for some ii in (T,)(T,\preceq). Without loss of generality, we may assume that t11=x1t_{1}^{1}=x_{1} and t12=x2t_{1}^{2}=x_{2}. If t21=t22=yit_{2}^{1}=t_{2}^{2}=y_{i} (i{1,2}i\in\{1,2\}), then we have

μ(x1,yi)+μ(x2,yi)=μ(x1,r)+μ(r,yi)+μ(x2,r)+μ(r,yi)\displaystyle\mu(x_{1},y_{i})+\mu(x_{2},y_{i})=\mu(x_{1},r)+\mu(r,y_{i})+\mu(x_{2},r)+\mu(r,y_{i})
1k+1μ(x1,r)+1k+1μ(r,yi)+kk+1μ(r,yi)+kk+1μ(x2,r)+1l+1μ(r,yi)+ll+1μ(r,yi)\displaystyle\geq\frac{1}{k+1}\mu(x_{1},r)+\frac{1}{k+1}\mu(r,y_{i})+\frac{k}{k+1}\mu(r,y_{i})+\frac{k}{k+1}\mu(x_{2},r)+\frac{1}{l+1}\mu(r,y_{i})+\frac{l}{l+1}\mu(r,y_{i})
=1l+1μ(x11x2,yi1yi)+ll+1μ(x12x2,yi2yi)+1k+1μ(x11x2,yi1yi)\displaystyle=\frac{1}{l+1}\mu(x_{1}\wedge_{1}x_{2},y_{i}\wedge_{1}y_{i})+\frac{l}{l+1}\mu(x_{1}\wedge_{2}x_{2},y_{i}\wedge_{2}y_{i})+\frac{1}{k+1}\mu(x_{1}\vee_{1}x_{2},y_{i}\vee_{1}y_{i})
+kk+1μ(x12x2,yi2yi).\displaystyle+\frac{k}{k+1}\mu(x_{1}\vee_{2}x_{2},y_{i}\vee_{2}y_{i}). (4.6)

Thus, (4.5) holds. Suppose that t21,t22{r,y1,y2}t_{2}^{1},t_{2}^{2}\in\{r,y_{1},y_{2}\} and (t21,t22)(yi,yi)(t_{2}^{1},t_{2}^{2})\neq(y_{i},y_{i}) for i=1,2i=1,2. Then we have t21it22{t21,t22}t_{2}^{1}\wedge_{i}t_{2}^{2}\in\{t_{2}^{1},t_{2}^{2}\} and t21it22=rt_{2}^{1}\vee_{i}t_{2}^{2}=r for i=1,2i=1,2. Hence, we have

μ(x1,t21)+μ(x2,t22)=μ(x1,r)+μ(x2,r)+μ(r,t21)+μ(r,t22)\displaystyle\mu(x_{1},t_{2}^{1})+\mu(x_{2},t_{2}^{2})=\mu(x_{1},r)+\mu(x_{2},r)+\mu(r,t_{2}^{1})+\mu(r,t_{2}^{2})
1k+1μ(x1,r)+kk+1μ(x2,r)+1l+1μ(r,t211t22)+ll+1μ(r,t212t22)\displaystyle\geq\frac{1}{k+1}\mu(x_{1},r)+\frac{k}{k+1}\mu(x_{2},r)+\frac{1}{l+1}\mu(r,t_{2}^{1}\wedge_{1}t_{2}^{2})+\frac{l}{l+1}\mu(r,t_{2}^{1}\wedge_{2}t_{2}^{2})
=1l+1μ(x11x2,t211t22)+ll+1μ(x12x2,t212t22)+1k+1μ(x11x2,t211t22)\displaystyle=\frac{1}{l+1}\mu(x_{1}\wedge_{1}x_{2},t_{2}^{1}\wedge_{1}t_{2}^{2})+\frac{l}{l+1}\mu(x_{1}\wedge_{2}x_{2},t_{2}^{1}\wedge_{2}t_{2}^{2})+\frac{1}{k+1}\mu(x_{1}\vee_{1}x_{2},t_{2}^{1}\vee_{1}t_{2}^{2})
+kk+1μ(x12x2,t212t22).\displaystyle+\frac{k}{k+1}\mu(x_{1}\vee_{2}x_{2},t_{2}^{1}\vee_{2}t_{2}^{2}). (4.7)

Thus, it suffices to consider the case when t21=x1t_{2}^{1}=x_{1} or x2x_{2}. We first consider the case when t21=x1t_{2}^{1}=x_{1}. If t22=x2t_{2}^{2}=x_{2}, then (4.5) holds trivially, since the both sides of (4.5) are 0. If t22{x1,r,y1,y2}t_{2}^{2}\in\{x_{1},r,y_{1},y_{2}\}, then we have

μ(x1,x1)+μ(x2,t22)=μ(x2,r)+μ(r,t22)\displaystyle\mu(x_{1},x_{1})+\mu(x_{2},t_{2}^{2})=\mu(x_{2},r)+\mu(r,t_{2}^{2})
=kk+1μ(x2,x1)+1l+1μ(r,t22)+ll+1μ(r,t22)\displaystyle=\frac{k}{k+1}\mu(x_{2},x_{1})+\frac{1}{l+1}\mu(r,t_{2}^{2})+\frac{l}{l+1}\mu(r,t_{2}^{2})
=1l+1μ(x11x2,x11t22)+ll+1μ(x12x2,x12t22)+1k+1μ(x11x2,x11t22)\displaystyle=\frac{1}{l+1}\mu(x_{1}\wedge_{1}x_{2},x_{1}\wedge_{1}t_{2}^{2})+\frac{l}{l+1}\mu(x_{1}\wedge_{2}x_{2},x_{1}\wedge_{2}t_{2}^{2})+\frac{1}{k+1}\mu(x_{1}\vee_{1}x_{2},x_{1}\vee_{1}t_{2}^{2})
+kk+1μ(x12x2,x12t22),\displaystyle+\frac{k}{k+1}\mu(x_{1}\vee_{2}x_{2},x_{1}\vee_{2}t_{2}^{2}), (4.8)

where we use μ(x2,r)=kb=kk+1μ(x2,x1)\mu(x_{2},r)=kb=\frac{k}{k+1}\mu(x_{2},x_{1}) and μ(x11x2,x11t22)=μ(x1,x1)=0\mu(x_{1}\vee_{1}x_{2},x_{1}\vee_{1}t_{2}^{2})=\mu(x_{1},x_{1})=0. Thus, (4.5) holds. We next consider the case when t21=x2t_{2}^{1}=x_{2}. If t22=x1t_{2}^{2}=x_{1}, then (4.5) holds trivially, since the right hand side of (4.5) is 0. If t22{r,y1,y2}t_{2}^{2}\in\{r,y_{1},y_{2}\}, then we have

μ(x1,x2)+μ(x2,t22)=μ(x1,x2)+μ(x2,r)+μ(r,t22)\displaystyle\mu(x_{1},x_{2})+\mu(x_{2},t_{2}^{2})=\mu(x_{1},x_{2})+\mu(x_{2},r)+\mu(r,t_{2}^{2})
1k+1μ(x1,x2)+1l+1μ(r,t22)+ll+1μ(r,t22)\displaystyle\geq\frac{1}{k+1}\mu(x_{1},x_{2})+\frac{1}{l+1}\mu(r,t_{2}^{2})+\frac{l}{l+1}\mu(r,t_{2}^{2})
=1l+1μ(x11x2,x21t22)+ll+1μ(x12x2,x22t22)+1k+1μ(x11x2,x21t22)\displaystyle=\frac{1}{l+1}\mu(x_{1}\wedge_{1}x_{2},x_{2}\wedge_{1}t_{2}^{2})+\frac{l}{l+1}\mu(x_{1}\wedge_{2}x_{2},x_{2}\wedge_{2}t_{2}^{2})+\frac{1}{k+1}\mu(x_{1}\vee_{1}x_{2},x_{2}\vee_{1}t_{2}^{2})
+kk+1μ(x12x2,x22t22).\displaystyle+\frac{k}{k+1}\mu(x_{1}\vee_{2}x_{2},x_{2}\vee_{2}t_{2}^{2}). (4.9)

Thus, (4.5) holds. This completes the proof of the case (i).

We next consider the case (ii). Let X={x1,x2,,xk}X=\{x_{1},x_{2},\ldots,x_{k}\} and Y={y1,y2}Y=\{y_{1},y_{2}\} (k3k\geq 3). Let \preceq be a partial order on TT defined by yirxj(i=1,2,j=1,2,k)y_{i}\prec r\prec x_{j}\ (i=1,2,\ j=1,2,\ldots k) (see Figure 4 (b)). Similar to the case (i), for i=1,2i=1,2, we extend \preceq to i\preceq_{i} by adding relation yiiyj(ji)y_{i}\prec_{i}y_{j}\ (j\neq i). Then each pair of two elements t1,t2t_{1},t_{2} in (T,i)(T,\preceq_{i}) has a unique meet, denoted by t1it2t_{1}\wedge_{i}t_{2} (i=1,2i=1,2). Let [k]:={1,2,,k}[k]:=\{1,2,\ldots,k\}. Let 𝒢\mathcal{G} be the set of all functions g:([k]2)[k]g:{\scriptsize\begin{pmatrix}[k]\\ 2\\ \end{pmatrix}}\rightarrow[k] that satisfy g(ij){i,j}g(ij)\in\{i,j\} for any ij([k]2)ij\in{\scriptsize\begin{pmatrix}[k]\\ 2\\ \end{pmatrix}}. Then, for g𝒢g\in\mathcal{G}, let g\vee_{g} be a binary operation on TT defined by xigxj:=xg(ij)x_{i}\vee_{g}x_{j}:=x_{g(ij)} for ij([k]2)ij\in{\scriptsize\begin{pmatrix}[k]\\ 2\\ \end{pmatrix}}, and t1gt2:=t1t2t_{1}\vee_{g}t_{2}:=t_{1}\vee t_{2} for t1t2(T2)(X2)t_{1}t_{2}\in{\scriptsize\begin{pmatrix}T\\ 2\\ \end{pmatrix}}\setminus{\scriptsize\begin{pmatrix}X\\ 2\\ \end{pmatrix}}, where t1t2t_{1}\vee t_{2} is a unique join in (T,)(T,\preceq). By |X|3|X|\geq 3, it holds that μ(xi,r)=μ(r,xi)\mu(x_{i},r)=\mu(r,x_{i}) for any i[k]i\in[k]. Since YY is unbiased, the lengths of the edges in HμH_{\mu} can be written as

μ(xi,r)=μ(r,xi)=ai(i[k]),\displaystyle\mu(x_{i},r)=\mu(r,x_{i})=a_{i}\ (i\in[k]),
μ(r,y1)=b,μ(y1,r)=c,μ(y2,r)=lb,μ(r,y2)=lc.\displaystyle\mu(r,y_{1})=b,\ \mu(y_{1},r)=c,\ \mu(y_{2},r)=lb,\ \mu(r,y_{2})=lc. (4.10)

Since i\wedge_{i} is a semilattice operation, by Theorem 2.1 it suffices to show that for any t1,t2T×Tt^{1},t^{2}\in T\times T,

μ(t1)+μ(t2)1l+1μ(t11t2)+ll+1μ(t12t2)+g𝒢ij([k]2)ag(ij)ai+ajμ(t1gt2).\displaystyle\mu(t^{1})+\mu(t^{2})\geq\frac{1}{l+1}\mu(t^{1}\wedge_{1}t^{2})+\frac{l}{l+1}\mu(t^{1}\wedge_{2}t^{2})+\sum_{g\in\mathcal{G}}\prod_{\tiny ij\in\begin{pmatrix}[k]\\ 2\\ \end{pmatrix}\normalsize}\frac{a_{g(ij)}}{a_{i}+a_{j}}\mu(t^{1}\vee_{g}t^{2}). (4.11)

Note that g𝒢ij([k]2)ag(ij)=ij([k]2)(ai+aj)\sum_{g\in\mathcal{G}}\prod_{\tiny ij\in\begin{pmatrix}[k]\\ 2\\ \end{pmatrix}\normalsize}a_{g(ij)}=\prod_{\tiny ij\in\begin{pmatrix}[k]\\ 2\\ \end{pmatrix}\normalsize}(a_{i}+a_{j}). Let t1=(t11,t21)t^{1}=(t_{1}^{1},t_{2}^{1}) and t2=(t12,t22)t^{2}=(t_{1}^{2},t_{2}^{2}). If |{t11,t21,t12,t22}X|2|\{t_{1}^{1},t_{2}^{1},t_{1}^{2},t_{2}^{2}\}\cap X|\leq 2, then (4.11) reduces to the case (i). Consider the case when |{t11,t21,t12,t22}X|3|\{t_{1}^{1},t_{2}^{1},t_{1}^{2},t_{2}^{2}\}\cap X|\geq 3. Without loss of generality, we may assume that t11=x1,t21=x2,t12=x3t_{1}^{1}=x_{1},t_{2}^{1}=x_{2},t_{1}^{2}=x_{3}. If t22=xit_{2}^{2}=x_{i} (i2,3i\neq 2,3), then we have

μ(x1,x2)+μ(x3,xi)=μ(x1,r)+μ(r,x2)+μ(x3,r)+μ(r,xi)\displaystyle\mu(x_{1},x_{2})+\mu(x_{3},x_{i})=\mu(x_{1},r)+\mu(r,x_{2})+\mu(x_{3},r)+\mu(r,x_{i})
a1a1+a3a2a2+aiμ(x1,r)+a1a1+a3aia2+aiμ(x1,r)+a1a1+a3a2a2+aiμ(r,x2)\displaystyle\geq\frac{a_{1}}{a_{1}+a_{3}}\cdot\frac{a_{2}}{a_{2}+a_{i}}\mu(x_{1},r)+\frac{a_{1}}{a_{1}+a_{3}}\cdot\frac{a_{i}}{a_{2}+a_{i}}\mu(x_{1},r)+\frac{a_{1}}{a_{1}+a_{3}}\cdot\frac{a_{2}}{a_{2}+a_{i}}\mu(r,x_{2})
+a3a1+a3a2a2+aiμ(r,x2)+a3a1+a3a2a2+aiμ(x3,r)+a3a1+a3aia2+aiμ(x3,r)\displaystyle+\frac{a_{3}}{a_{1}+a_{3}}\cdot\frac{a_{2}}{a_{2}+a_{i}}\mu(r,x_{2})+\frac{a_{3}}{a_{1}+a_{3}}\cdot\frac{a_{2}}{a_{2}+a_{i}}\mu(x_{3},r)+\frac{a_{3}}{a_{1}+a_{3}}\cdot\frac{a_{i}}{a_{2}+a_{i}}\mu(x_{3},r)
+a1a1+a3aia2+aiμ(r,xi)+a3a1+a3aia2+aiμ(r,xi)\displaystyle+\frac{a_{1}}{a_{1}+a_{3}}\cdot\frac{a_{i}}{a_{2}+a_{i}}\mu(r,x_{i})+\frac{a_{3}}{a_{1}+a_{3}}\cdot\frac{a_{i}}{a_{2}+a_{i}}\mu(r,x_{i})
a1a1+a3a2a2+aiμ(x1,x2)+a1a1+a3aia2+aiμ(x1,xi)+a3a1+a3a2a2+aiμ(x3,x2)\displaystyle\geq\frac{a_{1}}{a_{1}+a_{3}}\cdot\frac{a_{2}}{a_{2}+a_{i}}\mu(x_{1},x_{2})+\frac{a_{1}}{a_{1}+a_{3}}\cdot\frac{a_{i}}{a_{2}+a_{i}}\mu(x_{1},x_{i})+\frac{a_{3}}{a_{1}+a_{3}}\cdot\frac{a_{2}}{a_{2}+a_{i}}\mu(x_{3},x_{2})
+a3a1+a3aia2+aiμ(x3,xi)\displaystyle+\frac{a_{3}}{a_{1}+a_{3}}\cdot\frac{a_{i}}{a_{2}+a_{i}}\mu(x_{3},x_{i})
=1l+1μ(x11x3,x21xi)+ll+1μ(x12x3,x22xi)\displaystyle=\frac{1}{l+1}\mu(x_{1}\wedge_{1}x_{3},x_{2}\wedge_{1}x_{i})+\frac{l}{l+1}\mu(x_{1}\wedge_{2}x_{3},x_{2}\wedge_{2}x_{i})
+g𝒢jm([k]2)ag(jm)aj+amμ(x1gx3,x2gxi).\displaystyle+\sum_{g\in\mathcal{G}}\prod_{\tiny jm\in\begin{pmatrix}[k]\\ 2\\ \end{pmatrix}\normalsize}\frac{a_{g(jm)}}{a_{j}+a_{m}}\mu(x_{1}\vee_{g}x_{3},x_{2}\vee_{g}x_{i}). (4.12)

(The second inequality is strict when xi=x1x_{i}=x_{1}.) If t22=x2t_{2}^{2}=x_{2}, then we have

μ(x1,x2)+μ(x3,x2)=μ(x1,r)+μ(r,x2)+μ(x3,r)+μ(r,x2)\displaystyle\mu(x_{1},x_{2})+\mu(x_{3},x_{2})=\mu(x_{1},r)+\mu(r,x_{2})+\mu(x_{3},r)+\mu(r,x_{2})
a1a1+a3μ(x1,r)+a1a1+a3μ(r,x2)+a3a1+a3μ(r,x2)+a3a1+a3μ(x3,r)+1l+1μ(r,x2)\displaystyle\geq\frac{a_{1}}{a_{1}+a_{3}}\mu(x_{1},r)+\frac{a_{1}}{a_{1}+a_{3}}\mu(r,x_{2})+\frac{a_{3}}{a_{1}+a_{3}}\mu(r,x_{2})+\frac{a_{3}}{a_{1}+a_{3}}\mu(x_{3},r)+\frac{1}{l+1}\mu(r,x_{2})
+ll+1μ(r,x2)\displaystyle+\frac{l}{l+1}\mu(r,x_{2})
=a1a1+a3μ(x1,x2)+a3a1+a3μ(x3,x2)+1l+1μ(r,x2)+ll+1μ(r,x2)\displaystyle=\frac{a_{1}}{a_{1}+a_{3}}\mu(x_{1},x_{2})+\frac{a_{3}}{a_{1}+a_{3}}\mu(x_{3},x_{2})+\frac{1}{l+1}\mu(r,x_{2})+\frac{l}{l+1}\mu(r,x_{2})
=1l+1μ(x11x3,x21x2)+ll+1μ(x12x3,x22x2)\displaystyle=\frac{1}{l+1}\mu(x_{1}\wedge_{1}x_{3},x_{2}\wedge_{1}x_{2})+\frac{l}{l+1}\mu(x_{1}\wedge_{2}x_{3},x_{2}\wedge_{2}x_{2})
+g𝒢jm([k]2)ag(jm)aj+amμ(x1gx3,x2gx2).\displaystyle+\sum_{g\in\mathcal{G}}\prod_{\tiny jm\in\begin{pmatrix}[k]\\ 2\\ \end{pmatrix}\normalsize}\frac{a_{g(jm)}}{a_{j}+a_{m}}\mu(x_{1}\vee_{g}x_{3},x_{2}\vee_{g}x_{2}). (4.13)

If t22=x3t_{2}^{2}=x_{3}, then we have

μ(x1,x2)+μ(x3,x3)=a1+a2\displaystyle\mu(x_{1},x_{2})+\mu(x_{3},x_{3})=a_{1}+a_{2}
=a1a2(a1+a2)+a2a3(a2+a3)+a3a1(a3+a1)+2a1a2a3(a1+a3)(a2+a3)\displaystyle=\frac{a_{1}a_{2}(a_{1}+a_{2})+a_{2}a_{3}(a_{2}+a_{3})+a_{3}a_{1}(a_{3}+a_{1})+2a_{1}a_{2}a_{3}}{(a_{1}+a_{3})(a_{2}+a_{3})}
a1a2(a1+a2)+a2a3(a2+a3)+a3a1(a3+a1)(a1+a3)(a2+a3)\displaystyle\geq\frac{a_{1}a_{2}(a_{1}+a_{2})+a_{2}a_{3}(a_{2}+a_{3})+a_{3}a_{1}(a_{3}+a_{1})}{(a_{1}+a_{3})(a_{2}+a_{3})}
=a1a1+a3a2a2+a3μ(x1,x2)+a3a1+a3a2a2+a3μ(x3,x2)+a1a1+a3a3a2+a3μ(x1,x3)\displaystyle=\frac{a_{1}}{a_{1}+a_{3}}\cdot\frac{a_{2}}{a_{2}+a_{3}}\mu(x_{1},x_{2})+\frac{a_{3}}{a_{1}+a_{3}}\cdot\frac{a_{2}}{a_{2}+a_{3}}\mu(x_{3},x_{2})+\frac{a_{1}}{a_{1}+a_{3}}\cdot\frac{a_{3}}{a_{2}+a_{3}}\mu(x_{1},x_{3})
=1l+1μ(x11x3,x21x3)+ll+1μ(x12x3,x22x3)\displaystyle=\frac{1}{l+1}\mu(x_{1}\wedge_{1}x_{3},x_{2}\wedge_{1}x_{3})+\frac{l}{l+1}\mu(x_{1}\wedge_{2}x_{3},x_{2}\wedge_{2}x_{3})
+g𝒢jm([k]2)ag(jm)aj+amμ(x1gx3,x2gx3).\displaystyle+\sum_{g\in\mathcal{G}}\prod_{\tiny jm\in\begin{pmatrix}[k]\\ 2\\ \end{pmatrix}\normalsize}\frac{a_{g(jm)}}{a_{j}+a_{m}}\mu(x_{1}\vee_{g}x_{3},x_{2}\vee_{g}x_{3}). (4.14)

If t22{r,y1,y2}t_{2}^{2}\in\{r,y_{1},y_{2}\}, then we have

μ(x1,x2)+μ(x3,t22)=μ(x1,r)+μ(r,x2)+μ(x3,r)+μ(r,t22)\displaystyle\mu(x_{1},x_{2})+\mu(x_{3},t_{2}^{2})=\mu(x_{1},r)+\mu(r,x_{2})+\mu(x_{3},r)+\mu(r,t_{2}^{2})
=μ(x1,r)+a1a1+a3μ(r,x2)+a3a1+a3μ(r,x2)+μ(x3,r)+μ(r,t22)\displaystyle=\mu(x_{1},r)+\frac{a_{1}}{a_{1}+a_{3}}\mu(r,x_{2})+\frac{a_{3}}{a_{1}+a_{3}}\mu(r,x_{2})+\mu(x_{3},r)+\mu(r,t_{2}^{2})
a1a1+a3μ(x1,x2)+a3a1+a3μ(x3,x2)+1l+1μ(r,t22)+ll+1μ(r,t22)\displaystyle\geq\frac{a_{1}}{a_{1}+a_{3}}\mu(x_{1},x_{2})+\frac{a_{3}}{a_{1}+a_{3}}\mu(x_{3},x_{2})+\frac{1}{l+1}\mu(r,t_{2}^{2})+\frac{l}{l+1}\mu(r,t_{2}^{2})
=1l+1μ(x11x3,x21t22)+ll+1μ(x12x3,x22t22)\displaystyle=\frac{1}{l+1}\mu(x_{1}\wedge_{1}x_{3},x_{2}\wedge_{1}t_{2}^{2})+\frac{l}{l+1}\mu(x_{1}\wedge_{2}x_{3},x_{2}\wedge_{2}t_{2}^{2})
+g𝒢jm([k]2)ag(jm)aj+amμ(x1gx3,x2gt22).\displaystyle+\sum_{g\in\mathcal{G}}\prod_{\tiny jm\in\begin{pmatrix}[k]\\ 2\\ \end{pmatrix}\normalsize}\frac{a_{g(jm)}}{a_{j}+a_{m}}\mu(x_{1}\vee_{g}x_{3},x_{2}\vee_{g}t_{2}^{2}). (4.15)

This completes the proof of the case (ii).

We finally consider the case (iii). Let ω\omega be the fractional polymorphism constructed in the case (ii), where ={X,Y}\mathcal{F}=\{X^{\prime},Y^{\prime}\} and |X|3,|Y|2|X^{\prime}|\geq 3,|Y^{\prime}|\leq 2. For any operation φ\varphi in supp(ω)\mathrm{supp}(\omega) and any x1,x2X,y1,y2Yx_{1},x_{2}\in X^{\prime},\ y_{1},y_{2}\in Y^{\prime}, we have

φ(x1,r),φ(r,x1),φ(x1,x2)X{r},\displaystyle\varphi(x_{1},r),\varphi(r,x_{1}),\varphi(x_{1},x_{2})\in X^{\prime}\cup\{r\},
φ(y1,r),φ(r,y1),φ(y1,y2)Y{r}.\displaystyle\varphi(y_{1},r),\varphi(r,y_{1}),\varphi(y_{1},y_{2})\in Y^{\prime}\cup\{r\}.

Hence, ω\omega also works for the case of ={X}\mathcal{F}=\{X^{\prime}\} or ={Y}\mathcal{F}=\{Y^{\prime}\}, which completes the proof.

5 Proof of hardness

In this section, we give proofs of Theorem 1.3 and Theorem 1.5. We prove them with the aid of Proposition 2.3; for a directed metric μ\mu satisfying the condition in Theorem 1.3 or Theorem 1.5, and the corresponding constraint language Γ\Gamma defined in (2.1), we show that Γ\Gamma satisfies the condition (MC). To prove this, we construct a “gadget” which is a counterexample to submodularity of the objective function of 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] (in a certain sense). This type of hardness proof using a gadget and the condition (MC) originates from the hardness proof of the multiterminal cut problem [7]. Karzanov [13, 14] also adopted the similar approach to show the hardness results of 0-Ext[μ][\mu] (Theorem 1.1). We apply this type of hardness proof to the directed version 0\overrightarrow{\textbf{0}}-Ext[μ][\mu]. We first describe the sufficient condition for Γ\Gamma defined in (2.1) to satisfy the (MC) condition in Section 5.1. For the next, we prove NP-hardness of 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] by using this sufficient condition in Section 5.2.

5.1 Approach

Let μ\mu be a rational-valued directed metric on TT. Suppose that we are given VTV\supseteq T and c:V×V+c:V\times V\rightarrow\mathbb{Q}_{+} as an instance of 0\overrightarrow{\textbf{0}}-Ext[μ][\mu]. For s0,s1,,skTs_{0},s_{1},\ldots,s_{k}\in T and x0,x1,,xkVTx_{0},x_{1},\ldots,x_{k}\in V\setminus T, we denote by τc(s0,x0|s1,x1||sk,xk)\tau_{c}(s_{0},x_{0}|s_{1},x_{1}|\cdots|s_{k},x_{k}) the optimal value of 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] subject to γ(x0)=s0,γ(x1)=s1,,γ(xk)=sk\gamma(x_{0})=s_{0},\gamma(x_{1})=s_{1},\ldots,\gamma(x_{k})=s_{k}. We simply denote τ(s0,x0|s1,x1||sk,xk):=τc(s0,x0|s1,x1||sk,xk)\tau(s_{0},x_{0}|s_{1},x_{1}|\cdots|s_{k},x_{k}):=\tau_{c}(s_{0},x_{0}|s_{1},x_{1}|\cdots|s_{k},x_{k}) if cc is clear in the context. Let τ\tau^{*} be the optimal value of 0\overrightarrow{\textbf{0}}-Ext[μ][\mu]. Similar to the constructions in [7, 13, 14], we desire a pair (called gadget) (V,c)(V,c) satisfying the following properties (in other words, “violates submodularity,” cf. [7]) for specified elements s,tTs,t\in T and x,yVTx,y\in V\setminus T.

(i)\displaystyle\mathrm{(i)} τ(s,x|t,y)=τ(t,x|s,y)=τ,\displaystyle\mathrm{\ \ \ }\tau(s,x|t,y)=\tau(t,x|s,y)=\tau^{*},
(ii)\displaystyle\mathrm{(ii)} τ(s,x|s,y)=τ(t,x|t,y)=τ+δforsomeδ>0,\displaystyle\mathrm{\ \ \ }\tau(s,x|s,y)=\tau(t,x|t,y)=\tau^{*}+\delta\mathrm{\ \ \ }\mathrm{for\ some\ \delta>0}, (5.1)
(iii)\displaystyle\mathrm{(iii)} τ(s,x|t,y)τ+δforallotherpairs(s,t)T×T.\displaystyle\mathrm{\ \ \ }\tau(s^{\prime},x|t^{\prime},y)\geq\tau^{*}+\delta\mathrm{\ \ \ }\mathrm{for\ all\ other\ pairs}\ (s^{\prime},t^{\prime})\in T\times T.

Let Γ\Gamma be a constraint language defined in (2.1). We show that the existence of a gadget (V,c)(V,c) satisfying (5.1) implies NP-hardness of VCSP[Γ][\Gamma]. Suppose that there exists a gadget (V,c)(V,c) satisfying (5.1) for s,tTs,t\in T and x,yVTx,y\in V\setminus T. Then, we define a function f:T×T+f:T\times T\rightarrow\mathbb{Q}_{+} as follows:

f(t1,t2):=τ(t1,x|t2,y)(t1,t2T).\displaystyle f(t_{1},t_{2}):=\tau(t_{1},x|t_{2},y)\ \ \ (t_{1},t_{2}\in T). (5.2)

We have fΓf\in\langle\Gamma\rangle and argminf={(s,t),(t,s)}\mathrm{argmin}\ f=\{(s,t),(t,s)\}. Hence, Γ\Gamma satisfies the condition (MC), which implies that VCSP[Γ][\Gamma] is strongly NP-hard by Proposition 2.3.

5.2 Proofs

In this section, we show Theorem 1.3 and Theorem 1.5 by constructing gadgets (V,c)(V,c) satisfying (5.1). We can state Theorem 1.3 in the following equivalent form:

Theorem 5.1.

Let μ\mu be a rational-valued directed metric. 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] is strongly NP-hard if one of the following conditions holds:

  1. (i)

    μ\mu is not modular.

  2. (ii)

    μ\mu is modular and not directed orbit-invariant.

  3. (iii)

    μ\mu is modular and directed orbit-invariant, and HμH_{\mu} is not orientable.

We prove each case of Theorem 5.1 and Theorem 1.5.

5.2.1 Proof of Theorem 5.1 for the case (i)

We first show the following lemma, which originates from the proof of Theorem 1.1 (i) in [14].

Lemma 5.2.

Let μ\mu be a rational-valued directed metric on TT. If there exists a pair (V,c)(V,c) which satisfies the following properties for a non-collinear triple (s0,s1,s2)(s_{0},s_{1},s_{2}) in TT and distinct elements z0,z1,,z5VTz_{0},z_{1},\ldots,z_{5}\in V\setminus T, then 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] is strongly NP-hard.

(i)\displaystyle\mathrm{(i)} τ(si0+1,z0|si11,z1|si2,z2|si3+1,z3|si41,z4|si5,z5)=τ(ij{0,1}for eachj),\displaystyle\mathrm{\ \ \ }\tau(s_{i_{0}+1},z_{0}|s_{i_{1}-1},z_{1}|s_{i_{2}},z_{2}|s_{i_{3}+1},z_{3}|s_{i_{4}-1},z_{4}|s_{i_{5}},z_{5})=\tau^{*}\mathrm{\ \ \ }(i_{j}\in\{0,1\}\mathrm{\ \textit{for\ each}\ }j),
(ii)\displaystyle\mathrm{(ii)} τ(s0,z0|s1,z1|s2,z2|s3,z3|s4,z4|s5,z5)τ+δ\displaystyle\mathrm{\ \ \ }\tau(s^{\prime}_{0},z_{0}|s^{\prime}_{1},z_{1}|s^{\prime}_{2},z_{2}|s^{\prime}_{3},z_{3}|s^{\prime}_{4},z_{4}|s^{\prime}_{5},z_{5})\geq\tau^{*}+\delta
for all other sextuplets s0,s1,s2,s3,s4,s5T and someδ>0,\displaystyle\mathrm{\ \ \ }\mathrm{\textit{for\ all\ other\ sextuplets\ }}s_{0}^{\prime},s_{1}^{\prime},s_{2}^{\prime},s_{3}^{\prime},s_{4}^{\prime},s_{5}^{\prime}\in T\mathrm{\textit{\ and\ some}\ \delta>0}, (5.3)

where the indices of sis_{i} are taken modulo 3.

Proof.

Let (V,c)(V,c) be a pair which satisfies (5.2) with respect to a non-collinear triple (s0,s1,s2)(s_{0},s_{1},s_{2}) in TT and distinct elements ziVT(i=0,,5)z_{i}\in V\setminus T\ (i=0,\ldots,5). We show NP-hardness of 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] by constructing a gadget based on (V,c)(V,c) which satisfies (5.1). Let μi:=μ(si1,si+1)+μ(si+1,si1)\mu_{i}:=\mu(s_{i-1},s_{i+1})+\mu(s_{i+1},s_{i-1}) and hi:=(μi1+μi+1μi)/2h_{i}:=(\mu_{i-1}+\mu_{i+1}-\mu_{i})/2 for i=0,1,2i=0,1,2 (the indices of μi\mu_{i} are taken modulo 3). Then we define a function c:V×V+c^{\prime}:V\times V\rightarrow\mathbb{Q}_{+} as follows (see Figure 5):

c(zi,zi+1)=c(zi+1,zi):=hi1(0i5).\displaystyle c^{\prime}(z_{i},z_{i+1})=c^{\prime}(z_{i+1},z_{i}):=h_{i-1}\mathrm{\ \ \ }(0\leq i\leq 5). (5.4)
\begin{overpic}[width=398.33858pt]{nphfig1.eps} \put(23.0,4.0){$s_{0}$} \put(42.5,32.0){$s_{1}$} \put(66.0,4.0){$s_{2}$} \put(37.0,2.0){$z_{1}$} \put(53.0,2.0){$z_{4}$} \put(26.0,16.5){$z_{2}$} \put(64.0,16.5){$z_{3}$} \put(32.0,24.0){$z_{5}$} \put(60.0,24.0){$z_{0}$} \end{overpic}
Figure 5: the function cc^{\prime}

Here the indices of ziz_{i} are taken modulo 6, and the indices of hih_{i} are taken modulo 3. Also we define c(v0,v1):=0c^{\prime}(v_{0},v_{1}):=0 for all other pairs (v0,v1)V×V(v_{0},v_{1})\in V\times V (undefined values of other functions below are also 0). Let NN be a sufficiently large positive rational (for example, N:=1/δ+4(h0+h1+h2)max{μ(s,t)s,tT}/δN:=1/\delta+4(h_{0}+h_{1}+h_{2})\cdot\mathrm{max}\{\mu(s,t)\mid s,t\in T\}/\delta). We define a function c~\tilde{c} by c~:=Nc+c\tilde{c}:=Nc+c^{\prime}. Let s:=s0,t:=s2,x:=z1,y:=z4s:=s_{0},t:=s_{2},x:=z_{1},y:=z_{4}. We now show that the pair (V,c~)(V,\tilde{c}) satisfies (5.1) with respect to s,t,x,ys,t,x,y. For rsi1,si+1r\neq s_{i-1},s_{i+1}, the value τ(r,zi)\tau(r,z_{i}) is so large that a map γ\gamma with γ(zi)=r\gamma(z_{i})=r is not optimal or nearly optimal, since (V,c)(V,c) satisfies (5.2) and NN is sufficiently large. We call such a map infeasible. Therefore, it suffices to consider the case when γ(zi){si1,si+1}\gamma(z_{i})\in\{s_{i-1},s_{i+1}\} for every ii. Let ρ:=2(h0h1+h1h2+h2h0),α:=2min{h02,h12,h22}\rho:=2(h_{0}h_{1}+h_{1}h_{2}+h_{2}h_{0}),\ \alpha:=2\mathrm{min}\{h_{0}^{2},h_{1}^{2},h_{2}^{2}\}. Without loss of generality, we may assume that α=2h02\alpha=2h_{0}^{2}. In the case of γ(x)=s\gamma(x)=s and γ(y)=t\gamma(y)=t, we have

τc~(s1,z0|s0,z1|s0,z2|s2,z3|s2,z4|s1,z5)\displaystyle\tau_{\tilde{c}}(s_{1},z_{0}|s_{0},z_{1}|s_{0},z_{2}|s_{2},z_{3}|s_{2},z_{4}|s_{1},z_{5}) =h1μ1+h2μ2+h0μ0+Nτc\displaystyle=h_{1}\mu_{1}+h_{2}\mu_{2}+h_{0}\mu_{0}+N\tau_{c}^{*}
=h1(h2+h0)+h2(h0+h1)+h0(h1+h2)+Nτc\displaystyle=h_{1}(h_{2}+h_{0})+h_{2}(h_{0}+h_{1})+h_{0}(h_{1}+h_{2})+N\tau_{c}^{*}
=ρ+Nτc,\displaystyle=\rho+N\tau_{c}^{*}, (5.5)

where τc\tau_{c}^{*} is the optimal value of 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] with respect to (V,c)(V,c). Similarly, in the case of γ(x)=t\gamma(x)=t and γ(y)=s\gamma(y)=s, we have

τc~(s2,z0|s2,z1|s1,z2|s1,z3|s0,z4|s0,z5)=ρ+Nτc.\displaystyle\tau_{\tilde{c}}(s_{2},z_{0}|s_{2},z_{1}|s_{1},z_{2}|s_{1},z_{3}|s_{0},z_{4}|s_{0},z_{5})=\rho+N\tau_{c}^{*}. (5.6)

On the other hand, in the case of γ(x)=γ(y)=t\gamma(x)=\gamma(y)=t, we have

τc~(s2,z0|s2,z1|s0,z2|s2,z3|s2,z4|s0,z5)\displaystyle\tau_{\tilde{c}}(s_{2},z_{0}|s_{2},z_{1}|s_{0},z_{2}|s_{2},z_{3}|s_{2},z_{4}|s_{0},z_{5}) =(h0+h0+h1+h1)μ1+Nτc\displaystyle=(h_{0}+h_{0}+h_{1}+h_{1})\mu_{1}+N\tau_{c}^{*}
=2(h0+h1)(h2+h0)+Nτc\displaystyle=2(h_{0}+h_{1})(h_{2}+h_{0})+N\tau_{c}^{*}
=α+ρ+Nτc.\displaystyle=\alpha+\rho+N\tau_{c}^{*}. (5.7)

Also, in the case of γ(x)=γ(y)=s\gamma(x)=\gamma(y)=s, we have

τc~(s1,z0|s0,z1|s1,z2|s1,z3|s0,z4|s1,z5)\displaystyle\tau_{\tilde{c}}(s_{1},z_{0}|s_{0},z_{1}|s_{1},z_{2}|s_{1},z_{3}|s_{0},z_{4}|s_{1},z_{5}) =(h2+h2+h0+h0)μ2+Nτc\displaystyle=(h_{2}+h_{2}+h_{0}+h_{0})\mu_{2}+N\tau_{c}^{*}
=2(h2+h0)(h0+h1)+Nτc\displaystyle=2(h_{2}+h_{0})(h_{0}+h_{1})+N\tau_{c}^{*}
=α+ρ+Nτc.\displaystyle=\alpha+\rho+N\tau_{c}^{*}. (5.8)

Let τc(γ)\tau_{c}(\gamma) denote the value of the objective function of 0\overrightarrow{\textbf{0}}-Ext[μ][\mu] with a map γ\gamma, where the input is (V,c)(V,c). We simply denote τ(γ):=τc(γ)\tau(\gamma):=\tau_{c}(\gamma) if cc is clear in the context. To finish the proof, we show that τc~(γ)α+ρ+Nτc\tau_{\tilde{c}}(\gamma)\geq\alpha+\rho+N\tau_{c}^{*} if a map γ\gamma is distinct from the assignments in (5.2.1) and (5.6). Let ξ\xi be the contribution to the value τc~(γ)\tau_{\tilde{c}}(\gamma) from cc^{\prime}. The value ξ\xi is represented by a sum of hihj(0i,j2)h_{i}h_{j}\ (0\leq i,j\leq 2). Let gig_{i} be the contribution to the value τc~(γ)\tau_{\tilde{c}}(\gamma) from c(zi,zi+1)c^{\prime}(z_{i},z_{i+1}) and c(zi+1,zi)c^{\prime}(z_{i+1},z_{i}) for each ii. We have the following four cases:

  1. (i)

    γ(zi)=γ(zi+1)=si1\gamma(z_{i})=\gamma(z_{i+1})=s_{i-1},

  2. (ii)

    γ(zi)=si+1,γ(zi+1)=si\gamma(z_{i})=s_{i+1},\ \gamma(z_{i+1})=s_{i},

  3. (iii)

    γ(zi)=si+1,γ(zi+1)=si1\gamma(z_{i})=s_{i+1},\ \gamma(z_{i+1})=s_{i-1},

  4. (iv)

    γ(zi)=si1,γ(zi+1)=si\gamma(z_{i})=s_{i-1},\ \gamma(z_{i+1})=s_{i}.

We have gi=0,gi=hi1hi+hi1hi+1,gi=hi1hi+1+hi12,gi=hi1hi+hi12g_{i}=0,\ g_{i}=h_{i-1}h_{i}+h_{i-1}h_{i+1},\ g_{i}=h_{i-1}h_{i+1}+h_{i-1}^{2},\ g_{i}=h_{i-1}h_{i}+h_{i-1}^{2} in the cases of (i), (ii), (iii), (iv), respectively. We call a pair zizi+1z_{i}z_{i+1} slanting if ziz_{i} and zi+1z_{i+1} satisfy (iii) or (iv). If zizi+1z_{i}z_{i+1} is not slanting for any ii with respect to γ\gamma, then γ\gamma corresponds to the assignment in (5.2.1) or (5.6). If zizi+1z_{i}z_{i+1} is slanting for some i{0,,5}i\in\{0,\ldots,5\}, then zjzj+1z_{j}z_{j+1} is also slanting for another j{0,,5}j\in\{0,\ldots,5\}. Hence, ξ\xi contains hi12+hj12αh_{i-1}^{2}+h_{j-1}^{2}\geq\alpha in its representation. Also, focusing on the pairs z0z1z_{0}z_{1} and z1z2z_{1}z_{2}, we can see that g0g_{0} includes h2h0h_{2}h_{0} in its representation when γ(z1)=s0\gamma(z_{1})=s_{0}, and g1g_{1} includes h2h0h_{2}h_{0} when γ(z1)=s2\gamma(z_{1})=s_{2}. Similarly, we can see that gig_{i} or gi+1g_{i+1} includes hi1hih_{i-1}h_{i} for each i{0,,5}i\in\{0,\ldots,5\}. This completes the proof. ∎

We now show Theorem 5.1 for the case (i) by use of Lemma 5.2. The proof we describe below is a directed version of that of Theorem 1.1 (i) in [14]. Let μ\mu be a nonmodular rational-valued directed metric on TT. For x,y,zTx,y,z\in T, we denote Δ(x,y,z):=μ(x,y)+μ(y,x)+μ(y,z)+μ(z,y)+μ(z,x)+μ(x,z)\Delta(x,y,z):=\mu(x,y)+\mu(y,x)+\mu(y,z)+\mu(z,y)+\mu(z,x)+\mu(x,z). Let (s0,s1,s2)(s_{0},s_{1},s_{2}) be a medianless triple such that Δ(s0,s1,s2)\Delta(s_{0},s_{1},s_{2}) is minimum. Let Δ¯:=Δ(s0,s1,s2)\bar{\Delta}:=\Delta(s_{0},s_{1},s_{2}). Take six elements z0,z1,,z5z_{0},z_{1},\ldots,z_{5}, and let V:=T{z0,z1,,z5}V:=T\cup\{z_{0},z_{1},\ldots,z_{5}\}. Let μi:=μ(si1,si+1)+μ(si+1,si1)\mu_{i}:=\mu(s_{i-1},s_{i+1})+\mu(s_{i+1},s_{i-1}) and ai:=(μi1+μi+1μi)/μi1μi+1a_{i}:=(\mu_{i-1}+\mu_{i+1}-\mu_{i})/\mu_{i-1}\mu_{i+1} for i=0,1,2i=0,1,2, where the indices of sis_{i} and μi\mu_{i} are taken modulo 3. Then we define a function c:V×V+c:V\times V\rightarrow\mathbb{Q}_{+} as follows:

c(si,zi+1)=c(zi+1,si)=1(0i5),\displaystyle c(s_{i},z_{i+1})=c(z_{i+1},s_{i})=1\mathrm{\ \ \ }(0\leq i\leq 5),
c(si,zi+2)=c(zi+2,si)=1(0i5),\displaystyle c(s_{i},z_{i+2})=c(z_{i+2},s_{i})=1\mathrm{\ \ \ }(0\leq i\leq 5), (5.9)

where the indices of ziz_{i} are taken modulo 6. Also we define a function c:V×V+c^{\prime}:V\times V\rightarrow\mathbb{Q}_{+} as follows:

c(si,zj)=c(zj,si)=ai(0i2, 0j5).\displaystyle c^{\prime}(s_{i},z_{j})=c^{\prime}(z_{j},s_{i})=a_{i}\mathrm{\ \ \ }(0\leq i\leq 2,\ 0\leq j\leq 5). (5.10)

Let NN be a sufficiently large positive rational. We define a function c~\tilde{c} by c~:=Nc+c\tilde{c}:=Nc+c^{\prime}. We now show that the pair (V,c~)(V,\tilde{c}) satisfies (5.2). We first observe that τc~(γ)\tau_{\tilde{c}}(\gamma) is not the optimal or nearly optimal value if γ(zi)I(si1,si+1)I(si+1,si1)\gamma(z_{i})\notin I(s_{i-1},s_{i+1})\cap I(s_{i+1},s_{i-1}) for some ii. Consider the case when γ(zi)I(si1,si+1)I(si+1,si1)\gamma(z_{i})\in I(s_{i-1},s_{i+1})\cap I(s_{i+1},s_{i-1}) holds for each ii. We show the following claim:

Claim 1.

Let xI(si1,si+1)I(si+1,si1)x\in I(s_{i-1},s_{i+1})\cap I(s_{i+1},s_{i-1}). Then at least one of the following conditions holds:

  1. (i)

    Both of sequences (si,si1,x)(s_{i},s_{i-1},x) and (x,si1,si)(x,s_{i-1},s_{i}) are μ\mu-shortest.

  2. (ii)

    Both of sequences (si,si+1,x)(s_{i},s_{i+1},x) and (x,si+1,si)(x,s_{i+1},s_{i}) are μ\mu-shortest.

Proof.

Suppose that (ii) does not hold. Let i=1i=1. By the assumption, we have μ(s1,x)<μ(s1,s2)+μ(s2,x)\mu(s_{1},x)<\mu(s_{1},s_{2})+\mu(s_{2},x) or μ(x,s1)<μ(x,s2)+μ(s2,s1)\mu(x,s_{1})<\mu(x,s_{2})+\mu(s_{2},s_{1}). Then we have Δ(s0,s1,x)<Δ(s0,s1,s2)\Delta(s_{0},s_{1},x)<\Delta(s_{0},s_{1},s_{2}). Hence, there exists a median mm of s0,s1,xs_{0},s_{1},x. If m=s0m=s_{0}, then (i) holds. If ms0m\neq s_{0}, then we have

Δ(s1,m,s2)\displaystyle\Delta(s_{1},m,s_{2}) =μ(s1,s2)+μ(s2,s1)+μ(s1,m)+μ(m,s1)+μ(s2,m)+μ(m,s2)\displaystyle=\mu(s_{1},s_{2})+\mu(s_{2},s_{1})+\mu(s_{1},m)+\mu(m,s_{1})+\mu(s_{2},m)+\mu(m,s_{2})
<μ(s1,s2)+μ(s2,s1)+μ(s1,s0)+μ(s0,s1)+μ(s2,x)+μ(x,m)+μ(m,x)+μ(x,s2)\displaystyle<\mu(s_{1},s_{2})+\mu(s_{2},s_{1})+\mu(s_{1},s_{0})+\mu(s_{0},s_{1})+\mu(s_{2},x)+\mu(x,m)+\mu(m,x)+\mu(x,s_{2})
<Δ(s0,s1,s2).\displaystyle<\Delta(s_{0},s_{1},s_{2}). (5.11)

Hence, there exists a median ww of s1,m,s2s_{1},m,s_{2}. However, ww is also a median of s0,s1,s2s_{0},s_{1},s_{2}, and this is a contradiction. ∎

For each i{0,1,,5}i\in\{0,1,\ldots,5\}, let gig_{i} be the contribution to the value τc~(γ)\tau_{\tilde{c}}(\gamma) from c(zi,s0)c^{\prime}(z_{i},s_{0}), c(zi,s1)c^{\prime}(z_{i},s_{1}), c(zi,s2)c^{\prime}(z_{i},s_{2}), c(s0,zi)c^{\prime}(s_{0},z_{i}), c(s1,zi)c^{\prime}(s_{1},z_{i}), c(s2,zi)c^{\prime}(s_{2},z_{i}). If γ(zi)=s0\gamma(z_{i})=s_{0}, then we have gi=a1μ2+a2μ1=(μ0+μ2μ1)/μ0+(μ1+μ0μ2)/μ0=2g_{i}=a_{1}\mu_{2}+a_{2}\mu_{1}=(\mu_{0}+\mu_{2}-\mu_{1})/\mu_{0}+(\mu_{1}+\mu_{0}-\mu_{2})/\mu_{0}=2. Similarly, we have gi=2g_{i}=2 when γ(zi)=s1\gamma(z_{i})=s_{1} or s2s_{2}. We next consider the case when γ(zi)I(si1,si+1)I(si+1,si1){si1,si+1}\gamma(z_{i})\in I(s_{i-1},s_{i+1})\cap I(s_{i+1},s_{i-1})\setminus\{s_{i-1},s_{i+1}\}. Let i=0i=0 and ϵ:=μ(s1,γ(z0))+μ(γ(z0),s1)\epsilon:=\mu(s_{1},\gamma(z_{0}))+\mu(\gamma(z_{0}),s_{1}). By Claim 1, we may assume that μ(s0,γ(zo))+μ(γ(z0),s0)=μ2+ϵ\mu(s_{0},\gamma(z_{o}))+\mu(\gamma(z_{0}),s_{0})=\mu_{2}+\epsilon holds. Hence, we have

g0\displaystyle g_{0} =a0(μ2+ϵ)+a1ϵ+a2(μ0ϵ)\displaystyle=a_{0}(\mu_{2}+\epsilon)+a_{1}\epsilon+a_{2}(\mu_{0}-\epsilon)
=a0μ2+a2μ0+ϵ(a0+a1a2)\displaystyle=a_{0}\mu_{2}+a_{2}\mu_{0}+\epsilon(a_{0}+a_{1}-a_{2})
=2+ϵ(a0+a1a2).\displaystyle=2+\epsilon(a_{0}+a_{1}-a_{2}). (5.12)

Note that we have

μ0μ1μ2(a0+a1a2)\displaystyle\mu_{0}\mu_{1}\mu_{2}(a_{0}+a_{1}-a_{2}) =μ0(μ1+μ2μ0)+μ1(μ0+μ2μ1)μ2(μ0+μ1μ2)\displaystyle=\mu_{0}(\mu_{1}+\mu_{2}-\mu_{0})+\mu_{1}(\mu_{0}+\mu_{2}-\mu_{1})-\mu_{2}(\mu_{0}+\mu_{1}-\mu_{2})
=2μ0μ1μ12μ02+μ22\displaystyle=2\mu_{0}\mu_{1}-\mu_{1}^{2}-\mu_{0}^{2}+\mu_{2}^{2}
=μ22(μ0μ1)2>0.\displaystyle=\mu_{2}^{2}-(\mu_{0}-\mu_{1})^{2}>0. (5.13)

Hence, we have g0>2g_{0}>2. Similarly, for each i{0,1,,5}i\in\{0,1,\ldots,5\}, we have gi>2g_{i}>2 if γ(zi)I(si1,si+1)I(si+1,si1){si1,si+1}\gamma(z_{i})\in I(s_{i-1},s_{i+1})\cap I(s_{i+1},s_{i-1})\setminus\{s_{i-1},s_{i+1}\}. Hence, the gadget (V,c~)(V,\tilde{c}) satisfies (5.2).

5.2.2 Proof of Theorem 5.1 for the case (ii)

Since μ\mu is modular and not directed orbit-invariant, HμH_{\mu} contains a directed orbit-varying modular cycle by Lemma 3.4. Let (s0,s1,s2,s3,s0)(s_{0},s_{1},s_{2},s_{3},s_{0}) be a directed orbit-varying modular cycle in HμH_{\mu}. Take eight elements x0,x1,y0,y1,z0,z1,w0,w1x_{0},x_{1},y_{0},y_{1},z_{0},z_{1},w_{0},w_{1}, and let V:=T{x0,x1,y0,y1,z0,z1,w0,w1}V:=T\cup\{x_{0},x_{1},y_{0},y_{1},z_{0},z_{1},w_{0},w_{1}\}. Without loss of generality, we may assume that μ(s0,s1)μ(s3,s2)=μ(s2,s3)μ(s1,s0)=μ(s0,s3)μ(s1,s2)=μ(s2,s1)μ(s3,s0)=k>0\mu(s_{0},s_{1})-\mu(s_{3},s_{2})=\mu(s_{2},s_{3})-\mu(s_{1},s_{0})=\mu(s_{0},s_{3})-\mu(s_{1},s_{2})=\mu(s_{2},s_{1})-\mu(s_{3},s_{0})=k>0. We define a function c4:V×V+c_{4}:V\times V\rightarrow\mathbb{Q}_{+} as follows (also see Figure 6):

c4(s2,xi)=c4(xi,s2)=c4(s3,xi)=c4(xi,s3)=1(i=0,1),\displaystyle c_{4}(s_{2},x_{i})=c_{4}(x_{i},s_{2})=c_{4}(s_{3},x_{i})=c_{4}(x_{i},s_{3})=1\mathrm{\ \ \ }(i=0,1),
c4(s0,zi)=c4(zi,s0)=c4(s1,zi)=c4(zi,s1)=1(i=0,1),\displaystyle c_{4}(s_{0},z_{i})=c_{4}(z_{i},s_{0})=c_{4}(s_{1},z_{i})=c_{4}(z_{i},s_{1})=1\mathrm{\ \ \ }(i=0,1),
c4(sj,yi)=c4(yi,sj)=1(i=0,1,j=0,1,2,3),\displaystyle c_{4}(s_{j},y_{i})=c_{4}(y_{i},s_{j})=1\mathrm{\ \ \ }(i=0,1,\ j=0,1,2,3),
c4(s1,w0)=c4(w0,s1)=c4(s2,w0)=c4(w0,s2)=1,\displaystyle c_{4}(s_{1},w_{0})=c_{4}(w_{0},s_{1})=c_{4}(s_{2},w_{0})=c_{4}(w_{0},s_{2})=1,
c4(s0,w1)=c4(w1,s0)=c4(s3,w1)=c4(w1,s3)=1.\displaystyle c_{4}(s_{0},w_{1})=c_{4}(w_{1},s_{0})=c_{4}(s_{3},w_{1})=c_{4}(w_{1},s_{3})=1. (5.14)
\begin{overpic}[width=398.33858pt]{nphfig2.eps} \put(31.4,27.0){$s_{0}$} \put(62.0,27.0){$s_{1}$} \put(62.0,5.0){$s_{2}$} \put(31.4,5.0){$s_{3}$} \put(47.0,9.5){$x_{0}$} \put(47.0,1.0){$x_{1}$} \put(47.0,31.1){$z_{0}$} \put(47.0,22.0){$z_{1}$} \put(65.5,16.5){$w_{0}$} \put(27.0,16.5){$w_{1}$} \put(50.0,16.5){$y_{i}$} \end{overpic}
Figure 6: the function c4c_{4}

Let μ10:=μ(s1,s0),μ12:=μ(s1,s2),μ30:=μ(s3,s0),\mu_{10}:=\mu(s_{1},s_{0}),\ \mu_{12}:=\mu(s_{1},s_{2}),\ \mu_{30}:=\mu(s_{3},s_{0}), and μ32:=μ(s3,s2)\mu_{32}:=\mu(s_{3},s_{2}). We next define a function c3:V×V+c_{3}:V\times V\rightarrow\mathbb{Q}_{+} as follows:

c3(y0,s0)=k2+μ32k+μ32μ12,\displaystyle c_{3}(y_{0},s_{0})=k^{2}+\mu_{32}k+\mu_{32}\mu_{12},
c3(y0,s1)=(μ10+μ12)k+μ10μ12,\displaystyle c_{3}(y_{0},s_{1})=(\mu_{10}+\mu_{12})k+\mu_{10}\mu_{12},
c3(y0,s2)=k2+μ10k+μ10μ30,\displaystyle c_{3}(y_{0},s_{2})=k^{2}+\mu_{10}k+\mu_{10}\mu_{30},
c3(y0,s3)=(μ32+μ30)k+μ32μ30,\displaystyle c_{3}(y_{0},s_{3})=(\mu_{32}+\mu_{30})k+\mu_{32}\mu_{30},
c3(s0,y1)=(μ10+μ30)k+μ10μ30,\displaystyle c_{3}(s_{0},y_{1})=(\mu_{10}+\mu_{30})k+\mu_{10}\mu_{30},
c3(s1,y1)=k2+μ32k+μ32μ30,\displaystyle c_{3}(s_{1},y_{1})=k^{2}+\mu_{32}k+\mu_{32}\mu_{30},
c3(s2,y1)=(μ32+μ12)k+μ32μ12,\displaystyle c_{3}(s_{2},y_{1})=(\mu_{32}+\mu_{12})k+\mu_{32}\mu_{12},
c3(s3,y1)=k2+μ10k+μ10μ12.\displaystyle c_{3}(s_{3},y_{1})=k^{2}+\mu_{10}k+\mu_{10}\mu_{12}. (5.15)

Then we define a function c2:V×V+c_{2}:V\times V\rightarrow\mathbb{Q}_{+} as follows:

c2(xi,zi)=c2(zi,xi)=1(i=0,1).\displaystyle c_{2}(x_{i},z_{i})=c_{2}(z_{i},x_{i})=1\mathrm{\ \ \ }(i=0,1). (5.16)

Also we define a function c1:V×V+c_{1}:V\times V\rightarrow\mathbb{Q}_{+} as follows:

c1(xi,yi)=c1(yi,xi)=c1(zi,yi)=c1(yi,zi)=1(i=0,1).\displaystyle c_{1}(x_{i},y_{i})=c_{1}(y_{i},x_{i})=c_{1}(z_{i},y_{i})=c_{1}(y_{i},z_{i})=1\mathrm{\ \ \ }(i=0,1). (5.17)

Finally, we define a function c0:V×V+c_{0}:V\times V\rightarrow\mathbb{Q}_{+} as follows:

c0(yi,wj)=c0(wj,yi)=1(i=0,1,j=0,1).\displaystyle c_{0}(y_{i},w_{j})=c_{0}(w_{j},y_{i})=1\mathrm{\ \ \ }(i=0,1,\ j=0,1). (5.18)

Let N1N_{1} be a sufficiently large positive rational. In addition, let NiN_{i} be a sufficiently large positive rational with respect to Ni1N_{i-1} for i=2,3,4i=2,3,4. We define a function cc by c:=c0+N1c1+N2c2+N3c3+N4c4c:=c_{0}+N_{1}c_{1}+N_{2}c_{2}+N_{3}c_{3}+N_{4}c_{4}. We now show that the pair (V,c)(V,c) satisfies (5.1) with respect to s2,s3,x0,x1s_{2},s_{3},x_{0},x_{1}. Focusing on the contribution from N4c4N_{4}c_{4}, we see that γ\gamma is infeasible if γ(xi)s2,s3\gamma(x_{i})\neq s_{2},s_{3} for some ii. Similarly, γ\gamma is infeasible if γ(zi)s0,s1\gamma(z_{i})\neq s_{0},s_{1} holds for some ii, or γ(w0)s1,s2\gamma(w_{0})\neq s_{1},s_{2} holds, or γ(w1)s0,s3\gamma(w_{1})\neq s_{0},s_{3} holds. Hence, it suffices to consider the case when γ(xi){s2,s3}\gamma(x_{i})\in\{s_{2},s_{3}\} and γ(zi){s0,s1}\gamma(z_{i})\in\{s_{0},s_{1}\} hold for each ii, or the case when γ(w0){s1,s2}\gamma(w_{0})\in\{s_{1},s_{2}\} and γ(w1){s0,s3}\gamma(w_{1})\in\{s_{0},s_{3}\} hold. In addition, γ\gamma is infeasible if γ(yi)I(s0,s2)I(s2,s0)I(s1,s3)I(s3,s1)\gamma(y_{i})\notin I(s_{0},s_{2})\cap I(s_{2},s_{0})\cap I(s_{1},s_{3})\cap I(s_{3},s_{1}). Note that HμH_{\mu} is modular by Lemma 3.2, which implies that edges {s0,s2}\{s_{0},s_{2}\} and {s1,s3}\{s_{1},s_{3}\} are not contained in HμH_{\mu} due to the condition of Lemma 2.5 (i). If γ(yi)I(s0,s2)I(s2,s0)I(s1,s3)I(s3,s1)\gamma(y_{i})\in I(s_{0},s_{2})\cap I(s_{2},s_{0})\cap I(s_{1},s_{3})\cap I(s_{3},s_{1}) and γ(yi){s0,s1,s2,s3}\gamma(y_{i})\notin\{s_{0},s_{1},s_{2},s_{3}\}, then by Lemma 3.1 (2), HμH_{\mu} has edges {γ(yi),sj}\{\gamma(y_{i}),s_{j}\} for j=0,1,2,3j=0,1,2,3. However, this contradicts the condition of Lemma 2.5 (i). Therefore, it suffices to consider the case when γ(yi){s0,s1,s2,s3}\gamma(y_{i})\in\{s_{0},s_{1},s_{2},s_{3}\}. We next focus on the contribution from N3c3N_{3}c_{3}. Let N3ξ0N_{3}\xi_{0} be the contribution to the value τc(γ)\tau_{c}(\gamma) from N3c3(y0,s0)N_{3}c_{3}(y_{0},s_{0}), N3c3(y0,s1)N_{3}c_{3}(y_{0},s_{1}), N3c3(y0,s2)N_{3}c_{3}(y_{0},s_{2}), N3c3(y0,s3)N_{3}c_{3}(y_{0},s_{3}). If γ(y0)=s0\gamma(y_{0})=s_{0}, then we have

ξ0\displaystyle\xi_{0} =(k2+μ10k+μ10μ30)(μ32+μ12+k)+((μ10+μ12)k+μ10μ12)(μ32+k)\displaystyle=(k^{2}+\mu_{10}k+\mu_{10}\mu_{30})(\mu_{32}+\mu_{12}+k)+((\mu_{10}+\mu_{12})k+\mu_{10}\mu_{12})(\mu_{32}+k)
+((μ32+μ30)k+μ32μ30)(μ12+k)\displaystyle+((\mu_{32}+\mu_{30})k+\mu_{32}\mu_{30})(\mu_{12}+k)
=k3+(2μ32+2μ10+2μ12+μ30)k2\displaystyle=k^{3}+(2\mu_{32}+2\mu_{10}+2\mu_{12}+\mu_{30})k^{2}
+(2μ32μ10+μ10μ30+μ32μ30+2μ10μ12+2μ32μ12+μ12μ30)k\displaystyle+(2\mu_{32}\mu_{10}+\mu_{10}\mu_{30}+\mu_{32}\mu_{30}+2\mu_{10}\mu_{12}+2\mu_{32}\mu_{12}+\mu_{12}\mu_{30})k
+μ32μ10μ12+μ32μ12μ30+μ32μ10μ30+μ10μ12μ30.\displaystyle+\mu_{32}\mu_{10}\mu_{12}+\mu_{32}\mu_{12}\mu_{30}+\mu_{32}\mu_{10}\mu_{30}+\mu_{10}\mu_{12}\mu_{30}. (5.19)

If γ(y0)=s1\gamma(y_{0})=s_{1}, then we have

ξ0\displaystyle\xi_{0} =((μ32+μ30)k+μ32μ30)(μ10+μ12+k)+(k2+μ32k+μ32μ12)μ10+(k2+μ10k+μ10μ30)μ12\displaystyle=((\mu_{32}+\mu_{30})k+\mu_{32}\mu_{30})(\mu_{10}+\mu_{12}+k)+(k^{2}+\mu_{32}k+\mu_{32}\mu_{12})\mu_{10}+(k^{2}+\mu_{10}k+\mu_{10}\mu_{30})\mu_{12}
=(μ32+μ10+μ12+μ30)k2+(2μ32μ10+μ10μ30+μ32μ30+μ10μ12+μ32μ12+μ12μ30)k\displaystyle=(\mu_{32}+\mu_{10}+\mu_{12}+\mu_{30})k^{2}+(2\mu_{32}\mu_{10}+\mu_{10}\mu_{30}+\mu_{32}\mu_{30}+\mu_{10}\mu_{12}+\mu_{32}\mu_{12}+\mu_{12}\mu_{30})k
+μ32μ10μ12+μ32μ12μ30+μ32μ10μ30+μ10μ12μ30.\displaystyle+\mu_{32}\mu_{10}\mu_{12}+\mu_{32}\mu_{12}\mu_{30}+\mu_{32}\mu_{10}\mu_{30}+\mu_{10}\mu_{12}\mu_{30}. (5.20)

If γ(y0)=s2\gamma(y_{0})=s_{2}, then we have

ξ0\displaystyle\xi_{0} =(k2+μ32k+μ32μ12)(μ10+μ30+k)+((μ32+μ30)k+μ32μ30)(μ10+k)\displaystyle=(k^{2}+\mu_{32}k+\mu_{32}\mu_{12})(\mu_{10}+\mu_{30}+k)+((\mu_{32}+\mu_{30})k+\mu_{32}\mu_{30})(\mu_{10}+k)
+((μ10+μ12)k+μ10μ12)(μ30+k)\displaystyle+((\mu_{10}+\mu_{12})k+\mu_{10}\mu_{12})(\mu_{30}+k)
=k3+(2μ32+2μ10+μ12+2μ30)k2\displaystyle=k^{3}+(2\mu_{32}+2\mu_{10}+\mu_{12}+2\mu_{30})k^{2}
+(2μ32μ10+2μ10μ30+2μ32μ30+μ10μ12+μ32μ12+μ12μ30)k\displaystyle+(2\mu_{32}\mu_{10}+2\mu_{10}\mu_{30}+2\mu_{32}\mu_{30}+\mu_{10}\mu_{12}+\mu_{32}\mu_{12}+\mu_{12}\mu_{30})k
+μ32μ10μ12+μ32μ12μ30+μ32μ10μ30+μ10μ12μ30.\displaystyle+\mu_{32}\mu_{10}\mu_{12}+\mu_{32}\mu_{12}\mu_{30}+\mu_{32}\mu_{10}\mu_{30}+\mu_{10}\mu_{12}\mu_{30}. (5.21)

If γ(y0)=s3\gamma(y_{0})=s_{3}, then we have

ξ0\displaystyle\xi_{0} =((μ10+μ12)k+μ10μ12)(μ32+μ30+k)+(k2+μ32k+μ32μ12)μ30+(k2+μ10k+μ10μ30)μ32\displaystyle=((\mu_{10}+\mu_{12})k+\mu_{10}\mu_{12})(\mu_{32}+\mu_{30}+k)+(k^{2}+\mu_{32}k+\mu_{32}\mu_{12})\mu_{30}+(k^{2}+\mu_{10}k+\mu_{10}\mu_{30})\mu_{32}
=(μ32+μ10+μ12+μ30)k2+(2μ32μ10+μ10μ30+μ32μ30+μ10μ12+μ32μ12+μ12μ30)k\displaystyle=(\mu_{32}+\mu_{10}+\mu_{12}+\mu_{30})k^{2}+(2\mu_{32}\mu_{10}+\mu_{10}\mu_{30}+\mu_{32}\mu_{30}+\mu_{10}\mu_{12}+\mu_{32}\mu_{12}+\mu_{12}\mu_{30})k
+μ32μ10μ12+μ32μ12μ30+μ32μ10μ30+μ10μ12μ30.\displaystyle+\mu_{32}\mu_{10}\mu_{12}+\mu_{32}\mu_{12}\mu_{30}+\mu_{32}\mu_{10}\mu_{30}+\mu_{10}\mu_{12}\mu_{30}. (5.22)

Hence, we see that the value ξ0\xi_{0} in the case of γ(y0){s1,s3}\gamma(y_{0})\in\{s_{1},s_{3}\} is smaller than that in the case of γ(y0){s0,s2}\gamma(y_{0})\in\{s_{0},s_{2}\}. Therefore, it suffices to consider the case when γ(y0){s1,s3}\gamma(y_{0})\in\{s_{1},s_{3}\}. Similar to N3ξ0N_{3}\xi_{0}, let N3ξ1N_{3}\xi_{1} be the contribution to the value τc(γ)\tau_{c}(\gamma) from N3c3(s0,y1)N_{3}c_{3}(s_{0},y_{1}), N3c3(s1,y1)N_{3}c_{3}(s_{1},y_{1}), N3c3(s2,y1)N_{3}c_{3}(s_{2},y_{1}), N3c3(s3,y1)N_{3}c_{3}(s_{3},y_{1}). If γ(y1)=s0\gamma(y_{1})=s_{0}, then we have

ξ1\displaystyle\xi_{1} =((μ32+μ12)k+μ32μ12)(μ10+μ30+k)+(k2+μ32k+μ32μ30)μ10+(k2+μ10k+μ10μ12)μ30\displaystyle=((\mu_{32}+\mu_{12})k+\mu_{32}\mu_{12})(\mu_{10}+\mu_{30}+k)+(k^{2}+\mu_{32}k+\mu_{32}\mu_{30})\mu_{10}+(k^{2}+\mu_{10}k+\mu_{10}\mu_{12})\mu_{30}
=(μ32+μ10+μ12+μ30)k2+(2μ32μ10+μ10μ30+μ32μ30+μ10μ12+μ32μ12+μ12μ30)k\displaystyle=(\mu_{32}+\mu_{10}+\mu_{12}+\mu_{30})k^{2}+(2\mu_{32}\mu_{10}+\mu_{10}\mu_{30}+\mu_{32}\mu_{30}+\mu_{10}\mu_{12}+\mu_{32}\mu_{12}+\mu_{12}\mu_{30})k
+μ32μ10μ12+μ32μ12μ30+μ32μ10μ30+μ10μ12μ30.\displaystyle+\mu_{32}\mu_{10}\mu_{12}+\mu_{32}\mu_{12}\mu_{30}+\mu_{32}\mu_{10}\mu_{30}+\mu_{10}\mu_{12}\mu_{30}. (5.23)

If γ(y1)=s1\gamma(y_{1})=s_{1}, then we have

ξ1\displaystyle\xi_{1} =(k2+μ10k+μ10μ12)(μ32+μ30+k)+((μ32+μ12)k+μ32μ12)(μ30+k)\displaystyle=(k^{2}+\mu_{10}k+\mu_{10}\mu_{12})(\mu_{32}+\mu_{30}+k)+((\mu_{32}+\mu_{12})k+\mu_{32}\mu_{12})(\mu_{30}+k)
+((μ10+μ30)k+μ10μ30)(μ32+k)\displaystyle+((\mu_{10}+\mu_{30})k+\mu_{10}\mu_{30})(\mu_{32}+k)
=k3+(2μ32+2μ10+μ12+2μ30)k2\displaystyle=k^{3}+(2\mu_{32}+2\mu_{10}+\mu_{12}+2\mu_{30})k^{2}
+(2μ32μ10+2μ10μ30+2μ32μ30+μ10μ12+μ32μ12+μ12μ30)k\displaystyle+(2\mu_{32}\mu_{10}+2\mu_{10}\mu_{30}+2\mu_{32}\mu_{30}+\mu_{10}\mu_{12}+\mu_{32}\mu_{12}+\mu_{12}\mu_{30})k
+μ32μ10μ12+μ32μ12μ30+μ32μ10μ30+μ10μ12μ30.\displaystyle+\mu_{32}\mu_{10}\mu_{12}+\mu_{32}\mu_{12}\mu_{30}+\mu_{32}\mu_{10}\mu_{30}+\mu_{10}\mu_{12}\mu_{30}. (5.24)

If γ(y1)=s2\gamma(y_{1})=s_{2}, then we have

ξ1\displaystyle\xi_{1} =((μ10+μ30)k+μ10μ30)(μ32+μ12k)+(k2+μ10k+μ10μ12)μ32+(k2+μ32k+μ32μ30)μ12\displaystyle=((\mu_{10}+\mu_{30})k+\mu_{10}\mu_{30})(\mu_{32}+\mu_{12}k)+(k^{2}+\mu_{10}k+\mu_{10}\mu_{12})\mu_{32}+(k^{2}+\mu_{32}k+\mu_{32}\mu_{30})\mu_{12}
=(μ32+μ10+μ12+μ30)k2+(2μ32μ10+μ10μ30+μ32μ30+μ10μ12+μ32μ12+μ12μ30)k\displaystyle=(\mu_{32}+\mu_{10}+\mu_{12}+\mu_{30})k^{2}+(2\mu_{32}\mu_{10}+\mu_{10}\mu_{30}+\mu_{32}\mu_{30}+\mu_{10}\mu_{12}+\mu_{32}\mu_{12}+\mu_{12}\mu_{30})k
+μ32μ10μ12+μ32μ12μ30+μ32μ10μ30+μ10μ12μ30.\displaystyle+\mu_{32}\mu_{10}\mu_{12}+\mu_{32}\mu_{12}\mu_{30}+\mu_{32}\mu_{10}\mu_{30}+\mu_{10}\mu_{12}\mu_{30}. (5.25)

If γ(y1)=s3\gamma(y_{1})=s_{3}, then we have

ξ1\displaystyle\xi_{1} =(k2+μ32k+μ32μ30)(μ10+μ12+k)+((μ32+μ12)k+μ32μ12)(μ10+k)\displaystyle=(k^{2}+\mu_{32}k+\mu_{32}\mu_{30})(\mu_{10}+\mu_{12}+k)+((\mu_{32}+\mu_{12})k+\mu_{32}\mu_{12})(\mu_{10}+k)
+((μ10+μ30)k+μ10μ30)(μ12+k)\displaystyle+((\mu_{10}+\mu_{30})k+\mu_{10}\mu_{30})(\mu_{12}+k)
=k3+(2μ32+2μ10+2μ12+μ30)k2\displaystyle=k^{3}+(2\mu_{32}+2\mu_{10}+2\mu_{12}+\mu_{30})k^{2}
+(2μ32μ10+μ10μ30+μ32μ30+2μ10μ12+2μ32μ12+μ12μ30)k\displaystyle+(2\mu_{32}\mu_{10}+\mu_{10}\mu_{30}+\mu_{32}\mu_{30}+2\mu_{10}\mu_{12}+2\mu_{32}\mu_{12}+\mu_{12}\mu_{30})k
+μ32μ10μ12+μ32μ12μ30+μ32μ10μ30+μ10μ12μ30.\displaystyle+\mu_{32}\mu_{10}\mu_{12}+\mu_{32}\mu_{12}\mu_{30}+\mu_{32}\mu_{10}\mu_{30}+\mu_{10}\mu_{12}\mu_{30}. (5.26)

Hence, we see that the value ξ1\xi_{1} in the case of γ(y1){s0,s2}\gamma(y_{1})\in\{s_{0},s_{2}\} is smaller than that in the case of γ(y1){s1,s3}\gamma(y_{1})\in\{s_{1},s_{3}\}. Therefore, it suffices to consider the case when γ(y1){s0,s2}\gamma(y_{1})\in\{s_{0},s_{2}\}. Recall that we consider the case when γ(xi){s2,s3}\gamma(x_{i})\in\{s_{2},s_{3}\} and γ(zi){s0,s1}\gamma(z_{i})\in\{s_{0},s_{1}\}. Focus on the contribution from N2c2N_{2}c_{2}. Then we see that the value τc(γ)\tau_{c}(\gamma) is not optimal or nearly optimal if γ(xi)=s2\gamma(x_{i})=s_{2} and γ(zi)=s0\gamma(z_{i})=s_{0}, or if γ(xi)=s3\gamma(x_{i})=s_{3} and γ(zi)=s1\gamma(z_{i})=s_{1}. Hence, it suffices to consider the case when γ(xi)=s2\gamma(x_{i})=s_{2} and γ(zi)=s1\gamma(z_{i})=s_{1} hold, or the case when γ(xi)=s3\gamma(x_{i})=s_{3} and γ(zi)=s0\gamma(z_{i})=s_{0} hold. We next focus on the contribution from N1c1N_{1}c_{1}. Then we see that a map γ\gamma is infeasible if γ(x0)=s2,γ(z0)=s1,\gamma(x_{0})=s_{2},\ \gamma(z_{0})=s_{1}, and γ(y0)=s3\gamma(y_{0})=s_{3} hold. Similarly, we see that a map γ\gamma is infeasible if γ(x0)=s3,γ(z0)=s0,\gamma(x_{0})=s_{3},\ \gamma(z_{0})=s_{0}, and γ(y0)=s1\gamma(y_{0})=s_{1} hold. Hence, we only consider the case when γ(x0)=s2,γ(z0)=s1,\gamma(x_{0})=s_{2},\ \gamma(z_{0})=s_{1}, and γ(y0)=s1\gamma(y_{0})=s_{1} hold, or the case when γ(x0)=s3,γ(z0)=s0,\gamma(x_{0})=s_{3},\ \gamma(z_{0})=s_{0}, and γ(y0)=s3\gamma(y_{0})=s_{3} hold. Applying the similar argument to x1x_{1}, z1z_{1}, and y1y_{1}, we see that it suffices to consider the case when γ(x1)=s2,γ(z1)=s1,\gamma(x_{1})=s_{2},\ \gamma(z_{1})=s_{1}, and γ(y1)=s2\gamma(y_{1})=s_{2} hold, or the case when γ(x1)=s3,γ(z1)=s0,\gamma(x_{1})=s_{3},\ \gamma(z_{1})=s_{0}, and γ(y1)=s0\gamma(y_{1})=s_{0} hold. We finally focus on the contribution from c0c_{0}. Let σ\sigma be the contribution to τc(γ)\tau_{c}(\gamma) from c0c_{0}. Consider the case when γ(x0)=γ(x1)=s2\gamma(x_{0})=\gamma(x_{1})=s_{2}, γ(y0)=s1\gamma(y_{0})=s_{1}, γ(y1)=s2\gamma(y_{1})=s_{2}. Then we have σ=2(μ32+μ10+μ12+μ30+2k)\sigma=2(\mu_{32}+\mu_{10}+\mu_{12}+\mu_{30}+2k) for any γ(w0){s1,s2},γ(w1){s0,s3}\gamma(w_{0})\in\{s_{1},s_{2}\},\gamma(w_{1})\in\{s_{0},s_{3}\}. Similarly, if γ(x0)=γ(x1)=s3\gamma(x_{0})=\gamma(x_{1})=s_{3}, γ(y0)=s3\gamma(y_{0})=s_{3}, γ(y1)=s0\gamma(y_{1})=s_{0}, then we have σ=2(μ32+μ10+μ12+μ30+2k)\sigma=2(\mu_{32}+\mu_{10}+\mu_{12}+\mu_{30}+2k) for any γ(w0){s1,s2},γ(w1){s0,s3}\gamma(w_{0})\in\{s_{1},s_{2}\},\gamma(w_{1})\in\{s_{0},s_{3}\}. On the other hand, if γ(x0)=s2\gamma(x_{0})=s_{2}, γ(x1)=s3\gamma(x_{1})=s_{3}, γ(y0)=s1\gamma(y_{0})=s_{1}, γ(y1)=s0\gamma(y_{1})=s_{0}, then σ\sigma takes the minimum value σ=2(μ32+μ10+k)\sigma=2(\mu_{32}+\mu_{10}+k) when γ(w0)=s1\gamma(w_{0})=s_{1} and γ(w1)=s0\gamma(w_{1})=s_{0}. Similarly, if γ(x0)=s3\gamma(x_{0})=s_{3}, γ(x1)=s2\gamma(x_{1})=s_{2}, γ(y0)=s3\gamma(y_{0})=s_{3}, γ(y1)=s2\gamma(y_{1})=s_{2}, then σ\sigma takes the minimum value σ=2(μ32+μ10+k)\sigma=2(\mu_{32}+\mu_{10}+k) when γ(w0)=s2\gamma(w_{0})=s_{2} and γ(w1)=s3\gamma(w_{1})=s_{3}. Thus, the pair (V,c)(V,c) satisfies the condition (5.1).

5.2.3 Proof of Theorem 5.1 for the case (iii)

We extend the proof of Theorem 1.1 for the case (ii) in [14] to that of Theorem 5.1 for the case (iii). Since the underlying graph HμH_{\mu} is not orientable, there exists a sequence (e0,e1,,ek)(\overrightarrow{e_{0}},\overrightarrow{e_{1}},\ldots,\overrightarrow{e_{k}}) (ei=(si,ti)\overrightarrow{e_{i}}=(s_{i},t_{i}) is an oriented edge of HμH_{\mu} for each ii) such that HμH_{\mu} contains a 4-cycle (si,ti,ti+1,si+1,si)(s_{i},t_{i},t_{i+1},s_{i+1},s_{i}) for each i{0,,k1}i\in\{0,\ldots,k-1\}, and tk=s0,sk=t0t_{k}=s_{0},s_{k}=t_{0}. Then we have h:=μ(s0,t0)=μ(s1,t1)==μ(sk,tk)=μ(t0,s0)=μ(t1,s1)==μ(tk,sk)h:=\mu(s_{0},t_{0})=\mu(s_{1},t_{1})=\cdots=\mu(s_{k},t_{k})=\mu(t_{0},s_{0})=\mu(t_{1},s_{1})=\cdots=\mu(t_{k},s_{k}), and fi:=μ(si,si+1)=μ(ti,ti+1)f_{i}:=\mu(s_{i},s_{i+1})=\mu(t_{i},t_{i+1}), gi:=μ(si+1,si)=μ(ti+1,ti)g_{i}:=\mu(s_{i+1},s_{i})=\mu(t_{i+1},t_{i}) for i=0,,k1i=0,\ldots,k-1, since μ\mu is directed orbit-invariant. Take 2k2k elements z0,z1,,z2k1z_{0},z_{1},\ldots,z_{2k-1}, and let V:=T{z0,z1,,z2k1}V:=T\cup\{z_{0},z_{1},\ldots,z_{2k-1}\}. We now define a function c:V×V+c:V\times V\rightarrow\mathbb{Q}_{+} as follows:

c(zi,si)=c(si,zi)=c(zi,ti)=c(ti,zi)=1(i=0,1,,2k1),\displaystyle c(z_{i},s_{i})=c(s_{i},z_{i})=c(z_{i},t_{i})=c(t_{i},z_{i})=1\mathrm{\ \ \ }(i=0,1,\ldots,2k-1), (5.27)

where the indices of sis_{i} and tit_{i} are taken modulo kk. We also define a function c:V×V+c^{\prime}:V\times V\rightarrow\mathbb{Q}_{+} as follows (see Figure 7):

c(zi,zi+1)=c(zi+1,zi)=1(i=0,1,,2k1),\displaystyle c^{\prime}(z_{i},z_{i+1})=c^{\prime}(z_{i+1},z_{i})=1\mathrm{\ \ \ }(i=0,1,\ldots,2k-1), (5.28)

where the indices of ziz_{i} are taken modulo 2k2k.

\begin{overpic}[width=398.33858pt]{nphfig3.eps} \put(15.5,28.5){$s_{0}$} \put(36.0,28.5){$s_{1}$} \put(58.0,28.5){$s_{2}$} \put(80.0,28.5){$s_{3}$} \put(15.5,4.2){$t_{0}$} \put(36.0,4.2){$t_{1}$} \put(58.0,4.2){$t_{2}$} \put(80.0,4.2){$t_{3}$} \put(21.7,22.0){$z_{0}$} \put(42.0,22.0){$z_{1}$} \put(64.3,22.0){$z_{2}$} \put(86.0,22.0){$z_{3}$} \put(10.0,10.5){$z_{4}$} \put(30.0,10.5){$z_{5}$} \put(53.0,10.5){$z_{6}$} \put(75.0,10.5){$z_{7}$} \put(95.0,19.6){to $z_{4}$} \put(86.8,12.5){to $z_{0}$} \end{overpic}
Figure 7: the function cc^{\prime} (the case of k=4k=4)

Let NN be a sufficiently large positive rational. We define a function c~\tilde{c} by c~:=Nc+c\tilde{c}:=Nc+c^{\prime}. Then we show that the pair (V,c~)(V,\tilde{c}) satisfies (5.1) with respect to s0,t0,z0,zks_{0},t_{0},z_{0},z_{k}. Let si+k:=ti(i=0,,k1)s_{i+k}:=t_{i}\ (i=0,\ldots,k-1), and the indices of sis_{i} are taken modulo 2k2k below. We first observe that γ\gamma is infeasible if γ(zi)si,si+k\gamma(z_{i})\neq s_{i},s_{i+k} for some ii, due to the contribution from NcNc. Thus, it suffices to consider the case when γ(zi){si,si+k}\gamma(z_{i})\in\{s_{i},s_{i+k}\} for every i{0,,2k1}i\in\{0,\ldots,2k-1\}. Let σ\sigma be the contribution to τc~(γ)\tau_{\tilde{c}}(\gamma) from cc^{\prime}, and σi\sigma_{i} be the contribution from c(zi,zi+1)c^{\prime}(z_{i},z_{i+1}) and c(zi+1,zi)c^{\prime}(z_{i+1},z_{i}) for each ii. If (γ(zi),γ(zi+1))=(si,si+1)(\gamma(z_{i}),\gamma(z_{i+1}))=(s_{i},s_{i+1}) or (si+k,si+k+1)(s_{i+k},s_{i+k+1}), then we have σi=fi+gi\sigma_{i}=f_{i}+g_{i}. Otherwise, we have σi=fi+gi+2h\sigma_{i}=f_{i}+g_{i}+2h. Consider the case when γ(z0)=s0\gamma(z_{0})=s_{0} and γ(zk)=sk\gamma(z_{k})=s_{k}. In this case, if γ(zi)=si\gamma(z_{i})=s_{i} holds for i=0,,2k1i=0,\ldots,2k-1, then we have σi=fi+gi\sigma_{i}=f_{i}+g_{i} for each ii, and σ=2i=0k1(fi+gi)\sigma=2\sum_{i=0}^{k-1}(f_{i}+g_{i}). Similarly, in the case of γ(z0)=sk\gamma(z_{0})=s_{k} and γ(zk)=s0\gamma(z_{k})=s_{0}, we have σ=2i=0k1(fi+gi)\sigma=2\sum_{i=0}^{k-1}(f_{i}+g_{i}) if γ(zi)=si+k\gamma(z_{i})=s_{i+k} for i=0,,2k1i=0,\ldots,2k-1. Consider the case when γ(z0)=γ(zk)=s0\gamma(z_{0})=\gamma(z_{k})=s_{0}. In this case, there exist two or more integers i{0,,2k1}i\in\{0,\ldots,2k-1\} for any γ\gamma such that σi=fi+gi+2h\sigma_{i}=f_{i}+g_{i}+2h holds, and exactly two integers for some γ\gamma. Hence, the minimum value of σ\sigma is 2i=0k1(fi+gi)+4h2\sum_{i=0}^{k-1}(f_{i}+g_{i})+4h. In the case of γ(z0)=γ(zk)=sk\gamma(z_{0})=\gamma(z_{k})=s_{k}, we also see that the minimum value of σ\sigma is 2i=0k1(fi+gi)+4h2\sum_{i=0}^{k-1}(f_{i}+g_{i})+4h by the similar argument.

5.2.4 Proof of Theorem 1.5

Let (s0,s1,s2)(s_{0},s_{1},s_{2}) be a biased non-collinear triple in TT. For i=0,1,2i=0,1,2, since (si,si+1)(s_{i},s_{i+1}) is a biased pair, Rμ(si,x)>Rμ(x,si+1)R_{\mu}(s_{i},x)>R_{\mu}(x,s_{i+1}) holds for every xI(si,si+1)I(si+1,si){si,si+1}x\in I(s_{i},s_{i+1})\cap I(s_{i+1},s_{i})\setminus\{s_{i},s_{i+1}\}, or Rμ(si,x)<Rμ(x,si+1)R_{\mu}(s_{i},x)<R_{\mu}(x,s_{i+1}) holds for every xI(si,si+1)I(si+1,si){si,si+1}x\in I(s_{i},s_{i+1})\cap I(s_{i+1},s_{i})\setminus\{s_{i},s_{i+1}\}, where the indices of sis_{i} are taken modulo 3. Take six elements z0,z1,,z5z_{0},z_{1},\ldots,z_{5}, and let V:=T{z0,,z5}V:=T\cup\{z_{0},\ldots,z_{5}\}. We define a function c:V×V+c:V\times V\rightarrow\mathbb{Q}_{+} as follows:

c(si1,zi)=c(zi,si1)=c(si+1,zi)=c(zi,si+1)=1(i=0,,5).\displaystyle c(s_{i-1},z_{i})=c(z_{i},s_{i-1})=c(s_{i+1},z_{i})=c(z_{i},s_{i+1})=1\mathrm{\ \ \ }(i=0,\ldots,5). (5.29)

We next define a function c:V×V+c^{\prime}:V\times V\rightarrow\mathbb{Q}_{+} as follows. For each i{0,,5}i\in\{0,\ldots,5\},

  • if Rμ(si1,x)>Rμ(x,si+1)R_{\mu}(s_{i-1},x)>R_{\mu}(x,s_{i+1}) holds for every xI(si1,si+1)I(si+1,si1){si1,si+1}x\in I(s_{i-1},s_{i+1})\cap I(s_{i+1},s_{i-1})\setminus\{s_{i-1},s_{i+1}\}, then

    c(si1,zi)=μ(si+1,si1),\displaystyle c^{\prime}(s_{i-1},z_{i})=\mu(s_{i+1},s_{i-1}),
    c(si+1,zi)=μ(si1,si+1).\displaystyle c^{\prime}(s_{i+1},z_{i})=\mu(s_{i-1},s_{i+1}). (5.30)
  • if Rμ(si1,x)<Rμ(x,si+1)R_{\mu}(s_{i-1},x)<R_{\mu}(x,s_{i+1}) holds for every xI(si1,si+1)I(si+1,si1){si1,si+1}x\in I(s_{i-1},s_{i+1})\cap I(s_{i+1},s_{i-1})\setminus\{s_{i-1},s_{i+1}\}, then

    c(zi,si1)=μ(si1,si+1),\displaystyle c^{\prime}(z_{i},s_{i-1})=\mu(s_{i-1},s_{i+1}),
    c(zi,si+1)=μ(si+1,si1).\displaystyle c^{\prime}(z_{i},s_{i+1})=\mu(s_{i+1},s_{i-1}). (5.31)

Let NN be a sufficiently large positive rational. We define a function c~\tilde{c} by c~:=Nc+c\tilde{c}:=Nc+c^{\prime}. We now show that the pair (V,c~)(V,\tilde{c}) satisfies the condition (5.2) with respect to s0,s1,s2,z0,z1,z2,z3,z4,z5s_{0},s_{1},s_{2},z_{0},z_{1},z_{2},z_{3},z_{4},z_{5}. Focusing on the contribution from NcNc, we see that a map γ\gamma is infeasible if γ(zi)I(si1,si+1)I(si+1,si1)\gamma(z_{i})\notin I(s_{i-1},s_{i+1})\cap I(s_{i+1},s_{i-1}) holds for some ii. Thus, it suffices to consider the case when γ(zi)I(si1,si+1)I(si+1,si1)\gamma(z_{i})\in I(s_{i-1},s_{i+1})\cap I(s_{i+1},s_{i-1}) holds for each ii. We next focus on the contribution from cc^{\prime}. Consider the case when Rμ(si1,x)>Rμ(x,si+1)R_{\mu}(s_{i-1},x)>R_{\mu}(x,s_{i+1}) holds for every xI(si1,si+1)I(si+1,si1){si1,si+1}x\in I(s_{i-1},s_{i+1})\cap I(s_{i+1},s_{i-1})\setminus\{s_{i-1},s_{i+1}\}. Let σi\sigma_{i} be the contribution to τc~(γ)\tau_{\tilde{c}}(\gamma) from c(si1,zi)c^{\prime}(s_{i-1},z_{i}) and c(si+1,zi)c^{\prime}(s_{i+1},z_{i}). If γ(zi){si1,si+1}\gamma(z_{i})\in\{s_{i-1},s_{i+1}\}, then we have σi=μ(si1,si+1)μ(si+1,si1)\sigma_{i}=\mu(s_{i-1},s_{i+1})\mu(s_{i+1},s_{i-1}). If γ(zi)I(si1,si+1)I(si+1,si1){si1,si+1}\gamma(z_{i})\in I(s_{i-1},s_{i+1})\cap I(s_{i+1},s_{i-1})\setminus\{s_{i-1},s_{i+1}\}, then we have

σi\displaystyle\sigma_{i} =μ(si+1,si1)μ(si1,γ(zi))+μ(si1,si+1)μ(si+1,γ(zi))\displaystyle=\mu(s_{i+1},s_{i-1})\mu(s_{i-1},\gamma(z_{i}))+\mu(s_{i-1},s_{i+1})\mu(s_{i+1},\gamma(z_{i}))
>μ(si+1,si1)μ(si1,γ(zi))+μ(γ(zi),si+1)μ(si+1,si1)\displaystyle>\mu(s_{i+1},s_{i-1})\mu(s_{i-1},\gamma(z_{i}))+\mu(\gamma(z_{i}),s_{i+1})\mu(s_{i+1},s_{i-1})
μ(si1,si+1)μ(si+1,si1).\displaystyle\geq\mu(s_{i-1},s_{i+1})\mu(s_{i+1},s_{i-1}). (5.32)

Consider the case when Rμ(si1,x)<Rμ(x,si+1)R_{\mu}(s_{i-1},x)<R_{\mu}(x,s_{i+1}) holds for every xI(si1,si+1)I(si+1,si1){si1,si+1}x\in I(s_{i-1},s_{i+1})\cap I(s_{i+1},s_{i-1})\setminus\{s_{i-1},s_{i+1}\}. Let σi\sigma_{i}^{\prime} be the contribution to τc~(γ)\tau_{\tilde{c}}(\gamma) from c(zi,si1)c^{\prime}(z_{i},s_{i-1}) and c(zi,si+1)c^{\prime}(z_{i},s_{i+1}). Similar to the above case, we have σi=μ(si1,si+1)μ(si+1,si1)\sigma_{i}^{\prime}=\mu(s_{i-1},s_{i+1})\mu(s_{i+1},s_{i-1}) if γ(zi){si1,si+1}\gamma(z_{i})\in\{s_{i-1},s_{i+1}\}, and we have σi>μ(si1,si+1)μ(si+1,si1)\sigma_{i}^{\prime}>\mu(s_{i-1},s_{i+1})\mu(s_{i+1},s_{i-1}) if γ(zi)I(si1,si+1)I(si+1,si1){si1,si+1}\gamma(z_{i})\in I(s_{i-1},s_{i+1})\cap I(s_{i+1},s_{i-1})\setminus\{s_{i-1},s_{i+1}\}. Thus, (V,c~)(V,\tilde{c}) satisfies the condition (5.2). This completes the proof.

Acknowledgments

We thank the referees for helpful comments. The first author was supported by JSPS KAKENHI Grant Numbers JP17K00029 and JST PRESTO Grant Number JPMJPR192A, Japan.

References

  • [1] Hans-Jürgen Bandelt. Networks with condorcet solutions. European Journal of Operational Research, 20(3):314–326, 1985.
  • [2] Hans-Jürgen Bandelt, Marcel van de Vel, and Eric R. Verheul. Modular interval spaces. Mathematische Nachrichten, 163:177–201, 1993.
  • [3] Garrett Birkhoff. Lattice Theory. American Mathematical Society, third edition, 1967.
  • [4] Victor Chepoi. Classification of graphs by metric triangles. Metody Diskretnogo Analyza, 49:75–93, 1989.
  • [5] Victor Chepoi. A multifacility location problem on median spaces. Discrete Applied Mathematics, 64(1):1–29, 1996.
  • [6] David A. Cohen, Martin C. Cooper, Peter G. Jeavons, and Andrei A. Krokhin. The complexity of soft constraint satisfaction. Artificial Intelligence, 170:983–1016, 2006.
  • [7] Elias Dahlhaus, David S. Johnson, Christos H. Papadimitriou, Paul D. Seymour, and Mihalis Yannakakis. The complexity of multiterminal cuts. SIAM Journal on Computing, 23(4):864–894, 1994.
  • [8] Hiroshi Hirai. Tight spans of distances and the dual fractionality of undirected multiflow problems. Journal of Combinatorial Theory, Series B, 99(6):843–868, 2009.
  • [9] Hiroshi Hirai. Folder complexes and multiflow combinatorial dualities. SIAM Journal on Discrete Mathematics, 25(3):1119–1143, 2011.
  • [10] Hiroshi Hirai. Discrete convexity and polynomial solvability in minimum 0-extension problems. Mathematical Programming, 155(1-2):1–55, 2016.
  • [11] Hiroshi Hirai and Ryuhei Mizutani. Minimum 0-extension problems on directed metrics. In 45th International Symposium on Mathematical Foundations of Computer Science, MFCS 2020, volume 170 of LIPIcs, pages 46:1–46:13. Schloss Dagstuhl - Leibniz-Zentrum für Informatik, 2020.
  • [12] Alexander V. Karzanov. Metrics with finite sets of primitive extensions. Annals of Combinatorics, 2:213–243, 1998.
  • [13] Alexander V. Karzanov. Minimum 0-extensions of graph metrics. European Journal of Combinatrics, 19(1):71–101, 1998.
  • [14] Alexander V. Karzanov. Hard cases of the multifacility location problem. Discrete Applied Mathematics, 143(1-3):368–373, 2004.
  • [15] Alexander V. Karzanov. One more well-solved case of the multifacility location problem. Discrete Optimization, 1:51–66, 2004.
  • [16] Vladimir Kolmogorov, Johan Thapper, and Stanislav Živný. The power of linear programming for general-valued csps. SIAM Journal on Computing, 44(1):1–36, 2015.
  • [17] Barbaros C. Tansel, Richard L. Francis, and Timothy J. Lowe. Location on networks: A survey. part i, ii. Management Science, 29(4):482–511, 1983.
  • [18] Johan Thapper and Stanislav Živný. The power of linear programming for valued csps. In 53rd Annual IEEE Symposium on Foundations of Computer Science, FOCS 2012, pages 669–678. IEEE Computer Society, 2012.
  • [19] Johan Thapper and Stanislav Živný. The complexity of finite-valued csps. Journal of the ACM, 63(4):37:1–37:33, 2016.
  • [20] M.L.J. van de Vel. Theory of Convex Structures. North Holland, 1993.
  • [21] Stanislav Živný. The Complexity of Valued Constraint Satisfaction Problems. Springer, 2012.