Fact-checked by Grok 2 weeks ago

Google matrix

The Google matrix is a primitive, column-stochastic matrix fundamental to the algorithm, which ranks web pages according to their importance in the World Wide Web's structure. Developed as part of 's by founders and , it models web navigation as a where matrix entries represent transition probabilities between pages linked by . The matrix addresses challenges in the web graph, such as dangling nodes (pages with no outgoing links) and potential rank sinks, by incorporating a random mechanism to ensure reliable convergence. Mathematically, the Google matrix G is constructed as G = \alpha S + (1 - \alpha) \frac{1}{n} \mathbf{e} \mathbf{e}^T, where S is a column-stochastic matrix derived from the adjacency matrix H (with columns normalized for out-degree and dangling node columns replaced by uniform distributions), \alpha (typically 0.85) is the balancing link-following and random jumps, n is the total number of modeled web pages, and \mathbf{e} is the column vector of all ones. This formulation renders G irreducible, aperiodic, and positive, properties that guarantee the existence of a unique —the vector—obtainable as the dominant eigenvector corresponding to eigenvalue 1 via the power iteration method. In practice, the vector assigns a non-negative score to each page proportional to its estimated , aggregating importance from incoming links while mitigating biases from isolated components. The algorithm's iterative scales to billions of pages through techniques, and its influence extends beyond web search to applications in , social networks, and recommendation systems, underscoring the matrix's role in quantifying node centrality in large directed graphs.

Basic Components

Adjacency Matrix

The A for a with n s is defined as an n \times n matrix where the entry A_{ij} = 1 if there is a from j to i, and A_{ij} = 0 otherwise. This representation captures the raw hyperlink structure of networks such as the , treating pages as s and links as edges. The matrix A has non-negative entries by construction, with column sums equal to the out-degrees of the nodes (i.e., the number of outgoing from each ). In large-scale networks like the web, A is highly sparse, as the number of non-zero entries corresponds to the total number of , which is typically much smaller than n^2. To prepare for probabilistic transitions, A is normalized into a column-stochastic form by dividing each column by its out-degree, ensuring column sums equal 1 where possible; dangling nodes (pages with zero out-degree, resulting in zero-sum columns) require special handling, such as uniform redistribution across all nodes. This normalized version transitions to the stochastic matrix S, which models random walks on the graph. For illustration, consider a simple directed graph with two nodes and mutual links: node 1 links to node 2, and node 2 links to node 1. The adjacency matrix is A = \begin{pmatrix} 0 & 1 \\ 1 & 0 \end{pmatrix}, where each column sums to 1, reflecting out-degrees of 1 for both nodes.

Stochastic Matrix

The stochastic matrix S is obtained by normalizing the adjacency matrix of the web graph to form transition probabilities for a random walk. Given an adjacency matrix A where A_{ij} = 1 if page j links to page i and 0 otherwise, S is defined as S = A D^{-1}, with D being the diagonal matrix where each diagonal entry D_{jj} equals the out-degree of page j, or the number of outgoing links from j. This construction yields S_{ij} = \frac{A_{ij}}{D_{jj}} = \frac{1}{\text{out-degree}(j)} if page j links to page i, and 0 otherwise, provided the out-degree is positive. Pages with no outgoing links, termed dangling nodes, result in a column of zeros in S, violating the stochastic property since the column sum is 0 rather than 1. To resolve this, such columns are replaced by a uniform vector with each entry equal to \frac{1}{n}, where n is the total number of pages, thereby distributing the transition probability equally across all pages and ensuring every column sums to 1. The matrix S is column-stochastic, with non-negative entries and each column summing to 1, which encodes the probabilities of moving from the current page j to any linked page i during a random surf on the web. It serves as the transition matrix for a that models unperturbed random surfing behavior, where the surfer follows links uniformly at random. If the underlying is strongly connected, S is irreducible, meaning the chain can reach any state from any other state.

Definition and Construction

The Google Matrix Formula

The Google matrix G, central to the algorithm, is defined by the formula G = \alpha S + (1 - \alpha) E, where S is the row-stochastic derived from the web's structure, \alpha is the , and E is the n \times n matrix whose every entry is $1/n, with n denoting the total number of web pages. The matrix S is obtained by row-normalizing the to account for outgoing links, with adjustments for pages lacking out-links (dangling nodes) to ensure stochasticity. This formulation interprets the random surfer model: the term \alpha S captures the probability \alpha of continuing the walk by following an outgoing hyperlink, while (1 - \alpha) E models the complementary probability of teleporting uniformly to any page, preventing the walk from becoming trapped in disconnected components or rank sinks within the web graph. In web ranking applications, \alpha is typically set between 0.85 and 0.99, which balances link-following behavior with random resets and guarantees that G remains row-stochastic. The row-stochastic property of G follows directly from those of its components: since the rows of both S and E sum to 1, each row sum of G is \alpha \cdot 1 + (1 - \alpha) \cdot 1 = 1, maintaining the essential Markov chain structure for convergence to a unique stationary distribution.

Role of the Damping Factor

The damping factor, denoted as α, serves a critical purpose in the Google matrix by regulating the trade-off between link-following behavior, which facilitates exploration along the web's hyperlink structure, and random teleportations to arbitrary pages, which promote global mixing across the entire graph. This balance is essential for preventing the random surfer model from becoming trapped in dense clusters of mutually linking pages or infinite loops, where probability mass could otherwise accumulate indefinitely without dissipation. Historically, α was introduced to emulate realistic user navigation patterns, capturing the empirical observation that web surfers occasionally abandon their current path—estimated at about a 15% probability per step—to jump randomly to another page. In terms of matrix properties, setting α < 1 ensures that the G = α S + (1 - α) E, where S is the stochastic link matrix and E is the uniform matrix, becomes primitive, rendering it both irreducible (fully connected via paths) and aperiodic (no cyclic partitioning). This primitivity guarantees the existence of a unique stationary distribution, which corresponds to the PageRank vector, independent of starting conditions. However, as α approaches 1, the matrix's spectral gap narrows, leading to slower convergence in power iteration methods used to compute the stationary distribution, often requiring more iterations for numerical stability. The choice of α is largely empirical, with the original PageRank implementation adopting α = 0.85 to strike an optimal balance between ranking accuracy, which favors higher α for stronger emphasis on link structure, and computational efficiency, as lower α accelerates convergence. Sensitivity analyses reveal that variations in α can significantly affect rank stability, potentially causing rank reversals or shifts in the ordering of pages, particularly in networks with uneven connectivity; for instance, small deviations from 0.85 can cause rank reversals, though rankings in real web graphs tend to exhibit relative stability with a consistent core of top-ranked pages, underscoring the need for careful tuning based on the specific corpus.

Mathematical Properties

Eigenvalues and Spectrum

The Google matrix G, being column-stochastic, possesses a largest eigenvalue \lambda_1 = 1 with algebraic multiplicity one. This eigenvalue is simple due to the stochastic nature of G, ensuring a unique stationary distribution in the associated Markov chain. For all other eigenvalues \lambda_i with i > 1, the condition |\lambda_i| < 1 holds, guaranteeing convergence of iterative methods like the power iteration used in computation. These eigenvalues are bounded by the damping factor \alpha, specifically |\lambda_i| \leq \alpha, as G acts as a contraction mapping in the infinity norm. In non-symmetric directed networks such as the web graph, complex eigenvalues are prevalent, forming intricate patterns in the complex plane within the unit disk. The spectral gap, defined as $1 - |\lambda_2|, where \lambda_2 is the second-largest eigenvalue in modulus, governs the convergence rate of the power iteration algorithm. In typical web graphs, this gap is small—often on the order of $1 - \alpha \approx 0.15 for \alpha = 0.85—leading to slow mixing times and requiring numerous iterations for accurate computation. The Perron-Frobenius theorem applies to the Google matrix, which is positive due to the uniform damping term, ensuring that the dominant eigenvalue \lambda_1 = 1 is real, simple, and greater in modulus than all others, with a corresponding strictly positive eigenvector. In large-scale networks with dimension n \gg 1, the power-law structure and communities inherent to web graphs introduce deviations in the eigenvalue distribution, such as patterns reflecting network modularity.

Eigenvectors and Stationary Distribution

The principal eigenvector of the Google matrix G corresponds to the eigenvalue \lambda = 1 and is a left eigenvector v satisfying v G = v, normalized such that \sum_i v_i = 1. This vector represents the stationary probability distribution of the associated , with components v_i interpreted as scores indicating the relative importance of nodes in the network. The right eigenvector for \lambda = 1 is the uniform vector consisting of all entries equal to $1/n, where n is the matrix dimension, reflecting the stochastic nature of G where each column sums to 1. Secondary eigenvectors, associated with eigenvalues |\lambda| < 1, can reveal underlying network structures such as communities or cycles; for instance, they often localize on specific subsets of nodes, highlighting groups with strong internal connectivity like academic disciplines or regional clusters in directed networks. The principal eigenvector is unique and strictly positive component-wise (v_i > 0 for all i), guaranteed by the primitivity of G (ensured by the damping factor making the matrix irreducible and aperiodic), as per the Perron-Frobenius theorem for nonnegative primitive matrices. It is computed via the power method, iterating v^{(k+1)} = v^{(k)} G from an initial v^{(0)} (often uniform), with convergence to the at a rate determined by the $1 - |\lambda_2|, where \lambda_2 is the second-largest eigenvalue modulus.

Examples and Illustrations

Small-Scale Models

To illustrate the construction of the matrix, consider a small 3- ring , where nodes are labeled 1, 2, and 3, with directed links 1 → 2, 2 → 3, and 3 → 1. Each node has out-degree 1, so the column-stochastic S (obtained by normalizing the columns) is the S = \begin{pmatrix} 0 & 0 & 1 \\ 1 & 0 & 0 \\ 0 & 1 & 0 \end{pmatrix}. With \alpha = 0.85, the matrix is G = \alpha S + (1 - \alpha)/3 \cdot \mathbf{1}\mathbf{1}^T, where \mathbf{1} is the all-ones , yielding G = \begin{pmatrix} 0.05 & 0.05 & 0.90 \\ 0.90 & 0.05 & 0.05 \\ 0.05 & 0.90 & 0.05 \end{pmatrix}. The PageRank vector is the principal right eigenvector of G corresponding to eigenvalue 1, normalized to sum to 1; due to the symmetric cycle structure and damping, it is the uniform distribution [1/3, 1/3, 1/3]^T. To compute it via power iteration, start with initial vector r^{(0)} = [0.5, 0.3, 0.2]^T and iterate r^{(k+1)} = G r^{(k)} (the sums remain 1 since G is column-stochastic). The first three iterations are r^{(1)} = \begin{pmatrix} 0.220 \\ 0.475 \\ 0.305 \end{pmatrix}, \quad r^{(2)} = \begin{pmatrix} 0.309 \\ 0.237 \\ 0.454 \end{pmatrix}, \quad r^{(3)} = \begin{pmatrix} 0.280 \\ 0.266 \\ 0.454 \end{pmatrix}, approaching [0.333, 0.333, 0.333]^T (further iterations confirm convergence within machine precision after ~20 steps for this scale). For the undamped case (\alpha = 1), G = S, and power iteration fails to converge, instead cycling periodically due to the eigenvalues of S lying on the unit circle (roots of unity for the cycle). Using the same initial r^{(0)}, r^{(1)} = S r^{(0)} = \begin{pmatrix} 0.200 \\ 0.500 \\ 0.300 \end{pmatrix}, \quad r^{(2)} = \begin{pmatrix} 0.300 \\ 0.200 \\ 0.500 \end{pmatrix}, \quad r^{(3)} = \begin{pmatrix} 0.500 \\ 0.300 \\ 0.200 \end{pmatrix} = r^{(0)}, demonstrating perpetual rotation without damping to ensure aperiodicity and convergence. A dangling node (out-degree 0) requires uniform replacement in S to maintain stochasticity. Consider a 4-node chain graph with nodes 1 → 2 → 3 → 4, where 4 is the dangling sink (no outgoing links). The column-stochastic S has column 4 replaced by [0.25, 0.25, 0.25, 0.25]^T: S = \begin{pmatrix} 0 & 0 & 0 & 0.25 \\ 1 & 0 & 0 & 0.25 \\ 0 & 1 & 0 & 0.25 \\ 0 & 0 & 1 & 0.25 \end{pmatrix}. With \alpha = 0.85, G = \alpha S + (1 - \alpha)/4 \cdot \mathbf{1}\mathbf{1}^T = 0.85 S + 0.0375 \cdot \mathbf{1}\mathbf{1}^T. Power iteration starting from uniform r^{(0)} = [0.25, 0.25, 0.25, 0.25]^T yields r^{(1)} \approx \begin{pmatrix} 0.091 \\ 0.303 \\ 0.303 \\ 0.303 \end{pmatrix}, and converges (after ~30 steps) to approximately [0.117, 0.215, 0.296, 0.372]^T. The damping factor \alpha = 0.85 ensures convergence while balancing link-following (85% probability) against uniform teleportation (15%), resulting in the sink node 4 receiving elevated rank (~37%) from incoming links and redistributed probability, while earlier nodes gain indirectly via uniform replacement and teleports (node 1 ~12%, reflecting dilution along the chain). Without replacement (raw adjacency for column 4 as zeros), probability mass leaks at the sink, causing total rank sum to decay to 0 over iterations, precluding a stationary distribution.

Large-Scale Network Examples

As observed in a 2000 study by Broder et al., the web underlying the matrix in PageRank computations exhibits a power-law degree distribution, with in-degree and out-degree probabilities following P(k) \sim k^{-\gamma} where \gamma \approx 2.1 for in-degrees and \gamma \approx 2.7 for out-degrees, reflecting a scale-free structure dominated by a few high-degree hubs. This also displays a bow-tie , featuring a strongly connected core (SCC) comprising about 28% of nodes, incoming components (IN) that funnel into the SCC (≈21%), outgoing components (OUT) from the SCC (≈21%), and additional tendrils (≈22%) and disconnected regions (≈8%). Recent studies suggest the global bow-tie structure has evolved, with more emphasis on local bow-tie patterns within components. Such characteristics produce an adjacency matrix that is highly sparse, with modern web graphs encompassing roughly $4 \times 10^9 nodes and average degrees around 10, resulting in edge counts on the order of $10^{10} far below the theoretical maximum of $10^{18}. Given the vast scale, exact construction of the dense Google matrix proves impractical; instead, approximations like Monte Carlo methods simulate random surfer walks to estimate the stationary distribution, providing accurate PageRank values for prominent pages with reduced memory demands compared to full matrix operations. Sparse matrix formats, such as compressed sparse row (CSR), enable efficient storage of the transition matrix and iterative computations without densifying the Google matrix. An illustrative early application processed a 1998 Stanford web crawl of 24 million pages using these techniques, demonstrating feasible ranking on period hardware despite the graph's sparsity. In large-scale web graphs, scores correlate strongly with in-degrees, both adhering to power-law tails with similar exponents, yet the \alpha integrates global paths and teleportation, highlighting pages with broader structural significance over mere local link counts. Within the , rankings show heightened sensitivity to \alpha, where shifts from the standard 0.85 value can induce frequent rank reversals, altering perceived page importance. Computational demands include storing the in CSR format, which scales with edge volume—often exceeding $10^{11} for web-scale graphs—typically via distributed systems. for convergence requires hundreds of sparse matrix-vector multiplications to achieve precision, though the expedites this to tens of iterations in practice for graphs over 80 million nodes.

Applications and Extensions

Web Ranking with

The algorithm employs an iterative process to compute the principal eigenvector v of the Google matrix, which assigns an importance score to each based on the structure of hyperlinks. Higher values of v_i for a page i signify greater authority, reflecting the likelihood that a random surfer would visit that page after many steps in a simplified model of . This eigenvector, representing the of the associated , is obtained through repeated matrix-vector multiplications until convergence, typically using the power iteration method. In Google Search, these precomputed PageRank scores are integrated into the ranking of search results by combining them with query-specific relevance measures derived from content analysis. The system evaluates pages not only for their authority via PageRank but also for how well their text matches the user's query, using techniques such as term frequency-inverse document frequency (tf-idf) weighting. The final ranking score for a page is a linear combination of its PageRank and these relevance factors, ensuring that results prioritize both authoritative sources and topical pertinence. This hybrid approach significantly enhances the quality of search outcomes over purely content-based methods. Personalized variants of adapt the algorithm to individual users by modifying the teleportation component of the Google matrix, biasing the random surfer's jumps toward a user-specific set of pages, such as bookmarks or previously visited sites. This shifts the to emphasize pages more relevant to the user's interests, allowing for tailored search rankings without altering the core link structure analysis. Such adaptations were proposed in the original to accommodate diverse surfing behaviors. Despite its effectiveness, PageRank exhibits vulnerabilities to manipulation through link spam, particularly via link farms—networks of low-quality pages created solely to interlink and artificially inflate scores for targeted sites. These tactics exploit the reliance on incoming links, leading to undeserved high rankings for spammed pages and degrading overall search quality. Subsequent refinements, including trust-based modifications, have mitigated these issues by incorporating additional signals of page credibility.

Analyses in Other Networks

The Google matrix has been extended to biological networks, where it models evolutionary dynamics and interaction flows in systems such as protein interaction graphs and food webs. In protein interaction networks, the reduced Google matrix algorithm analyzes effective interactions between subsets of proteins, accounting for indirect pathways through the larger network; for instance, in the bi-functional SIGNOR database of signaling pathways, it reveals hidden causal relations and ranks protein influence beyond direct links. Similarly, in fibrosis-related protein-protein interaction networks derived from data, the reduced Google matrix identifies key regulatory proteins by quantifying their indirect impacts on disease progression. For food webs and ecological systems, the Google matrix assesses in mutualism-competition models governed by Lotka-Volterra dynamics, where it decomposes interaction matrices to predict coexistence; a 2016 study on structured ecological networks found that feasibility (positive abundances) is lost before as mutualism strength increases, with empirical mutualistic networks (e.g., plant-pollinator webs with connectance q ≈ 0.13) showing alignment to randomized thresholds. In social and citation networks, the Google matrix enables ranking of influence while adapting the damping factor α to capture community structures. For citation networks, analysis of the Physical Review articles (over 400,000 papers) using the Google matrix spectrum reveals PageRank vectors that prioritize seminal works, with the distribution in the PageRank-CheiRank plane highlighting bidirectional citation flows and outperforming simple indegree metrics for impact assessment. In social platforms like Twitter, the Google matrix constructed from the 2009 network (41 million users) quantifies user influence via PageRank and CheiRank, where adjusting α (typically 0.85) accounts for sparse community ties and retweet asymmetries, identifying top influencers with eigenvector centralities that reflect real-world virality patterns. For ranking scientists via arXiv links, adaptations of the Google matrix in co-citation graphs adjust α to emphasize disciplinary communities, yielding prestige scores correlated with h-index but enhanced by indirect collaboration paths. Generalizations of the Google matrix extend to directed weighted graphs and time-varying links, preserving key properties like Perron-Frobenius eigenvectors for measures. In directed weighted networks, the matrix incorporates edge weights into the stochastic transition operator, enabling analysis of flow imbalances; a comprehensive review demonstrates its application to economic trade networks, where weighted links represent transaction volumes, and the leading eigenvector approximates in undirected projections. Recent extensions leverage the Google matrix in recommender systems and epidemic spreading models, with damping factor α optimized for network diameter to enhance prediction accuracy. In recommender systems, the reduced Google matrix has been explored to infer user-item affinities in sparse graphs by propagating indirect preferences.

Historical Development

Origins and Invention

The development of the Google matrix emerged from early explorations in web graph analysis and link-based ranking. Prior to its invention, the , proposed by in 1998, represented a foundational method that employed separate and matrices to evaluate the importance of web pages based on incoming and outgoing . This approach highlighted the potential of hyperlink structures to infer page authority, influencing later algorithms by shifting focus from content keywords to relational graph properties. In 1996, Ph.D. students and launched the BackRub project under the Stanford Digital Library Initiative, with the goal of systematically downloading and indexing the rapidly expanding while prioritizing hyperlinks as indicators of page relevance over textual content alone. Motivated by the limitations of existing search engines, which struggled with and low-quality results, they sought to exploit the web's hypertextual structure to generate more accurate rankings. The core idea of the Google matrix was first articulated in Brin and Page's 1998 paper, where they introduced the random surfer model to compute page importance, incorporating a damping mechanism to account for users occasionally abandoning their browsing path and jumping to a random page. This model transformed the web's link graph into a , with the damping factor ensuring convergence and mimicking realistic navigation behavior. The invention was protected through U.S. Patent 6,285,999, filed on January 9, 1998, and issued on September 4, 2001, to Lawrence Page, which detailed a method for ranking nodes in linked databases using a modified derived from citations. The patent described the matrix as a combination of the web's adjacency structure and uniform teleportation probabilities, establishing the foundational framework for scalable web ranking.

Key Publications and Evolution

The foundational description of the Google matrix appeared in the 1998 paper "The Anatomy of a Large-Scale Hypertextual Search Engine" by and Lawrence Page, presented at the 7th International Conference. This work introduced the stochastic matrix G = \alpha S + (1 - \alpha) E, where S is the column-stochastic transition matrix derived from web hyperlinks, E is the uniform teleportation matrix, and the damping factor \alpha = 0.85 was selected to balance link-following with random jumps, mimicking typical user navigation patterns. A key follow-up publication, the 1999 Stanford "The PageRank Citation Ranking: Bringing Order to the Web" by Lawrence Page, , , and , formalized the computation of the Google matrix's principal eigenvector as the vector, emphasizing efficient methods for large-scale implementation. This report solidified the mathematical framework, proving convergence under the matrix's Perron-Frobenius properties and highlighting its superiority over content-based ranking for web-scale data. Subsequent evolutions included the 2002 introduction of personalized PageRank variants, such as topic-sensitive PageRank by Taher H. Haveliwala, which modifies the teleportation vector to incorporate user-specific or topic-based biases, enabling context-aware ranking without altering the core matrix structure. In the 2010s, integrations with machine learning enhanced spam detection; for instance, the 2012 MaxRank algorithm proposed by Olivier Fercoq et al. uses optimization and seed sets to detect link spam and apply PageRank demotion. By 2025, the fundamental Google matrix formulation remained unchanged in core search applications, but hybrid models emerged, fusing it with graph neural networks for AI-driven search, as in the 2022 Personalized PageRank Graph Neural Network (PPRGNN) that extends propagation to infinite depths via neural approximations. The Google matrix's academic impact extends beyond web search, with extensive citations in random matrix theory for analyzing eigenvalue distributions in sparse stochastic systems. In physics, by the 2010s, it inspired extensions to and chaotic dynamics, notably through the reduced Google matrix framework for modeling open and network flows, as reviewed in the 2015 work by Leonardo Ermann, Klaus M. Frahm, and Dima L. Shepelyansky.

References

  1. [1]
    [PDF] The PageRank Citation Ranking: Bringing Order to the Web
    Jan 29, 1998 · This paper describes PageRank, a method for rating Web pages objectively and mechanically, effectively measuring the human interest and.
  2. [2]
  3. [3]
    [PDF] An Introduction to Google's PageRank - Search Engine Rankings
    Properties of Google matrix. G = αS + (1 − α)1/n e eT. G is ... table: Langville and Meyer. ▻ H is the largest component to be stored. ▻ if ...
  4. [4]
    None
    ### Summary of PageRank Algorithm Description
  5. [5]
  6. [6]
  7. [7]
    [PDF] Google PageRank - Australian Mathematical Sciences Institute
    Brin and Page suggested that the matrix equation v = v H given in the previous section should be solved iteratively, that is, by calculating a sequence of ...
  8. [8]
    [PDF] PageRank as a Function of the Damping Factor∗ - Sebastiano Vigna
    The original value suggested by Brin and Page (α = 0.85) is the most common choice. Intuitively, 1 − α is an amount of ranking that we agree to give uniformly ...
  9. [9]
    [PDF] The Anatomy of a Large-Scale Hypertextual Web Search Engine
    Abstract. In this paper, we present Google, a prototype of a large-scale search engine which makes heavy use of the structure present in hypertext.
  10. [10]
    Choose the damping, choose the ranking? - ScienceDirect.com
    This paper addresses the fundamental question of how the ranking induced by PageRank can be affected by variations of the damping factor.
  11. [11]
    [PDF] The Second Eigenvalue of the Google Matrix - Stanford NLP Group
    Abstract. We determine analytically the modulus of the second eigenvalue for the web hyperlink matrix used by Google for computing PageRank. Specifically,.
  12. [12]
    [PDF] the google spectrum - UChicago Math
    Dec 14, 2017 · The adjacency matrix is both real and symmetric, by definition. Since the la- beling of a graph's vertices is arbitrary, properties of the ...
  13. [13]
    Google matrix - Scholarpedia
    The Google matrix belongs to the class of Perron-Frobenius operators which appear in the description of dynamical chaotic systems.
  14. [14]
    Spectral properties of the Google matrix of the World Wide Web and ...
    Feb 17, 2010 · We also determine specific properties of eigenstates of the Google matrix, including the PageRank. The fidelity of the PageRank is proposed as a ...
  15. [15]
    [PDF] The PageRank Citation Ranking: Bringing Order to the Web
    Jan 29, 1998 · This paper describes PageRank, a method for rating Web pages objectively and mechanically, effectively measuring the human interest and.Missing: adjacency | Show results with:adjacency
  16. [16]
  17. [17]
    WorldWideWebSize.com | The size of the World Wide Web (The ...
    The Indexed Web contains at least 3.98 billion pages (Wednesday, 15 January, 2025). The Dutch Indexed Web contains at least 2042.74 million pages (Wednesday, 15 ...
  18. [18]
    Monte Carlo Methods in PageRank Computation: When One ...
    We present a simple algorithm for computing the PageRank (stationary distribution) of the stochastic Google matrix G. The algorithm lumps all dangling nodes ...
  19. [19]
    The Anatomy of a Large-Scale Hypertextual Web Search Engine
    In this paper, we present Google, a prototype of a large-scale search engine which makes heavy use of the structure present in hypertext.
  20. [20]
    In-Degree and PageRank: Why Do They Follow Similar Power Laws?
    They have shown that the PageRank of the top 10% of the nodes always follows a power law with the same exponent independent of the value of the damping factor.
  21. [21]
    PageRank and rank-reversal dependence on the damping factor
    EFFECTS OF DAMPING FACTOR ON PAGERANK. We first examine the responses of PR ... When the damping factor changes from 0.05 to 0.99, we observe how the rank changes ...
  22. [22]
    [PDF] Adaptive Methods for the Computation of PageRank
    The Power Method on a web data set of over 80 million pages converges in about 50 iterations. Improving on this already fast convergence rate is a difficult ...Missing: CSR | Show results with:CSR
  23. [23]
    Google matrix analysis of bi-functional SIGNOR network of protein ...
    Sep 24, 2019 · This group of proteins is analyzed by the reduced Google matrix algorithm which allows to determine the effective interactions between them due ...
  24. [24]
    Fibrosis Protein-Protein Interactions from Google Matrix Analysis of ...
    The reduced Google matrix G R of the fibrosis network describes effective interactions between N r nodes, taking into account all direct and indirect ...
  25. [25]
    The Google matrix controls the stability of structured ecological and ...
    Sep 30, 2016 · For competition communities, the ecological interaction matrix A may be decomposed into three components: a background network of uniform all-to ...
  26. [26]
    Google matrix of the citation network of Physical Review | Phys. Rev. E
    May 28, 2014 · We study the statistical properties of spectrum and eigenstates of the Google matrix of the citation network of Physical Review for the ...Missing: graph | Show results with:graph
  27. [27]
    [1207.3414] Google matrix of Twitter - arXiv
    Jul 14, 2012 · We construct the Google matrix of the entire Twitter network, dated by July 2009, and analyze its spectrum and eigenstate properties including ...
  28. [28]
    [PDF] PageRank for ranking authors in co-citation networks - arXiv
    PageRank, used for ranking web pages, is applied to author co-citation networks, giving higher weights to highly cited publications and those cited by highly ...
  29. [29]
    Google matrix analysis of directed networks | Rev. Mod. Phys.
    Nov 2, 2015 · This review describes the Google matrix analysis of directed complex networks demonstrating its efficiency using various examples.<|control11|><|separator|>
  30. [30]
    Estimating time-varying networks for high-dimensional time series
    We explore time-varying networks for high-dimensional locally stationary time series, using the large VAR model framework with both the transition and ...
  31. [31]
    Predicting the Speed of Epidemics Spreading in Networks
    Feb 12, 2020 · A new analysis predicts the speed at which an infectious disease spreads to specific individuals in a network.
  32. [32]
    The anatomy of a large-scale hypertextual Web search engine
    In this paper, we present Google, a prototype of a large-scale search engine which makes heavy use of the structure present in hypertext.
  33. [33]
    US6285999B1 - Method for node ranking in a linked database
    A method assigns importance ranks to nodes in a linked database, such as any database of documents containing citations, the world wide web or any other ...
  34. [34]
    [PDF] The anatomy of a large-scale hypertextual Web search engine '
    This paper addresses this question of how to build a practical large-scale system which can exploit the additional information present in hypertext. Also we ...
  35. [35]
    Topic-sensitive PageRank | Proceedings of the 11th international ...
    Topic-sensitive PageRank. Author: Taher H. Haveliwala. Taher H. Haveliwala ... Published: 07 May 2002 Publication History. 912citation5,219Downloads.
  36. [36]
    [PDF] PageRank optimization applied to spam detection - arXiv
    Mar 7, 2012 · ABSTRACT. We give a new link spam detection and PageRank demotion algorithm called MaxRank. Like TrustRank and AntiTrust-.
  37. [37]
    Transforming PageRank into an Infinite-Depth Graph Neural Network
    Sep 19, 2022 · Adopting this idea, we propose the Personalized PageRank Graph Neural Network (PPRGNN), which extends the graph convolutional network to an ...