Fact-checked by Grok 2 weeks ago

Perfect Match

In , a is a set of edges without common such that every in the is incident to exactly one edge in the matching. This requires the graph to have an even number of vertices. Perfect matchings are a special case of matchings and play a central role in , with applications in areas such as , network routing, and molecular chemistry. The concept applies to both bipartite and general graphs, though conditions for existence differ: characterizes perfect matchings in bipartite graphs, while Tutte's theorem provides a criterion for general graphs. Efficient algorithms exist for finding perfect matchings, including the for bipartite cases and Edmonds' for general graphs. Key properties are captured by theorems like the König-Egerváry theorem and the Berge-Tutte formula.)

Definition

Bipartite Graphs

A is an undirected graph whose set can be partitioned into two disjoint subsets U and V such that every edge in the graph connects a in U to a in V, with no edges within U or within V. In a G = (U \cup V, E) where |U| = |V|, a is a set of edges M \subseteq E such that every in U and every in V is incident to exactly one edge in M. Formally, M is a if |M| = |U| = |V| and no two edges in M share a common . For example, the K_{n,n}, which has n in each and every possible between them, always admits a , as one can pair each in U directly to a distinct in V. Consider a simple with partitions U = \{u_1, u_2\} and V = \{v_1, v_2\}. If the are \{u_1v_1, u_1v_2, u_2v_1, u_2v_2\} (forming K_{2,2}), then M = \{u_1v_1, u_2v_2\} is a , covering all . In contrast, if the are only \{u_1v_1, u_2v_2\}, this set forms a , but removing one , say u_2v_2, leaves no , as u_2 cannot be covered without sharing .

General Graphs

In a general undirected graph G = (V, E), a perfect matching is a matching M \subseteq E such that every v \in V is incident to exactly one edge in M. This requires |V| to be even, since the size of the matching satisfies |M| = |V|/2. The notation for incidence remains standard: an edge e \in E is incident to vertices u, v \in V if e = uv, and in non-bipartite settings, edges may connect vertices without restriction to predefined partitions, allowing for structures like odd-length cycles across potentially multiple connected components. A differs from a maximum matching in that it demands full saturation of all vertices, whereas a maximum matching merely maximizes the number of covered vertices without necessarily including every one. For instance, the C_4 possesses a perfect matching formed by its two non-adjacent edges, fully pairing its four vertices. In contrast, the C_3 (a triangle) lacks a perfect matching, as its odd number of vertices prevents complete pairing. Unlike the bipartite case, where perfect matchings align with balanced partitions, general graphs introduce complexities from arbitrary connectivity and cycles of any length. The presence of a perfect matching in such graphs is often hinted at by the absence of structural obstructions known as Tutte barriers, which capture deficiencies in vertex coverage across components.

Existence Conditions

Hall's Marriage Theorem

Hall's Marriage Theorem states that in a G = (U \cup V, E) where |U| = |V| = n, a exists if and only if for every S \subseteq U, the neighborhood N(S) of S in V satisfies |N(S)| \geq |S|. This condition, known as Hall's condition, ensures that no group of vertices in U is underserved by connections to V, preventing bottlenecks in pairing. The theorem was originally formulated in terms of systems of distinct representatives for families of sets, where each set corresponds to possible matches for an in U. The theorem derives its name from an analogy to the "marriage problem," where vertices in U and V represent two groups of individuals (e.g., men and women), and edges indicate compatible pairings; Hall's condition guarantees that stable, complete pairings are possible without leaving anyone unmatched. Philip Hall proved the theorem in 1935 in his seminal paper "On Representatives of Subsets," published in the Journal of the London Mathematical Society, establishing it as a cornerstone of combinatorial matching theory. A proof outline, attributed to Dénes Kőnig, leverages the . Construct a with a connected to all vertices in U by edges of capacity 1, edges from U to V with capacity 1 matching the graph's edges, and V connected to a by capacity-1 edges. A maximum flow of value n corresponds to a . If the maximum flow is less than n, the min-cut theorem identifies a saturating cut whose U-side S violates Hall's condition, as the cut capacity equals |N(S)| + (|U| - |S|) < n, implying |N(S)| < |S|. An alternative proof by contradiction assumes a maximum matching of size less than n and uses augmenting paths (alternating paths starting from unmatched vertices in U) to construct a violating set S, showing that the endpoints in V form a neighborhood smaller than S. To illustrate non-existence, consider a bipartite graph with U = \{u_1, u_2\}, V = \{v_1, v_2\}, and edges u_1 - v_1, u_2 - v_1. For S = \{u_1, u_2\}, N(S) = \{v_1\}, so |N(S)| = 1 < 2, violating Hall's condition; indeed, no perfect matching exists, as v_2 remains isolated from pairings. This example highlights how a single overloaded vertex in V blocks completeness. For instance, in symmetric balanced incomplete block designs (BIBDs), the theorem underpins constructions of Youden squares by verifying that block neighborhoods cover points sufficiently.

Tutte's Theorem

Tutte's theorem provides a necessary and sufficient condition for the existence of a perfect matching in an arbitrary finite undirected graph, generalizing Hall's marriage theorem from bipartite graphs to general graphs. Specifically, a graph G = (V, E) has a perfect matching if and only if for every subset S \subseteq V, the number of odd components in the subgraph G - S (induced on V \setminus S) is at most |S|, denoted as o(G - S) \leq |S|. This condition captures the structural barriers to perfect matchings through the parity of components after vertex removals. The theorem was established by William T. Tutte in 1947, building on earlier work in graph factorization and responding to Petersen's conjecture on 1-factors in cubic bridgeless graphs. Tutte's result resolved the general case for perfect matchings, proving that the condition holds equivalently across all simple graphs without multiple edges. The proof proceeds in two directions. Necessity follows directly: if G has a perfect matching M, then for any S \subseteq V, each odd component in G - S must have at least one edge from M incident to a vertex in S, implying at most |S| such components since M matches vertices injectively. For sufficiency, assume the condition holds but no perfect matching exists; extend G to an edge-maximal graph G' without a perfect matching (by adding edges if needed). In G', identify a barrier set S where o(G' - S) > |S|, leading to a contradiction via the existence of universal vertices or shortest augmenting paths that would imply a larger matching. This construction often invokes on maximum matchings and alternating paths, or ear decompositions, to resolve the barrier. A classic example of violation occurs in the complete bipartite graph K_{1,3} (a with three leaves). Removing the center S = \{c\} leaves three isolated vertices, each an odd component, so o(G - S) = 3 > 1 = |S|, confirming no exists in this four-vertex graph. Related to Tutte's are factor-critical graphs, which are connected graphs on an odd number of vertices such that G - v admits a for every v \in V; these graphs have no themselves but satisfy the theorem's with equality for S = \emptyset (one odd component, the whole ). In bipartite graphs, Tutte's theorem reduces to , as removing S from one side cannot produce odd components disconnected from the other side in a balanced bipartition.

Algorithms

Edmonds' Blossom Algorithm

Edmonds' matching algorithm, published in 1965, provides a polynomial-time solution for computing a maximum matching in general undirected graphs, extending the augmenting path technique from bipartite cases to handle non-bipartite structures. The algorithm iteratively builds the matching by searching for augmenting paths relative to the current matching M, but unlike bipartite algorithms, it must address odd-length alternating cycles—termed blossoms—that can obstruct the search. These blossoms are detected and temporarily contracted into supervertices, allowing the search to proceed in a simplified graph until an augmenting path is found, after which the contractions are reversed to update the original matching. This process continues until no further augmenting paths exist, yielding a maximum cardinality matching. The algorithm operates as follows: Start with an empty matching M and a set of free (unmatched) vertices X. Construct a directed auxiliary graph where non-matching edges point away from matched vertices and matching edges point toward them, facilitating the search from free vertices. Use breadth-first search (BFS) to explore layers of even and odd distances from X, building an alternating forest. If a free vertex is reached, an augmenting path exists and M is symmetric-differenced with it to increase its size by one. If during the search a non-tree edge connects two vertices in the same layer (indicating an odd alternating cycle), a blossom is identified: the cycle is the unique paths from the lowest common ancestor to the endpoints plus the connecting edge. This blossom is contracted into a single supervertex, the graph is reduced accordingly, and the search recurses on the new graph with the adjusted matching (removing internal matching edges from the blossom). Upon finding an augmenting path in the contracted graph, it is lifted back to the original by expanding the blossom along an even-length alternating path within it relative to the contracted matching. The entire phase of finding and augmenting along one path, including all contractions and expansions, repeats up to |V|/2 times. The core innovation lies in the blossom contraction mechanism, which transforms problematic odd alternating cycles into a single vertex while preserving the matching structure, effectively "simplifying" the non-bipartite graph to mimic bipartite behavior during the search without losing augmenting opportunities. This contraction identifies the blossom's base (the attachment point to the rest of the forest) and ensures that any augmenting path found post-contraction can be uniquely expanded within the blossom to form a valid path in the original graph, maintaining the algorithm's correctness. The supervertex is treated as free if the blossom contains an odd number of free vertices. When expanding an augmenting path ending at the supervertex, if the base is free, simply add the attachment edge; if the base is matched internally, flip the matching along an odd-length alternating path from the base to a free vertex (the tip) within the blossom to free the base before adding the external edge. In terms of efficiency, the original formulation runs in O(|V|^2 |E|) time, which is O(|V|^4) for dense graphs, arising from O(|V|) augmentations, each involving O(|V|) BFS traversals and O(|V|^2) time for handling contractions and expansions. Later refinements, such as those avoiding explicit recursive contractions through careful labeling and union-find structures, achieve O(|V|^3) time overall. To illustrate, consider a graph with vertices 1, 2, 3, 4 and edges 1-2, 2-3, 3-1 (triangle), 3-4. Initial matching M = {2-3}, so free vertices 1 and 4. The search starts from free vertex 4 (level 0). Add 3 at level 1 via non-matching edge 4-3. Since 3 is matched, add 2 at level 2 via matching edge 3-2. From 2 (level 2 even), add 1 at level 3 odd via non-matching edge 2-1. From 1 (level 3 odd), the non-matching edge 1-3 connects to 3 at level 1 (odd), same parity, detecting the blossom: cycle 3-2-1-3 (alternating: 3(M)-2(nonM)-1(nonM to 3), but precisely the paths 3-2-1 plus back 1-3). Contract blossom B = {1,2,3} into supervertex s (base at 3). Reduced graph: s-4, reduced M = ∅ (internal {2-3} shrunk), s free (odd number of free vertices in B: only 1). Augmenting path 4-s (add 4-3). For expansion: base 3 was matched internally, so flip along odd-length alternating path from 3 to free vertex 1: 3(M)-2(nonM)-1 (edges 3-2, 2-1). Flip removes 3-2, adds 2-1. Now internal M = {1-2}, freeing 3. Then add external 4-3. New M = {1-2, 3-4}, size increased by 1. High-level pseudocode for the algorithm is as follows:
function MAXIMUM_MATCHING(G):
  M ← ∅
  while true:
    forest ← BUILD_ALTERNATING_FOREST(G, M)
    if no augmenting path in forest:
      return M
    P ← FIND_AUGMENTING_PATH(forest)  # with blossom handling
    M ← M Δ P  # symmetric difference to augment

function FIND_AUGMENTING_PATH(forest):
  use BFS to build layers from free vertices
  while blossoms detected:
    contract blossom into supervertex
    recurse on reduced graph
  if free vertex reached, trace path P
  expand contractions along P to get original path
  return P
This outline captures the recursive nature of contraction and expansion without full implementation details.

Hopcroft-Karp Algorithm

The , developed by John E. Hopcroft and Richard M. Karp in 1973, computes a in a by iteratively identifying and augmenting along multiple vertex-disjoint augmenting paths in successive phases. Unlike earlier methods that augment a single path per iteration, this approach leverages the structure of bipartite graphs to find a collection of shortest disjoint augmenting paths simultaneously, significantly reducing the number of iterations required. The algorithm terminates when no augmenting paths remain, yielding a maximum matching as per Berge's lemma. The algorithm operates in phases, each consisting of two main steps. First, a (BFS) is performed from all unmatched in one (say, the left side) to build a layered , where levels are assigned based on the shortest path distances to unmatched vertices in the other partition. This layering captures all minimal-length augmenting paths. Second, multiple depth-first searches (DFS) are conducted within the layered graph to discover a maximal set of vertex-disjoint augmenting paths of this minimal length; each such path is traversed only along edges that respect the level structure (forward edges increase level by 1, backward edges decrease by 1). These paths are then used to flip the matching edges simultaneously, augmenting the overall matching by the number of paths found. Phases continue until the BFS finds no path to an unmatched vertex. A distinguishing feature of the is its phase-based augmentation strategy, which ensures that each phase increases the matching size by at least one but typically by many disjoint paths, all of equal shortest length. This avoids the inefficiency of single-path augmentations, which can require up to O(|V|) iterations in the worst case, by bounding the number of phases to O(\sqrt{|V|}) through the progressive shortening of augmenting paths. The use of DFS for allows efficient and reuse of partial searches, further optimizing the process within each phase. The time complexity of the Hopcroft–Karp algorithm is O(|E| \sqrt{|V|}), where |V| denotes the number of vertices and |E| the number of edges, achieved via O(\sqrt{|V|}) phases, each taking O(|E|) time for BFS and DFS traversals. This complexity outperforms prior algorithms like the Hungarian method's O(|V|^3) bound, especially in dense graphs where |E| \approx |V|^2 / 2, making it practical for large-scale bipartite matching problems. To illustrate, consider a bipartite graph with left partition U = {u_1, u_2, u_3} and right partition W = {w_1, w_2, w_3}, connected by edges u_1–w_1, u_1–w_2, u_2–w_2, u_2–w_3, u_3–w_1, u_3–w_3. With an initial empty matching, the first phase's BFS creates levels reaching all w_i at distance 1. DFS then identifies two disjoint paths, such as u_1–w_1 and u_2–w_2, augmenting the matching to size 2. In the second phase, BFS from the remaining free u_3 finds the path u_3–w_3 (length 1), augmenting it alone to reach the maximum matching of size 3. This example highlights how the algorithm processes multiple paths in parallel within phases, completing in two iterations rather than three if using single-path augmentation. The models bipartite matching as a in a with unit capacities: a source connects to all left vertices, a sink to all right vertices, and partition edges have capacity 1. It corresponds to Dinic's blocking-flow algorithm applied to this network, where each phase computes a blocking flow in the level graph built by BFS, ensuring the O(|E| \sqrt{|V|}) bound.

Key Properties and Theorems

König-Egerváry Theorem

The König-Egerváry theorem asserts that, for any bipartite graph G, the cardinality of a maximum matching \nu(G) equals the cardinality of a minimum vertex cover \tau(G). This equality distinguishes bipartite graphs from general graphs, where \nu(G) \leq \tau(G) holds but equality does not always occur. The theorem was established by Hungarian mathematician Dénes König in his 1916 paper, where he proved the result as part of broader investigations into graph theory and its applications to determinant theory and set theory. Independently, Jenő Egerváry rediscovered and extended the result in 1931 to the weighted case, demonstrating a similar duality between maximum-weight matchings and minimum-weight vertex covers in weighted bipartite graphs. A standard proof proceeds in two directions. The inequality \nu(G) \leq \tau(G) is immediate, as the endpoints of any matching form a of the same size. For the reverse, consider a maximum matching M^* in the with partitions A and B. Direct the edges as follows: non-matching edges from A to B, and matching edges from B to A. Let L_A \subseteq A be the vertices reachable from unsaturated (exposed) vertices in A via alternating paths in this , and let L_B = B \cap L be the corresponding reachable vertices in B. The set C^* = (A \setminus L_A) \cup L_B forms a of size |M^*|, since any edge not covered would imply an augmenting path relative to M^*, contradicting maximality; moreover, no unsaturated vertices in A reach outside C^*, ensuring all edges are incident to C^*. This construction relies on implicitly, as the absence of augmenting paths ensures the cover's minimality. As an implication, in a balanced with |V| = 2n vertices, a exists \tau(G) = n, since such a matching saturates all vertices and thus equals the minimum cover size under the theorem. For illustration, consider a with partitions A = \{u_1, u_2\} and B = \{v_1, v_2, v_3\}, and edges u_1v_1, u_1v_2, u_2v_2, u_2v_3. A maximum matching is \{u_1v_1, u_2v_3\} of size 2, and a minimum is \{u_1, u_2\} (or equivalently \{v_1, v_2, v_3\} but minimized to size 2), confirming \nu(G) = \tau(G) = 2. Egerváry's extension to weighted bipartite graphs underpins the solution to the , where the maximum-weight equals the minimum cost in a formulation, enabling efficient algorithms like the Hungarian method.

Berge-Tutte

The Berge-Tutte provides a min-max of the size of a maximum matching in a general , extending the conditions for perfect matchings to arbitrary matching sizes. Developed by Claude Berge in 1958, building on W. T. Tutte's 1947 theorem, the formula expresses the matching number \nu(G) of a G = (V, E) as \nu(G) = \frac{1}{2} \min_{S \subseteq V} \left( |V| + |S| - o(G - S) \right), where o(G - S) denotes the number of odd-sized connected components in the subgraph induced by V \setminus S. This formulation defines the deficiency of G as \mathrm{def}(G) = \max_{S \subseteq V} (o(G - S) - |S|), yielding the equivalent expression \nu(G) = \frac{1}{2} (|V| - \mathrm{def}(G)). The proof derives from Tutte's theorem, which ensures a perfect matching exists if and only if o(G - S) \leq |S| for all S \subseteq V. For general matchings, consider a maximum matching M in G; the unsaturated vertices form odd components in G - S for appropriate S, leading to the upper bound \nu(G) \leq \frac{1}{2} (|V| + |S| - o(G - S)) by analyzing edges incident to S and components. Equality holds by minimizing the deficiency over subsets, often proven via induction on |V| and application of Hall's theorem in an auxiliary constructed from maximum-deficiency sets. This formula enables computation of the matching number without explicitly enumerating matchings, serving as a theoretical tool for analyzing structural obstructions to larger matchings in non-bipartite graphs. For instance, in the star graph K_{1,3} with |V| = 4, \nu(G) = 1; taking S as the center yields o(G - S) = 3 (three isolated leaves), so \frac{1}{2} (4 + 1 - 3) = 1, achieving the minimum, while other S give larger values. The Berge-Tutte formula generalizes the König-Egerváry theorem from , where the matching number equals the number; in general , the odd-component term accounts for non-bipartite barriers, reducing to the bipartite case when o(G - S) aligns with unsaturated vertices.

Applications

Combinatorial Optimization

Perfect matchings play a central role in problems, where they correspond to common bases in specific matroids defined on the edge set of a . In , the set of matchings forms the independent sets of the of two matroids: one partitioning the edges by their left endpoints and the other by their right endpoints. A in such a graph is thus a common basis of maximum size in this , enabling the use of matroid algorithms to find or verify their existence. Weighted perfect matchings are instrumental in solving the (), which seeks a minimum-length closed tour traversing every edge of an undirected at least once. For a connected , the solution involves identifying the odd-degree vertices and constructing an auxiliary where vertices represent these odd-degree nodes, and edge weights are the shortest-path distances in the original . A minimum-weight in this auxiliary determines the minimum-length paths to add (as duplicate edges) to make all degrees even, yielding an Eulerian whose tour solves the . This reduction leverages efficient algorithms for minimum-weight s, such as Edmonds' algorithm, to achieve polynomial-time solvability for undirected s. The enumeration of perfect matchings in graphs is captured by matrix functions analogous to the . For bipartite graphs, the number of perfect matchings equals the permanent of the biadjacency matrix. In general (non-bipartite) graphs, it is given by the hafnian of the symmetric A, defined as \haf(A) = \sum_{\substack{\pi \\ \text{perfect matching}}} \prod_{(i,j) \in \pi} A_{i,j}, where the sum is over all s \pi of the complete graph on the vertices labeled by the rows/columns of A, and the product is over the pairs in \pi. For unweighted graphs with A_{i,j} = 1 if an edge exists and 0 otherwise, \haf(A) directly counts the perfect matchings. Counting the number of perfect matchings is computationally challenging, as Valiant proved in 1979 that this problem is #P-complete, even for bipartite graphs where it reduces to computing the permanent of a (0,1)-matrix. This hardness underscores the difficulty of exact enumeration, motivating approximation algorithms and randomized methods for large instances.90044-6) A representative application arises in scheduling problems, such as assigning n tasks to n processors without conflicts, modeled as finding a perfect matching in a bipartite graph where one part represents tasks, the other processors, and edges indicate compatible assignments. The existence of a perfect matching ensures an optimal conflict-free schedule, solvable via maximum matching algorithms like Hopcroft-Karp. Perfect matchings also connect to tilings in through dimer models, where a dimer covering of a —such as a planar —is precisely a , with each dimer occupying an to cover all vertices. Seminal work by Kasteleyn in provided a Pfaffian-based to count such coverings on lattices, linking combinatorial enumeration to physical models of adsorption and transitions.90130-2)

Economics and Bipartite Matching

In economic models, perfect matchings in bipartite graphs play a central role in analyzing two-sided markets where agents on one side (e.g., workers) are paired with agents on the other (e.g., firms) based on preferences or costs. The stable marriage problem, introduced by Gale and Shapley, models such pairings as a bipartite graph where edges represent acceptable matches ranked by mutual preferences, and a perfect stable matching exists if Hall's marriage condition holds, ensuring no subset of agents on one side lacks sufficient connections to the other. The Gale-Shapley algorithm, also known as deferred acceptance, computes such a perfect stable matching by having one side propose in order of preference, with the other side tentatively accepting or rejecting based on their rankings; this guarantees stability—meaning no blocking pair of agents both prefer each other over their assigned matches—and optimality for the proposing side, assuming equal-sized partitions. This framework ensures efficient, incentive-compatible allocations in preference-based economies, as participants cannot benefit by misrepresenting preferences in the proposing role. The extends this to cost-minimizing s in weighted bipartite graphs, where edge weights represent assignment costs (e.g., wages in job markets). The , developed by Kuhn, solves this by iteratively adjusting dual variables (prices) to find a minimum-cost , equivalent to a where the dual provides shadow prices for resources. In economic contexts like job-worker assignments, it optimizes total surplus by pairing high- matches while respecting budget constraints, and its polynomial-time complexity enables practical implementation in large . For instance, in labor , it models firm-worker allocations to minimize costs subject to productivity, yielding competitive equilibria where prices clear the . In housing markets modeled as bipartite graphs (buyers vs. sellers with indivisible ), perfect matchings underpin the of competitive equilibria, where prices adjust to supply and demand, ensuring Pareto-efficient allocations. Shapley and Shubik's assignment game demonstrates that —stable outcomes resistant to coalitional deviations—consists of all perfect matchings with supporting price vectors that make unmatched trades unprofitable, guaranteeing equilibrium under transferable utility. This bipartite structure captures one-sided preferences over , with equilibria emerging from the of stable matchings, where buyer-optimal and seller-optimal outcomes bound the set. Such models inform policy in rental or housing voucher systems, where perfect matchings prevent shortages via price-mediated clearing. A prominent real-world application is kidney exchange programs, where patient-donor pairs form a with edges indicating compatibility for transplants. Roth, Sönmez, and Ünver model pairwise exchanges as maximum-weight perfect matchings to maximize successful transplants, using algorithms like the Hungarian method to pair incompatible donors with recipients while ensuring cycle fairness and efficiency. Empirical implementations in the U.S., such as through national clearinghouses like the National Kidney Registry and UNOS, have significantly increased the number of living donor transplants, with kidney paired donation accounting for nearly 20% of such transplants as of 2021 and facilitating over 1,400 transplants in 2023. This approach leverages bipartite perfect matchings to resolve blood-type incompatibilities, treating exchanges as costless trades in a stability-focused market. In game-theoretic terms, perfect stable matchings correspond to in strategic games where simultaneously report preferences over potential partners. In the induced matching game from Gale-Shapley, truth-telling by the proposing side yields a subgame-perfect , as deviations lead to worse outcomes due to ; more generally, any stable perfect matching is a pure in the direct revelation game, where no benefits unilaterally by altering their . This linkage highlights in economic designs, preventing strategic misrepresentation that could destabilize allocations. Empirically, bipartite perfect matchings inform 21st-century applications like and . In , the Gale-Shapley algorithm has been adopted in systems like City's since 2003, assigning students to seats via student-proposing deferred acceptance to achieve , envy-free perfect matchings when capacities allow, improving equity over prior serial dictatorship methods. In platforms, such as those analyzed by Hitsch et al., bipartite graphs model user matches based on revealed preferences from swipes or messages, with matching algorithms enhancing recommendation systems to increase pairing rates by simulating market clearing.

References

  1. [1]
    Perfect Match (TV Series 2023– ) - IMDb
    Rating 5.8/10 (2,758) Perfect Match: With Nick Lachey, Dom Gabriel, Francesca Farago, Joey Sasso. Couples who prove their compatibility gain the power to make or break other ...Full cast & crew · Episode list · Parents guide · Ratings
  2. [2]
    Perfect Match Season 2 Cast: Meet the Singles Entering the Villa
    Jun 24, 2024 · Yes, the beachside retreat for the boldest and most bikini-forward among us is back for Season 2, hard launching on June 7. The rules of the ...
  3. [3]
    The Perfect Match: the rules behind Netflix's new dating show
    Feb 14, 2023 · The Perfect Match is, in a nutshell, The Hunger Games of reality television. Released today on Netflix (perfect for a V-Day binge watch), ...
  4. [4]
    What is Perfect Match on Netflix? Rules explained - Radio Times
    Feb 15, 2023 · What is Perfect Match on Netflix? Rules explained. The stars of Netflix's biggest reality shows are hoping to find love – here's everything ...
  5. [5]
    'Perfect Match' Episode Release Schedule And Premiere Dates
    Feb 14, 2023 · The next four episodes of Perfect Match will be added to Netflix on Tuesday, February 14 at 3:01 a.m. ET/12:01 a.m. PT. This is a change of pace ...
  6. [6]
    Perfect Match season 2 release schedule | When are new episodes ...
    Jun 14, 2024 · Episodes 1-6 - Friday 7th June (out now); Episodes 7-9 - Friday 14th June (out now); Episode 10 (finale) - Friday 21st June. Perfect Match ...
  7. [7]
    Everything You Need to Know About Perfect Match Season 3 - Netflix
    When is the Perfect Match Season 3 release date? Starting Aug. 1, new episodes will roll out every Wednesday following these NRU singles on their mission to ...
  8. [8]
    'Perfect Match' Season 3 Release Schedule—When Do New ...
    Aug 2, 2025 · Following the Perfect Match Season 3 premiere on August 1, the next batch of episodes (Episodes 7 - 9) will drop one week later on Friday, ...
  9. [9]
    'Perfect Match' Season 3: Episode Release Schedule and Cast Guide
    Aug 4, 2025 · Netflix has confirmed that Perfect Match season 3 will begin on August 1, giving reality fans a week or so to collect themselves after Love Island USA season 7.<|control11|><|separator|>
  10. [10]
    When Was Perfect Match Season 3 Filmed? Filming ... - Netflix
    Aug 5, 2025 · Curious about when your favorite pairings coupled up? Find out everything you need to know about when and where the Season 3 of Perfect ...
  11. [11]
    Perfect Match Season 2 Release Date, Cast, Trailer, News - Netflix
    What is the Perfect Match Season 2 release date? Summer just got a lot more fun. Starting on June 7, new episodes chronicling the singles' journeys to the ...
  12. [12]
    [PDF] Graph Theory: Matchings and Hall's Theorem - cs.Princeton
    If G(V1,V2,E) is a bipartite graph than a matching M of G that saturates all the vertices in V1 is called a complete matching (also called a perfect matching).<|control11|><|separator|>
  13. [13]
    [PDF] Bipartite Graphs and Matchings - University of Toronto Mathematics
    If a graph has a perfect matching, then clearly it must have an even number of vertices. Further- more, if a bipartite graph G = (L, R, E) has a perfect ...
  14. [14]
    [PDF] Hall's Matching Theorem 1. Perfect Matching in Bipartite Graphs
    Define a perfect matching in a bipartite graph G to be an injective mapping f : VI → VO such that for every x ∈ VI there is an edge e ∈ E with endpoints ...
  15. [15]
    [PDF] Matching Algorithms
    In particular, a maximum matching is called perfect if every vertex of G is matched. A bipartite graph G is a graph in which the vertices of G can be ...
  16. [16]
    Perfect Matching -- from Wolfram MathWorld
    A perfect matching of a graph is a matching (ie, an independent edge set) in which every vertex of the graph is incident to exactly one edge of the matching.
  17. [17]
    [PDF] Math 301: Matchings in Graphs
    Note that a perfect matching can only occur in a graph with evenly many vertices. A matching M is called maximal if M ∪ {e} is not a matching for any e ∈ E(G).
  18. [18]
    [PDF] Matchings in Simple Graphs - math.binghamton.edu
    A partial matching is a set of edges where no two share endpoints. A perfect matching is a partial matching where every vertex is an endpoint of an edge.
  19. [19]
    [PDF] Discrete Mathematics - Homework 7 Solutions - math.binghamton.edu
    a) A nonbipartite graph in which every node has the same degree need not contain a perfect matching. Solution: Any odd cycle will do (though C3 is the example ...
  20. [20]
    [PDF] Section 16.4. Perfect Matchings and Factors
    Feb 7, 2022 · Every 3-regular graph without cut edges has a perfect matching. Note. The condition that G has no cut edges in Petersen's Theorem (Theorem 16. ...
  21. [21]
    [PDF] On Representatives of Subsets - Semantic Scholar
    On Representatives of Subsets · P. Hall · Published 1935 · Mathematics · Journal of The London Mathematical Society-second Series.
  22. [22]
    [PDF] Hall's Marriage Theorem and Related Aspects
    Abstract. The main focus of this thesis is to study Hall's Marriage Theorem, which was named after, and proven by the English mathematician Philip Hall in 1935.Missing: original | Show results with:original
  23. [23]
    [2503.23159] Hall's marriage theorem - arXiv
    Mar 29, 2025 · In 1935, Philip Hall published what is often referred to as ``Hall's marriage theorem'' in a short paper (P.~Hall, On Representatives of Subsets ...Missing: original | Show results with:original
  24. [24]
    [PDF] Tutte-factorization-of-linear-graphs.pdf
    A factor is a regular subgraph of the first degree. If N has no factor it is prime. Clearly all graphs of odd order are prime. Let the O-simplexes of a graph ...
  25. [25]
    [PDF] Tutte's theorem - CSA – IISc Bangalore
    Nov 19, 2014 · Proof (Sketch) of Sufficiency of Tutte's Condition ... This ends the proof of Tutte's theorem for existence of a perfect matching in a graph.Missing: statement history
  26. [26]
    [PDF] 1 Tutte's Theorem - CSE IITM
    Tutte's theorem is the analogue of Halls' Theorem which gives the necessary and sufficient condition for existence of a perfect matching.Missing: statement history sketch
  27. [27]
    [PDF] 1 Matchings in graphs - Stanford CS Theory
    Apr 28, 2014 · The following theorem is at the core of all the algorithms for Perfect Matching that we will discuss. Theorem 1.1 (Tutte). For any graph G = (V, ...Missing: statement history
  28. [28]
    [PDF] 1 Edmonds-Gallai Decomposition and Factor-Critical Graphs
    Feb 11, 2010 · Definition 2 A graph G = (V,E) is factor-critical if G has no perfect matching but for each v ∈ V ,. G − v has a perfect matching. Factor ...
  29. [29]
    [PDF] paths, trees, and flowers - jack edmonds
    PATHS, TREES, AND FLOWERS. JACK EDMONDS. 1. Introduction. A graph G for purposes here is a finite set of elements called vertices and a finite set of elements ...
  30. [30]
    [PDF] 1 2 Lecture 2 Recapitulation Flowers, Stems, and Blossoms
    Feb 5, 2004 · In this lecture, we will: • Present Edmonds' algorithm for computing a maximum matching in a (not necessarily bipar- tite) graph G. • Use ...<|control11|><|separator|>
  31. [31]
    Finding and Contracting Blossoms - Algorithms II
    To find an augmenting path for the current matching M, Edmonds's algorithm starts by computing an alternating forest F of G with all unmatched vertices in G as ...<|control11|><|separator|>
  32. [32]
    [PDF] Edmonds' Blossom Algorithm - Stanford University
    Jun 6, 2016 · Edmonds' Blossom algorithm is a polynomial time algorithm for finding a maximum matching in a graph. Definition 1.1. In a graph G, a matching ...
  33. [33]
    [PDF] Maximum Matching in General Graphs Without Explicit ...
    Hence, Edmonds original algorithm uses only O(n3) time. In [2, 17, 27, 37], it is shown how to avoid explicit shrinking of odd cycles. All these algorithms use.
  34. [34]
    [PDF] A new implementation of a minimum cost perfect matching algorithm
    In this paper we describe a new implementation of Edmonds's algorithm to which we refer as Blossom V. Our motivation was to incorporate both the variable δ ...
  35. [35]
    (PDF) Dinitz' Algorithm: The Original Version and Even's Version
    Aug 7, 2025 · In this paper, we prove two new results related to finding maximum flows in directed graphs and finding maximum matchings in bipartite graphs.
  36. [36]
    Über Graphen und ihre Anwendung auf Determinantentheorie und ...
    König, D.. "Über Graphen und ihre Anwendung auf Determinantentheorie und Mengenlehre." Mathematische Annalen 77 (1916): 453-465.Missing: Dénes 1931
  37. [37]
    [PDF] 1. Lecture notes on bipartite matching
    Feb 8, 2011 · Theorem 1.1 (König 1931) For any bipartite graph, the maximum size of a matching is equal to the minimum size of a vertex cover. We shall prove ...Missing: statement history
  38. [38]
    Jenő Egerváry: from the origins of the Hungarian algorithm to ...
    Dec 24, 2009 · We discuss some relevant results obtained by Egerváry in the early Thirties, whose importance has been recognized several years later.
  39. [39]
    [PDF] 5.1 Min-Max Theorem for General Matching - acs.uwinnipeg.ca
    Tutte-Berge formula makes this precise. In the following o(G) denotes the number of odd components of G and ν(G) denotes the maximum size of a matching.
  40. [40]
    [PDF] A short proof of the Berge–Tutte Formula and the Gallai–Edmonds ...
    Berge [2] extended Tutte's 1-Factor Theorem to a min-max formula (known as the Berge–Tutte Formula) for the maximum size of a matching in a general graph. In ...
  41. [41]
    [PDF] 5. Lecture notes on matroid intersection 5.1 Examples
    Mar 30, 2011 · Matchings in a bipartite graph G = (V,E) with bipartition (A, B) do not form the indepen- dent sets of a matroid. However, they can be viewed as ...
  42. [42]
    Arc Routing Problems, Part I: The Chinese Postman Problem
    The optimal solution to this problem can be obtained by finding a minimum weight cut in the graph, ... This model can be solved as a perfect matching problem on ...
  43. [43]
    [PDF] Lecture 4 (based on [3] and [6]) 1 Combinatorial Optimization
    Mar 9, 2015 · Bipartite matching is a basic combinatorial optimization problem arising in many different applications, like: • scheduling - consider ...
  44. [44]
    [PDF] College Admissions and the Stability of Marriage
    College Admissions and the Stability of Marriage. Author(s): D. Gale and L. S. Shapley. Source: The American Mathematical Monthly, Vol. 69, No. 1 (Jan., 1962), ...
  45. [45]
    The Hungarian method for the assignment problem - Kuhn - 1955
    The “assignment problem” is the quest for an assignment of persons to jobs so that the sum of the n scores so obtained is as large as possible.
  46. [46]
    The Assignment Game I: The Core - RAND
    The assignment game is a model for a two-sided market in which a product that comes in large, indivisible units is exchanged for money.
  47. [47]
    Kidney Exchange* | The Quarterly Journal of Economics
    The Quarterly Journal of Economics, Volume 119, Issue 2, May 2004 ... Article contents. Cite. Cite. Alvin E. Roth, Tayfun Sönmez, M. Utku Ünver, Kidney Exchange ...
  48. [48]
    Stable Nash Equilibria in the Gale-Shapley Matching Game - arXiv
    Sep 14, 2015 · In this article we study the stable marriage game induced by the men-proposing Gale-Shapley algorithm. Our setting is standard: all the lists ...
  49. [49]
    The Stable Marriage Problem and School Choice
    There is, however, a W W -optimal stable matching: we simply apply the Gale-Shapley algorithm with the roles interchanged so that women propose to men.