Min-Cost Multicast Networks in Euclidean Space

Similar documents
Min-Cost Multicast Networks in Euclidean Space

Space Information Flow

2386 IEEE TRANSACTIONS ON INFORMATION THEORY, VOL. 52, NO. 6, JUNE 2006

The Encoding Complexity of Network Coding

On Benefits of Network Coding in Bidirected Networks and Hyper-networks

WHILE traditional information transmission in a data network. Bounding the Coding Advantage of Combination Network Coding in Undirected Networks

The Geometry of Carpentry and Joinery

The Encoding Complexity of Network Coding

A Connection between Network Coding and. Convolutional Codes

Treewidth and graph minors

K 4,4 e Has No Finite Planar Cover

Connected Components of Underlying Graphs of Halving Lines

Approximating Fault-Tolerant Steiner Subgraphs in Heterogeneous Wireless Networks

Lecture 15: The subspace topology, Closed sets

Network Coding in Planar Networks

12.1 Formulation of General Perfect Matching

The Full Survey on The Euclidean Steiner Tree Problem

Monotone Paths in Geometric Triangulations

Lecture 2 - Introduction to Polytopes

EXTREME POINTS AND AFFINE EQUIVALENCE

CPSC 536N: Randomized Algorithms Term 2. Lecture 10

AVALANCHE: A NETWORK CODING ANALYSIS

3 No-Wait Job Shops with Variable Processing Times

A Systematic Approach to Network Coding Problems using Conflict Graphs

The Geodesic Integral on Medial Graphs

Network Routing Capacity

On median graphs and median grid graphs

Notes 4 : Approximating Maximum Parsimony

Discrete mathematics

PACKING DIGRAPHS WITH DIRECTED CLOSED TRAILS

2 The Fractional Chromatic Gap

Matching Algorithms. Proof. If a bipartite graph has a perfect matching, then it is easy to see that the right hand side is a necessary condition.

Straight-line Drawability of Embedded Graphs

MOST attention in the literature of network codes has

arxiv: v5 [cs.dm] 9 May 2016

FUTURE communication networks are expected to support

FOUR EDGE-INDEPENDENT SPANNING TREES 1

Multicast Network Coded Flow in Grid Graphs

A Reduction of Conway s Thrackle Conjecture

7. The Gauss-Bonnet theorem

Problem Set 2 Solutions

On the Max Coloring Problem

Pebble Sets in Convex Polygons

The Multiple Unicast Network Coding Conjecture. and a geometric framework for studying it

V1.0: Seth Gilbert, V1.1: Steven Halim August 30, Abstract. d(e), and we assume that the distance function is non-negative (i.e., d(x, y) 0).

Network Coding in Planar Networks

The Structure of Bull-Free Perfect Graphs

Matching and Planarity

Line Arrangement. Chapter 6

arxiv: v2 [cs.cg] 24 Jul 2011

Approximation Algorithms: The Primal-Dual Method. My T. Thai

On Covering a Graph Optimally with Induced Subgraphs

Approximating Node-Weighted Multicast Trees in Wireless Ad-Hoc Networks

Ice-Creams and Wedge Graphs

Network Coding for Joint Storage and Transmission with Minimum Cost

Joint Entity Resolution

Computer Science 280 Fall 2002 Homework 10 Solutions

In this lecture, we ll look at applications of duality to three problems:

6. Lecture notes on matroid intersection

Connecting face hitting sets in planar graphs

Introduction to Graph Theory

CMSC 451: Lecture 22 Approximation Algorithms: Vertex Cover and TSP Tuesday, Dec 5, 2017

Geometric Steiner Trees

Chapter 12 and 11.1 Planar graphs, regular polyhedra, and graph colorings

Which n-venn diagrams can be drawn with convex k-gons?

Notes for Lecture 24

CSE 331: Introduction to Algorithm Analysis and Design Graphs

On Minimum Weight Pseudo-Triangulations

Lecture 5: Duality Theory

CSE 548: Analysis of Algorithms. Lecture 13 ( Approximation Algorithms )

Distributed minimum spanning tree problem

2. Optimization problems 6

Overlapping Cluster Planarity.

Mathematical and Algorithmic Foundations Linear Programming and Matchings

EULER S FORMULA AND THE FIVE COLOR THEOREM

Points covered an odd number of times by translates

Computational Geometry: Lecture 5

Assignment 4 Solutions of graph problems

Lecture 20 : Trees DRAFT

Cross-Monotonic Multicast

Counting the Number of Eulerian Orientations

γ 2 γ 3 γ 1 R 2 (b) a bounded Yin set (a) an unbounded Yin set

Hamiltonian cycles in bipartite quadrangulations on the torus

Notes on Binary Dumbbell Trees

Bottleneck Steiner Tree with Bounded Number of Steiner Vertices

The Multi-Sender Multicast Index Coding

On the Relationships between Zero Forcing Numbers and Certain Graph Coverings

Crossing Families. Abstract

Lecture 7. s.t. e = (u,v) E x u + x v 1 (2) v V x v 0 (3)

Lecture 25: Bezier Subdivision. And he took unto him all these, and divided them in the midst, and laid each piece one against another: Genesis 15:10

Consistency and Set Intersection

Planar graphs. Math Prof. Kindred - Lecture 16 Page 1

Tutte s Theorem: How to draw a graph

On the Multiple Unicast Network Coding Conjecture

Chapter 15 Introduction to Linear Programming

Some Open Problems in Graph Theory and Computational Geometry

Math 311. Trees Name: A Candel CSUN Math

Midpoint Routing algorithms for Delaunay Triangulations

LOW-DENSITY PARITY-CHECK (LDPC) codes [1] can

ADJACENCY POSETS OF PLANAR GRAPHS

Transcription:

Min-Cost Multicast Networks in Euclidean Space Xunrui Yin, Yan Wang, Xin Wang, Xiangyang Xue School of Computer Science Fudan University {09110240030,11110240029,xinw,xyxue}@fudan.edu.cn Zongpeng Li Dept. of Computer Science, University of Calgary and Institute of Network Coding, CUHK zongpeng@ucalgary.ca Abstract Space information flow is a new field of research recently proposed by Li and Wu [1], [2]. It studies the transmission of information in a geometric space, where information flows can be routed along any trajectories, and can be encoded wherever they meet. The goal is to satisfy given endto-end unicast/multicast throughput demands, while minimizing a natural bandwidth-distance sum-product (network volume). Space information flow models the design of a blueprint for a minimum-cost network. We study the multicast version of the space information flow problem, in Euclidean spaces. We present a simple example that demonstrates the design of an information network is indeed different from that of a transportation network. We discuss properties of optimal multicast network embedding, prove that network coding does not make a difference in the basic case of 1-to-2 multicast, and prove upper-bounds on the number of relay nodes required in an optimal multicast network in general. I. INTRODUCTION The classic Minimum Euclidean Steiner tree problem studies the shortest tree that connects a given set of nodes in an Euclidean space. It has a broad range of applications in operations research and communication networks. In particular, it has been believed that the design of a min-cost multicast network is equivalent to constructing a minimum Steiner tree. More specifically, given the positions of n nodes in an Euclidean space R d, the minimum Steiner tree problem is to find a min-cost network connecting the n nodes. Relay nodes can be inserted wherever desired, and a link can be drawn between any pair of nodes. An optimal solution always has a tree structure, and the cost of the tree is the total length of its edges/links. With the conventional store-andforward principle, a minimum multicast network connecting given terminals is always a minimum Euclidean Steiner tree. In a Steiner tree, every link has a unit capacity 1. With network coding, it is natural to weight the cost of a link by its capacity. The total cost of a multicast network is then its volume : the product of capacity and distance at each link, summed over all links in the network. Li and Wu [1] show that given network coding, a min-cost multicast network is no longer necessarily a Steiner tree. Below we explain this fact with a new and simpler example. Fig. 1 shows six terminals lying on a equilateral triangle. Nodes s 1, s 2, and s 3 are sources having the same information, which is requested by receivers t 1, t 2, and t 3 at unit rate 1 bps. A min-cost multicast network without network coding is shown in (a), which is a geometric embedding of a Steiner forest, with capacity 1 at each link. Edge length of the Fig. 1. An example where the optimal multicast network is not a minimum Steiner tree. Black nodes are terminals: s 1, s 2, and s 3 are sources with the same information, requested by receivers t 1, t 2, and t 3. White nodes are relay nodes inserted. Each of a, b, and hence a + b, is an information flow of rate 0.5. Thick solid edges have capacity 1, thinner solid edges have capacity 0.5. large equilateral triangle is 2. Fig. 1(a) shows the minimum multicast network based on network coding, with flow rate 0.5 at each link. The total cost of this multicast network is 3 3/2. Fig. 1(b) shows a minimum network based on a Steiner tree, for supporting the same multicast rate of 1 to the three receivers. Each link flow rate here is 1, the three angles at each relay are all 120. The total cost of the Steiner tree is 7. The ratio of the two costs is 7/(3 3/2) = 1.018. It is worth pointing out that, although the gap between tree based and network coding based multicast network costs is small, the gap reveals a fundamental difference in the underlying structure of tree based multicast and network coding based multicast in a geometric space. Consequently, these two problems may have drastically different computational complexity. Fig. 2. t 1 s 1 s s 2 s 3 t 2 Transform the multiple sources example into a single source example. While the above example includes multiple sources, it can be transformed into a single source example. As shown in Fig. 2, add a new source node s, connect s to each of s 1, s 2, and s 3 with a curved path where receivers are densely t 3

distributed. When these new receivers are close enough to each other on the three new paths, an optimal multicast network will connect them along the three curved paths. Consequently, from the point of view of t 1, t 2 and t 3, s 1, s 2, and s 3 become information sources holding common information of interest, and we are back at the scenario in Fig. 1. In this work, we concentrate on the new problem of min-cost multicast network design with network coding, in Euclidean spaces. As one of the first work along this direction, we study basic problems including (1) what properties must an optimal multicast network have, (2) when does an optimal multicast network require and not require network coding, and (3) the computational complexity of constructing the optimal multicast network. In particular, we prove that network coding is not necessary in the basic case of 1-to-2 multicast (three terminals, including one source and two receivers). We also prove upper-bounds on the number of relay nodes required in an optimal multicast network. The paper is organized as follows. In Sec. II, we present the formal formulation of the problem and develop an equivalent discrete model that is helpful in the following analysis. In Sec. III, we prove that network coding is unnecessary for the case of three terminals. In Sec. IV, we study the number of relay nodes required in an optimal multicast network. We conclude the paper and discuss open questions in Sec. V. II. PROBLEM MODEL AND PRELIMINARIES A. Problem Formulation Let T denote the set of n terminals and s T be the source. A network is specified by a finite directed graph D(V,A), T V, the capacity c( uv) of each link uv A, and the position ρ(v) of each relay node v V. A network is called a feasible multicast network if it can support a unit rate multicast session from s to each receiver t R = T \{s}. The cost of link uv is given by the product of its capacity c( uv) and its length uv, i.e., the Euclidean distance between u and v. The cost of introducing a relay node is ignored. Hence the cost of a network is the summation of its link costs. The multicast version of the space information flow problem, or the design of a min-cost multicast network in a geometric space, is to construct a feasible multicast network of minimum cost. With network coding, feasibility is checked by verifying whether the network can support a multicast flow of unit rate. A celebrated result on network coding states that, for multicast in a directed network, a given multicast rate is feasible if it is feasible as a unicast rate to each receiver separately [3]. Let f t ( uv) be the s-t flow from u to v. The problem can be formulated as follows: minimize uv A c( uv) uv subject to : (1) t R, u V f t ( uv) f t ( vu) = δ t (u) t R, u, v V v V 0 f t ( uv) c( uv) (2) where 1 if u = s δ t (u) = 1 if u = t 0 otherwise Remarks. Note that in the geometric version of the multicast problem, the number and locations of relay nodes are not specified in the input, and the mathematical program above can not be solved using regular optimization techniques. Which terminal assumes the role of the multicast source is actually not important for constructing an optimal multicast network. If the role between a receiver and the source is switched, we only need to reverse some links (or reverse parts of the link s capacity) of the original optimal multicast network to support the new multicast session [4]. B. Discrete Model We next transform the optimal multicast network design problem into its discrete version, which will be utilized in our later analysis. In a practical network coding or routing solution, information flow rates are rational numbers. We can define a unit flow rate as one symbol (from a finite field F) per unit time. Then all the information flow rates and link capacities are integral. A link with capacity larger than 1 can be replaced with parallel links of unit capacity. Let D(V,A) denote such a directed network, and λ D (s, t) denote the maximum number of linkdisjoint paths from s to t in D. Then with linear network coding [5], h = min t R λ D (s, t) symbols can be transmitted to each receiver per unit time. Therefore, the average cost for a unit rate multicast session is 1 h uv A uv, and the problem to find such a min cost discrete network can be formulated as: 1 min h N h uv A uv + subject to : (3) λ D (s, t) h, t R Note that the Euclidean Steiner tree problem can be viewed as a special case with h =1. To show that the two formulations (1) and (2) are equivalent, we only need to verify that the optimal solution of (1), if existent, can always be achieved by rational link capacities. This is true because all the coefficients and constants of the linear constraints in (1) are rational. So we can scale each link s capacity with the least common multiple of their denominators to make each link s capacity integral to obtain an optimal solution of (2). C. Optimal embedding of a multicast network By definition, a solution to a space information flow problem has two components: a multicast network D(V,A) and an embedding (location assignment of each relay node). In this subsection, we discuss properties that an optimal embedding of a multicast network must possess. Some of these properties are utilized in our proofs in the next section. In the literature of the Euclidean Steiner Tree problem [6], Gilbert and Pollak study the properties of relatively minimal

Fig. 3. Wedge property. If W contains no terminal nodes and each relay node has three non-zero length links, W contains no relay nodes. Steiner trees. Some of such properties can be extended to the scenario of space information flow. Besides the goal of minimizing total network cost, we have no constraints on the embedding, e.g., the embedding does not have to be planar even if the space is 2-D, and links may intersect. We say an embedding is relatively minimal with respect to network D(V,A), if we can not reduce the overall network cost by perturbing a relay node, i.e., by changing the position of some relay node slightly. When the position of a relay node v is changed by dx, the length of link uv is changed by uv uv dx. If uv =0, the length is increased by dx. Therefore, we have the following properties: Stable at Relay. Imagine each link of non-zero length as an elastic band with the unusual property of having unit tension regardless of how much it is stretched. For links of zero length, imagine them as glue binding the two nodes, which can provide a unit passive force at most. Suppose there is some external force at each terminal to hold its position. Then in any relatively minimal embedding, the position of each relay node is stable. In fact, the combined external force at any set of relay nodes is zero. Convex Hull. In any relatively minimal embedding, all relay nodes lie in the convex hull of T. Convexity. Let ρ 1 and ρ 2 be two embeddings of the same network, with cost C 1 and C 2, respectively. For any p, q 0,p+q =1, the embedding ρ = pρ 1 +qρ 2 has cost C pc 1 + qc 2. Therefore, any relatively minimal embedding achieves the minimum cost of the given multicast network. Unlike in the minimum Steiner tree problem, a relay node may have more than three adjacent links in the optimal multicast network enabled by network coding. Thus some properties of the relatively minimal Steiner tree [6] do not hold for the relatively minimal embedding, because these properties depend on the assumption that a Steiner (relay) node has three non-zero length links. We translate one of these properties into a conditioned form for the optimal embedding in a plane, which is useful in our proof to the three terminals case: Wedge Property. Let W R 2 be any open wedge-shaped region with an angle of at least 120 (Fig. 3). In any relatively minimal embedding on a plane where each relay node has three non-zero length links, if W does not contain any terminal node, W contains no relay node [6]. III. THE EQUIVALENCE BETWEEN STEINER TREES AND MIN-COST MULTICAST NETWORKS FOR THREE TERMINALS Space information flow is a new field of research, and basic problems remain open on the design of a min-cost multicast network in a geometric space. Three important ones are listed below. Achievability. Given a finite set of terminals, is the minimum cost achievable with a finite network, i.e., a multicast network with a finite number of relay nodes? In the problem formulation, no bound is enforced on the size of a solution D(V,A). There is a possibility that no finite network is optimal, in that for any finite multicast network, there exists another network that achieves a less cost by including more relay nodes. Benefits of network coding. How much is the difference between the cost of a minimum Steiner tree and a minimum network coding based solution? In Li and Wu s work [1] that proposes the space information flow problem, such a gap is studied, with upper-bound on the gap proven for special cases. Complexity. Is there an efficient (polynomial-time) algorithm that computes the optimal solution? While Euclidean Steiner Tree is a well known NP-hard problem, it is unclear whether the new problem of min-cost multicast network design in (1) is NP-hard or not. The example in Fig. 1 reveals that these two problems have different underlying structures, and may have different computational complexity. These questions are indeed co-related. In this work, we present two results toward these directions. In the rest of this section, we prove that in the most basic scenario, when the input size is small (three terminals), network coding does not make a difference. This leads to an interesting contrast to the network information flow case (optimal multicast in a fixed, existing network), where in the well-known threeterminal butterfly network [3], network coding outperforms routing. In the next section, we prove that finitely many relays always suffice, and present closed-form upper-bounds on the number of relays required for achieving minimum cost. As a basic scenario, the three terminal case is well-studied in the Euclidean Steiner tree literature. The minimum Steiner tree for 3 terminals contains at most 1 relay node of degree 3. From the Stable at Relay property, the three links must meet at the relay with three 120 angles. Such a location is called the Fermat Point. Let s define the cost advantage as the ratio of the minimum network cost without coding over that with network coding. By showing that the cost advantage is 1 for three terminals, we can see that network coding is unnecessary for this case, and the min-cost multicast network is a minimum Steiner tree. Theorem 1. If T =3, the cost advantage is always 1. In other words, the minimum Steiner tree achieves the minimum cost, which cannot be improved by network coding. Proof: We prove the theorem by way of contradiction. Suppose there is a configuration T = {s, t 1,t 2 } with cost advantage strictly larger than 1. Then there must be a network

starts to make a difference for three terminals already in the butterfly network, there are one source and two terminals. Fig. 4. Splitting a relay node. with total cost strictly less than the cost of a minimum Steiner tree. Let D(V,A) be such a minimal network which has the minimum number of links. We will show that we can construct a network D with the same number of links and cost no more than the cost of D, while its optimal embedding has no zero length links and each relay has 3 links at most. Therefore, applying the Wedge Property concludes that all the relay nodes locate on the edge of the minimal Steiner tree, which contradicts the assumption. We construct the desired topology D from D by splitting a relay node of degree larger than 3 repeatedly, while preserving the max flow to each receiver. As illustrated in Fig. 4, splitting a relay node corresponds to replacing the relay node with two nodes, and dividing the adjacent links into two none empty parts connected to the two nodes respectively. Note that if such splitting exists, the total cost does not increase and there is no zero length links in the optimal embedding of D. Otherwise, we can merge the two ends of a zero length link to obtain a new topology with the same cost but less links, which conflicts with the fact that D has the minimum number of links. Therefore, we can obtain the desired network satisfying the condition of Wedge Property. Let (f 1,f 2 ) be a max flow solution for receiver t 1 and t 2. As we are considering the discrete model, all links can be divided into three classes: (A) with flow (f 1 (e),f 2 (e)) = (0, 1); (B) with flow (1, 0); (C) with flow (1, 1). For a relay node of degree larger than 3, if there exist an incoming link and an outgoing link belonging to the same class, we complete the splitting by connecting these two links to a new node. Otherwise, all incoming links and outgoing links belong to different classes. According to the flow conversation law, it can be verified that classes A and B always belong to the same side, i.e., either incoming links belong to class A and B, and outgoing links belong to class C, or the other way around. In any case, we find a class A link, a class B link and a class C link to complete the splitting operation. Discussion. As n grows larger, eventually there will be configurations with cost advantage strictly larger than 1. Recall that in the example illustrated in Fig. 1, six terminals are used to construct a multiple source example where cost advantage is strictly larger than 1. It is unknown whether the cost advantage can be larger than 1 for the cases n =4and n =5. For multicast in a network (instead of in a space), network coding IV. UPPER-BOUNDS ON THE NUMBER OF RELAY NODES In the literature of network coding, the case of two integral flows has attracted considerable research interests [7][8]. It represents the most basic scenario where network coding can make a difference from routing. For the case h =1, we are back to multicast trees. In this section, we analyze the maximum number of relay nodes that can be required in an optimal solution to the multicast problem in space. We focus on the two flow case first, and then extend our discussions to more general cases of h>2. Once the number of relay nodes is upper bounded, the problem in (1) becomes an optimization problem of finite variables, simplifying its solution algorithm design. For the case h =1, studies on the Euclidean Steiner tree problem have shown that the minimum cost can be achieved by adding n 2 relay nodes at most. Our result essentially generalizes the study to the case h =2. Before stating the theorem, we first introduce the concepts of a h-minimal network and a subtree decomposition. While searching for the optimal solution, it is reasonable to consider only the h-minimal networks [9], where deleting any link will cause λ D (s, t) <hfor some receiver t. According to Li, et al. [5], for any h-minimal network, there is a linear network code where each link is assigned a global coding vector from the h dimension linear space F h. Therefore in a h-minimal network, each node has in-degree at most h, since the coding vector on an extra link would be linearly dependent with the other h coding vectors and thus redundant. For further analyzing the structure of a h-minimal network, Fragouli and Soljanin introduce the feasible subtree decomposition [7], which divides all the links into subtrees such that there exists a feasible coding solution with links from the same subtree allocated the same coding vector. These subtrees are classified into source subtrees, rooted at the source, and coding subtrees, rooted at non-source nodes. Theorem 2. When h =2, there exists an optimal multicast network that has at most (2n 3)(2n 2)+n 1 relay nodes. Proof: Consider an optimal solution with h = 2 that has the minimum number of links. If there exists a splittable relay node, we apply the splitting operation to it. During the splitting, the cost of the network and the number of links in the network both remain unchanged, and the number of nodes increases by 1. Since the total number of nodes is upper-bounded by twice the number of links, which is not changing, sooner or later we have no more relay nodes to split. Let D(V,A) be the optimal network after such splitting. In network D, a relay node with in-degree larger than 1 must be an encoding node, since otherwise it can be splitted. As an optimal solution, D must be 2-minimal. Hence each node has in-degree no more than 2. The relay nodes can be categorized into two sets: 1) inner Steiner nodes with in-degree 1, which appear exactly once in a minimum Steiner subtree;

2) coding nodes with in-degree 2. We prove an upper-bound on the size of each set separately. For the minimal subtree decomposition, we have the following lemma: Lemma. (C. Fragouli and E. Soljanin [7]) For minimal subtree decomposition of a 2-minimal network, each coding subtree contains at least two receivers and each source subtree contains at least one receiver. As each receiver appears in exactly two different subtrees, the number of coding subtrees is no more than n 1. Since a coding node is the root of at least one coding subtree, the number of coding nodes is also no more than n 1. On the other hand, each subtree is a minimum Euclidean Steiner tree connecting a subset of the n terminals and no more than n 1 coding nodes. Therefore, each subtree contains 2n 3 inner Steiner nodes at most. According to the previous analysis, there are no more than 2(n 1) subtrees, since each subtree contains at least one receiver node. To conclude, the number of inner Steiner nodes is upper-bounded by (2n 3)(2n 2). Acyclic networks with arbitrary h. According to the result that the number of coding nodes is upper bounded by h 3 (n 1) 2 in an acyclic multicast network [9], we can show that the number of relay nodes is upper bounded in such networks in a similar way: in an h-minimal network, any relay node with in-degree larger than 1 must be a coding node and thus the total number of relays is bounded by h 3 (n 1) 2 ; for the relay nodes with in-degree 1, they appear exactly once in each coding subtree. The number of coding forests, each of which is composed of coding subtrees of the same coding vector, is no more than n h, since a finite field of size n is sufficient and the number of different coding vectors is no more than n h. As each coding forest has n + h 3 (n 1) 2 2 inner Steiner nodes (relay node with in-degree 1) at most, we can see that the number of relay nodes with in-degree 1 is no more than n h (n + h 3 (n 1) 2 2). [4] Z. Li, B. Li, and L. C. Lau, A constant bound on throughput improvement of multicast network coding in undirected networks, IEEE Transactions on Information Theory, vol. 55, no. 3, pp. 1016 1026, Mar. 2009. [5] S.-Y. Li, R. Yeung, and N. Cai, Linear network coding, IEEE Transactions on Information Theory, vol. 49, no. 2, pp. 371 381, Feb. 2003. [6] E. Gilbert and H. Pollak, Steiner minimal trees, Society for Industrial and Applied Mathematics, vol. 16, no. 1, pp. 1 29, Jan. 1968. [7] C. Fragouli and E. Soljanin, Information flow decomposition for network coding, IEEE Transactions on Information Theory, vol. 52, no. 3, pp. 829 848, 2006. [8] S. el Rouayheb, C. Georghiades, and A. Sprintson, Network coding in minimal multicast networks, in Information Theory Workshop, 2006. ITW 06 Punta del Este. IEEE, march 2006, pp. 307 311. [9] M. Langberg, A. Sprintson, and J. Bruck, The encoding complexity of network coding, IEEE Transactions on Information Theory, vol. 52, no. 6, pp. 2386 2397, june 2006. V. CONCLUSION We studied the multicast version of the space information flow problem in this work, which models the design of the blueprint of a min-cost multicast network. We showed through a simple example that the design of an information network is different from that of a transportation network. We discussed a number of properties that an optimal multicast network embedding must possess. For the basic case of three multicast terminals, we prove that network coding does not make a difference, contrasting the problem of multicast in networks. We finally prove upper-bounds on the number of relay nodes required in min-cost multicast networks. REFERENCES [1] Z. Li and C. Wu, Space Information Flow, submitted to NetCod 2012. [2], Space Information Flow: Multiple Unicast, submitted to IEEE ISIT 2012. [3] R. Ahlswede, N. Cai, S.-Y. Li, and R. Yeung, Network information flow, IEEE Transactions on Information Theory, vol. 46, no. 4, pp. 1204 1216, Jul. 2000.