Fact-checked by Grok 2 weeks ago

Algorithmic paradigm

An algorithmic paradigm is a fundamental strategy or pattern for designing algorithms to solve computational problems, offering a reusable that addresses a broad class of similar challenges by emphasizing common techniques such as , optimization, or . These paradigms guide algorithm development by providing structured methods to break down problems, analyze efficiency, and ensure correctness, often drawing from mathematical principles like or . Key algorithmic paradigms include divide-and-conquer, which recursively divides a problem into smaller subproblems, solves them independently, and combines the results—for instance, in , which achieves O(n log n) for . Another prominent paradigm is dynamic programming, used for optimization problems with overlapping subproblems and optimal substructure, such as computing the in O(mn) time by building a table of solutions to subproblems. Greedy algorithms make locally optimal choices at each step to approximate a global optimum, exemplified by for minimum spanning trees, running in O(E log V) time with a union-find . Additional paradigms encompass brute-force methods, which exhaustively enumerate all possibilities to guarantee an exact solution, though often at high computational cost, as in generating all permutations for the traveling salesman problem. Randomized algorithms incorporate randomness to achieve expected efficiency, such as randomized with average O(n log n) performance. Reduction transforms one problem into another known problem, facilitating solutions or complexity proofs, like reducing 3-SAT to the in studies. These paradigms are foundational in , enabling efficient algorithm design across domains like , , and optimization, while their analysis often relies on asymptotic notations such as Big-O to evaluate time and .

Definition and Fundamentals

Definition

An algorithmic paradigm is a generic model or that underlies the design of a class of , focusing on high-level strategies for problem-solving rather than specific implementations. It represents common concepts and ideas behind algorithm , providing a structured approach to tackling computational problems. This distinguishes algorithmic paradigms from individual algorithms, which are concrete procedures for solving specific instances of problems, as paradigms offer abstract templates applicable across multiple related problems. Unlike programming paradigms, which emphasize styles of code organization and execution (such as imperative or ), algorithmic paradigms center on the logical strategies for devising solutions, independent of how they are coded. Algorithmic paradigms play a key role in guiding efficiency analysis by enabling the evaluation of time and , often expressed in , to compare solution viability for large inputs.

Key Characteristics

Algorithmic paradigms serve as reusable blueprints for solving a wide array of computational problems, enabling designers to apply proven strategies across diverse scenarios rather than starting from scratch for each instance. This reusability stems from the paradigms' of common structural patterns, such as or incremental decision-making, which can be adapted to problems sharing similar properties, thereby significantly reducing development time and effort in algorithm creation. A core emphasis of algorithmic paradigms lies in optimizing , particularly through the and resolution of recurrence relations that govern their time and complexities. For instance, in the divide-and-conquer paradigm, the provides a systematic method to solve recurrences of the form T(n) = aT(n/b) + f(n), where a \geq 1, b > 1, and f(n) represents the cost of dividing and combining subproblems, allowing for the determination of asymptotic bounds without exhaustive case-by-case computation. This focus on ensures that paradigms guide the selection of approaches that minimize resource usage while maintaining correctness. Algorithmic paradigms strike a balance between generality and specificity, offering broad frameworks that apply to multiple problem classes while permitting tailored adaptations to unique constraints or objectives. This duality allows paradigms to provide versatile templates—such as selection for optimization problems—that can be specialized with domain-specific heuristics or data structures, enhancing both applicability and without sacrificing foundational rigor. The choice of paradigm influences the criteria employed to assess algorithmic , with metrics like worst-case (bounding the maximum runtime over all inputs), average-case (averaging over probable input distributions), and (smoothing costs over a sequence of operations) tailored to the paradigm's operational characteristics. For example, is particularly useful for paradigms involving dynamic data structures, where occasional expensive operations are offset by frequent inexpensive ones, providing a more realistic view of overall efficiency than isolated worst-case bounds.

Historical Development

Early Concepts

The origins of algorithmic paradigms trace back to ancient mathematics, where systematic procedures for solving problems emerged without the aid of mechanical devices. One of the earliest examples is the , documented by the Greek mathematician in his seminal work around 300 BCE. This method computes the of two integers by iteratively replacing the larger number with the remainder of the division of the two numbers, effectively reducing the problem size until the remainder is zero; the last non-zero remainder is the GCD. As a precursor to divide-and-conquer strategies, it demonstrates a foundational approach to problem decomposition, where a complex task is broken into simpler subproblems through repeated application of a basic operation, highlighting the timeless efficiency of recursive reduction in computation. In the 9th century, the Persian mathematician Muhammad ibn Musa developed systematic procedures for solving linear and quadratic equations in his treatise Al-Kitab al-Mukhtasar fi Hisab al-Jabr wal-Muqabala, laying foundational algorithmic methods that influenced later computational thinking. In the , the advent of computing devices marked a transition toward more structured algorithmic thinking, influenced by the need for automated calculation in scientific and engineering contexts. Babbage's design for the , conceived in the 1830s and detailed through the 1840s, introduced key concepts such as punched cards for programming instructions and a separation between (the "store") and (the "mill"), enabling systematic decomposition of problems into programmable sequences. This architecture allowed for conditional branching and looping, facilitating the breakdown of computations into modular steps that could be executed non-sequentially based on intermediate results, thus laying early groundwork for general-purpose algorithmic frameworks beyond fixed-function machines like his earlier . Augmenting Babbage's vision, Ada Lovelace's extensive notes published in 1843 on the emphasized its potential for universal computation. She articulated that the machine was not limited to numerical tabulation but could manipulate symbols and perform abstract operations, such as composing and resolving complex algebraic forms, likening it to a weaving patterns from logical threads. Lovelace's insights, including her algorithm for computing Bernoulli numbers, underscored the recognition of general methods applicable to non-numerical domains, foreshadowing the versatility of algorithmic paradigms in handling diverse symbolic manipulations. A pivotal formalization occurred concurrently with , introduced by in his 1847 treatise The Mathematical Analysis of Logic. Boole transformed traditional Aristotelian logic into an algebraic system using symbols to represent propositions, where operations like addition denoted disjunction and multiplication , enabling the mechanical resolution of logical inferences through . This framework provided a rigorous basis for logical paradigms, shifting from enumerative syllogisms to algorithmic manipulation of classes and relations, which would later underpin binary decision-making in computational systems.

20th Century Advancements

The foundations of algorithmic paradigms in the were profoundly shaped by Alan Turing's 1936 introduction of the universal , which formalized the concept of and provided a theoretical model for understanding the limits and structures of algorithmic processes. This machine, capable of simulating any other , established the basis for evaluating the efficiency and applicability of various paradigm models, influencing subsequent developments in algorithm design by emphasizing the role of step-by-step computation in problem-solving. In the 1950s and 1960s, key paradigms emerged through targeted innovations in optimization and . Richard Bellman developed dynamic programming in 1953 as a method for solving complex multistage decision problems by breaking them into overlapping subproblems, with the approach formalized in his early report. The divide-and-conquer paradigm had earlier gained prominence through algorithms like mergesort, first proposed by in 1945 and detailed in a 1948 report by Goldstine and , which recursively divides data into halves before merging sorted subarrays to achieve efficient . Donald Knuth's multivolume series , beginning with Volume 1 in 1968 and extending to Volume 3 () in 1973, offered a systematic classification of algorithmic paradigms, analyzing their structures, efficiencies, and implementations in depth. Knuth's work categorized techniques such as brute-force, divide-and-conquer, and greedy methods within concrete applications like and , providing rigorous proofs and examples that standardized paradigm evaluation in . The 1970s and 1980s saw the introduction and scrutiny of paradigms like greedy algorithms and amid the era, initiated by Stephen Cook's 1971 paper demonstrating that the satisfiability problem is NP-complete. This era highlighted the limitations of these paradigms for intractable problems, as greedy approaches often yield suboptimal approximations for optimization tasks, while enables exhaustive searches but faces exponential time complexities on NP-hard instances, prompting shifts toward and approximation strategies.

Classification of Paradigms

By Design Strategy

Algorithmic paradigms can be classified based on their strategies, which outline the fundamental approaches to solving computational problems. This emphasizes the for developing algorithms, focusing on how problems are approached and resolved rather than the underlying computational resources or models. A prominent groups these strategies into exhaustive, , and optimization categories, providing a structured way to understand and select appropriate techniques for different problem types. Exhaustive strategies involve systematically exploring all possible solutions or a significant portion of the search space to guarantee finding an optimal or correct answer. Brute-force methods represent the simplest form, generating and checking every potential solution without pruning, making them complete but often inefficient for large inputs; for instance, they are used in small-scale problems like solving the traveling salesman problem by evaluating all routes. extends this by building partial solutions incrementally and abandoning unproductive paths early through constraint checking, thus avoiding full enumeration while remaining exhaustive in principle; a classic example is the , where invalid board placements are discarded during recursive placement attempts. These strategies are particularly suited to combinatorial problems where is essential, though their practicality diminishes with problem size due to in possibilities. Decomposition strategies break down complex problems into simpler subproblems, solving them recursively or iteratively and then combining the results. Divide-and-conquer partitions the problem into multiple independent subproblems of roughly equal size, solves each separately, and merges the solutions, as seen in where an array is recursively divided, sorted, and merged. Decrease-and-conquer, in contrast, reduces the problem to a single smaller instance (often by removing or simplifying one element), solves it, and extends the solution to the original; variable-size decrease examples include , which builds a sorted list by incrementally placing elements, while fixed-size variants like Euclid's algorithm for GCD repeatedly reduce the input pair. Both approaches leverage to manage complexity, promoting efficiency through modular problem-solving. Transform-and-conquer, another decomposition variant, transforms the problem into a more convenient form before solving, such as balancing a or using for linear systems. Optimization strategies aim to find high-quality solutions to problems seeking minima, maxima, or optimal selections, often trading for . algorithms make irrevocable local optimal choices at each step, hoping they lead to a global optimum; for example, for minimum spanning trees selects the smallest edge that does not form a cycle until the tree is complete. Dynamic programming addresses global optima by solving overlapping subproblems systematically, storing intermediate results to avoid recomputation—either top-down with or bottom-up via tabulation—as in the computation where subproblem values are cached. While methods are faster but may yield suboptimal results (e.g., in some variants), dynamic programming ensures optimality for suitable problems like knapsack but at higher space and time costs. An example within strategies further groups them by implementation style: recursive strategies like divide-and-conquer and rely on function calls to handle subproblems; iterative ones, such as certain or brute-force implementations, use loops for sequential processing; and randomized strategies incorporate probability for approximation or efficiency, as in methods for exhaustive-like searches. This layering highlights how design strategies intersect with computational models, such as sequential versus probabilistic execution.

By Computational Model

Algorithmic paradigms can be classified based on the underlying computational models they assume, which dictate the resources, execution behavior, and environmental interactions available to the algorithm. A primary distinction lies between deterministic and non-deterministic models. In deterministic models, such as the standard deterministic Turing machine (DTM), the computation follows a unique path for any given input, producing a fixed output without ambiguity. This model underpins paradigms where predictability is essential, ensuring that the same input always yields the same result through sequential state transitions governed by a finite set of rules. In contrast, non-deterministic models, exemplified by the non-deterministic Turing machine (NTM), allow multiple possible transitions from a given state, effectively branching the computation to explore parallel possibilities in a single step. This framework supports paradigms that handle uncertainty or search spaces efficiently in theory, as NTMs characterize the complexity class NP, where verification is polynomial-time but solving may require exploring exponential paths. For instance, paradigms like backtracking can be analyzed under NTM models to assess worst-case behaviors, though practical implementations often simulate nondeterminism deterministically at higher cost. Non-deterministic models are particularly useful for proving lower bounds and understanding inherent computational hardness, as they abstract away the need for explicit branching mechanisms. Another key classification separates sequential and parallel computational models, reflecting differences in how computations exploit concurrency. Sequential models, based on single-processor Turing machines or , assume one computation thread, making them suitable for paradigms like dynamic programming where state dependencies enforce linear progression. Parallel models extend this by allowing simultaneous operations across multiple processors. The model, introduced as an idealized shared-memory architecture, formalizes this by enabling synchronous access to a common memory by p processors, with variants like Concurrent Read Concurrent Write (CRCW) or Exclusive Read Exclusive Write (EREW) to handle contention. PRAM-based paradigms adapt sequential strategies, such as divide-and-conquer, to parallel settings; for example, merging in mergesort can be parallelized to achieve logarithmic time with linear processors under the EREW PRAM model. This classification highlights how parallel paradigms reduce from O(n log n) to O(log n) for certain problems, assuming idealized without communication overhead. Resource-bounded models further classify paradigms by imposing constraints on computational resources beyond time and space, often to address intractability in practical settings. Fixed-parameter tractability (FPT) emerges as a core concept in , where run in time f(k) * n^c, with k as a small , c constant, and f arbitrary but independent of input size n. This model supports paradigms that exploit structural parameters, like in graph algorithms, to achieve tractability for otherwise NP-hard problems. Downey and Fellows established the foundational theory, defining FPT as a generalization of polynomial-time solvability and introducing completeness under parameterized reductions for classes like W and W. For example, admits an FPT algorithm with k as the solution size, solving it in 1.2738^k * n^{O(1)} time, demonstrating how resource bounds enable efficient computation when parameters are bounded. This approach contrasts with classical complexity by slicing problems into tractable kernels, emphasizing preprocessing to reduce instance size before applying exponential-time methods on the parameter. Hybrid models integrate standard Turing machines with external aids, such as oracles, to explore paradigms that approximate solutions under computational limitations. An oracle Turing machine augments a base TM with a query tape to an oracle set A, allowing instantaneous decision on membership in A, thus modeling relativized computation. In approximation paradigms, oracles facilitate the study of classes like APX, where polynomial-time approximation within a constant factor is achievable; for instance, an oracle for the traveling salesman problem can enable PTAS (polynomial-time approximation schemes) by providing near-optimal subproblem solutions. Papadimitriou and Yannakakis formalized such hierarchies for approximation classes, including MAX-SNP completeness under L-reductions. This hybrid framework reveals limitations of derandomization and approximation preservation, proving that P ≠ NP relative to oracles where approximation is hard. By combining deterministic computation with idealized oracles, these models underpin paradigms for optimization under uncertainty, such as local search with approximate oracles for set cover.

Core Paradigms

Brute-Force

The brute-force paradigm, also known as exhaustive search, is an algorithmic strategy that solves problems by systematically generating and evaluating every possible candidate solution until the correct or optimal one is identified. This approach relies on direct enumeration based on the problem's definition, without employing any optimization or shortcuts, making it a straightforward method applicable to any decidable problem. A classic example is the traveling salesman problem (TSP), where the goal is to find the shortest route visiting each city exactly once and returning to the start. In the brute-force implementation, all possible permutations of the cities are generated—totaling n! for n cities—and each permutation's total distance is computed and compared to identify the minimum. This exhaustive checking ensures the exact optimal solution but highlights the paradigm's core mechanism of complete exploration. The of brute-force algorithms for combinatorial problems like TSP is typically O(n!), reflecting the growth in the number of candidates to evaluate. is generally O(1) if solutions are generated iteratively without storage, or O(n) if permutations or paths are held in memory during computation. Brute-force methods are suitable for problems with small input sizes, such as n \leq 10 in permutation-based tasks, where computation remains feasible on modern hardware. They also serve as a correctness to validate more efficient algorithms by comparing outputs on small instances. However, the paradigm's primary limitation is its exponential time complexity, which causes rapid degradation in performance as input size increases; for instance, TSP with n=20 requires over 2.4 quintillion evaluations, rendering it impractical for large-scale applications. For a concrete illustration outside , consider brute-force string matching, which searches for occurrences of a pattern string P of length m in a text string T of length n by sliding P across T and comparing characters position by position. The following outlines the process:
for i from 0 to n - m
    j = 0
    while j < m and T[i + j] == P[j]
        j = j + 1
    if j == m
        report match at position i
This naive approach has a worst-case time complexity of O(nm).

Divide-and-Conquer

The divide-and-conquer paradigm is a fundamental approach in algorithm design where a problem is recursively broken down into smaller, non-overlapping subproblems of the same form, solved independently, and then their solutions are combined to yield the overall solution. This strategy typically involves three key steps: divide, which partitions the problem into subproblems; conquer, which recursively solves the subproblems until they reach base cases that can be solved directly (often in constant time); and combine, which merges the subproblem solutions into a solution for the original problem. The paradigm is particularly effective for problems that exhibit a natural recursive structure, allowing for efficient decomposition without requiring subproblem overlap or local optimization choices. A classic example of divide-and-conquer is mergesort, an efficient sorting algorithm that divides an array of n elements into two halves, recursively sorts each half, and then merges the sorted halves in linear time. The recurrence relation for mergesort's time complexity is T(n) = 2T(n/2) + O(n), which solves to O(n \log n) in both the average and worst cases, providing a stable and predictable performance superior to quadratic brute-force sorting methods. Another representative example is binary search, which locates a target element in a sorted array by repeatedly dividing the search interval in half and discarding the portion that cannot contain the target, reducing the problem size exponentially until the base case is reached. This yields a time complexity of O(\log n), making it vastly more efficient than linear search for large datasets. To analyze the time complexity of divide-and-conquer algorithms, the Master Theorem provides a straightforward method for solving recurrences of the form T(n) = a T(n/b) + f(n), where a \geq 1 is the number of subproblems, b > 1 is the factor by which the input size is divided, and f(n) represents the cost of the divide and combine steps. The theorem classifies solutions into three cases based on the growth rate of f(n) relative to n^{\log_b a}: if f(n) = O(n^{\log_b a - \epsilon}) for some \epsilon > 0, then T(n) = \Theta(n^{\log_b a}); if f(n) = \Theta(n^{\log_b a} \log^k n) for k \geq 0, then T(n) = \Theta(n^{\log_b a} \log^{k+1} n); and if f(n) = \Omega(n^{\log_b a + \epsilon}) with the regularity condition a f(n/b) \leq c f(n) for some c < 1 and large n, then T(n) = \Theta(f(n)). For mergesort, with a=2, b=2, and f(n) = O(n), it falls into the second case, confirming the O(n \log n) bound. One key advantage of the divide-and-conquer paradigm is its inherent parallelizability, as subproblems can often be solved independently across multiple processors, making it suitable for distributed computing environments. Additionally, the recursive decomposition simplifies complex problems into manageable parts, facilitating easier implementation and analysis. However, a notable disadvantage is the overhead from recursive calls, which can lead to increased stack space usage—up to O(n) in the worst case for unbalanced divisions—and potential performance degradation due to function call costs in deep recursion.

Greedy Algorithms

Greedy algorithms constitute a design paradigm in which the solution is constructed by repeatedly selecting the locally optimal choice at each step, with the intention of achieving a global optimum. This approach hinges on two fundamental properties: the , asserting that a locally optimal choice can be part of a globally optimal solution, and the , indicating that the optimal solution incorporates optimal solutions to subproblems. These properties ensure that the algorithm's decisions do not compromise the overall optimality, allowing for efficient construction without revisiting prior choices. A prominent example is for computing the minimum spanning tree of an undirected, connected, weighted graph. The algorithm sorts all edges in non-decreasing order of weight and iteratively adds the smallest edge that connects two distinct components in the forest, avoiding cycles through . This greedy selection guarantees the minimum total weight, as proven by the matroid structure of graphic forests. Another illustrative application is , used for lossless data compression. The method constructs a prefix code by building a binary tree where each internal node represents a merge of the two lowest-probability symbols or subtrees, assigning shorter codes to more frequent symbols. This results in an optimal variable-length code that minimizes the expected code length, achieving the entropy bound for uniquely decodable codes. To establish optimality, proofs for greedy algorithms often employ the exchange argument. This technique demonstrates that if an optimal solution differs from the greedy one, edges or choices can be swapped iteratively—replacing a non-greedy choice with the greedy alternative—without increasing the total cost, eventually yielding the greedy solution. Such arguments are particularly effective for problems like minimum spanning trees, where they show equivalence between any optimal solution and the . Despite their efficiency, greedy algorithms are limited to problems exhibiting the requisite properties; absent these, they may yield suboptimal results. For example, the , where items can be divided, admits a greedy solution by sorting items by value-to-weight ratio and filling the knapsack sequentially, achieving optimality due to the continuous nature. In contrast, the , prohibiting fractions, lacks the greedy choice property, as selecting the highest-ratio item first can miss the global optimum, necessitating instead.

Dynamic Programming

Dynamic programming is an algorithmic paradigm that solves optimization problems by breaking them into overlapping subproblems, computing solutions to these subproblems once, and storing them to avoid redundant calculations, thereby achieving efficient polynomial-time solutions for otherwise exponential problems. The method was developed by in the early 1950s at the to address multistage decision processes in . It applies to problems with two fundamental properties: optimal substructure, where the optimal solution to the overall problem incorporates optimal solutions to its subproblems, and overlapping subproblems, where the same subproblems recur multiple times in a recursive decomposition. These properties enable dynamic programming to extend divide-and-conquer techniques by caching results, transforming recursive calls into a directed acyclic graph of dependencies. Bellman's principle of optimality underpins optimal substructure, stating that an optimal policy has the property that whatever the initial state and initial decision are, the remaining decisions must constitute an optimal policy with regard to the state resulting from the first decision. Overlapping subproblems ensure that without caching, naive recursion would recompute the same values exponentially many times, but with storage, the computation becomes linear in the number of unique subproblems. Dynamic programming implementations follow two primary approaches: top-down memoization and bottom-up tabulation. Top-down memoization starts with the original problem and recursively solves subproblems, using a map or array to store and retrieve previously computed results, ensuring each subproblem is solved at most once. Bottom-up tabulation iteratively fills a table from base cases to the full problem, relying on the dependency order to build solutions progressively. Both achieve the same asymptotic efficiency but differ in implementation: memoization naturally handles sparse subproblems, while tabulation uses contiguous storage for dense cases. A representative example is computing the nth Fibonacci number, defined by the recurrence F_n = F_{n-1} + F_{n-2} for n \geq 2, with base cases F_0 = 0 and F_1 = 1. In bottom-up tabulation, initialize an array dp of size n+1 with dp[0] = 0 and dp[1] = 1, then for i = 2 to n, set \text{dp} = \text{dp}[i-1] + \text{dp}[i-2]; the result is dp[n], running in O(n) time and space. This avoids the O(2^n) time of naive recursion by resolving the overlapping subproblems F_k for k < n. The longest common subsequence (LCS) problem, which finds the longest subsequence present in two given sequences, exemplifies for string alignment. Let X = x_1 x_2 \dots x_m and Y = y_1 y_2 \dots y_n; define C[i,j] as the LCS length of the first i characters of X and first j of Y. The recurrence is: C[i,j] = \begin{cases} 0 & \text{if } i = 0 \text{ or } j = 0 \\ C[i-1,j-1] + 1 & \text{if } i > 0, j > 0, x_i = y_j \\ \max(C[i-1,j], C[i,j-1]) & \text{if } i > 0, j > 0, x_i \neq y_j \end{cases} A bottom-up table fills row-by-row in O(mn) time and space, with the solution at C[m,n]. This captures , as the LCS of prefixes builds from smaller aligned or skipped prefixes. The 0-1 maximizes the total value of items selected for a knapsack of W, where each of n items has weight w_i and value v_i, and each item can be taken at most once. Define C(i,w) as the maximum value using the first i items and w. The recurrence is: C(i,w) = \begin{cases} C(i-1,w) & \text{if } w < w_i \\ \max\left( C(i-1,w),\ C(i-1, w - w_i) + v_i \right) & \text{if } w \geq w_i \end{cases} With base cases C(0,w) = 0 and C(i,0) = 0, a 2D table computes this in O(nW) time and space. Optimal substructure holds because the best selection for i items either excludes item i or includes it with the best for the reduced capacity. Space optimization for the 0-1 knapsack reduces the 2D table to a 1D array of size W+1, iterating over items and updating the array backward from w = W to w_i to prevent overwriting values needed for subsequent computations within the same item. This maintains correctness while using O(W) space instead of O(nW), suitable when W is smaller than nW.

Advanced Paradigms

Backtracking

Backtracking is an algorithmic paradigm used to solve computational problems by systematically exploring possible solutions through a depth-first search strategy, incrementally building candidate solutions and abandoning partial solutions that cannot lead to a valid complete solution. This approach is particularly effective for constraint satisfaction problems (CSPs), where variables must be assigned values satisfying a set of constraints. Unlike brute-force enumeration, which exhaustively checks all possibilities without early termination, backtracking prunes the search space by detecting failures early and retracting choices—a process known as "backtracking"—to avoid exploring invalid paths. The core mechanism of backtracking involves constructing partial solutions step by step, typically by assigning values to variables in a sequential order. At each step, the algorithm attempts to extend the current partial solution by trying feasible values for the next variable. If a value leads to a constraint violation, the algorithm immediately abandons that branch and reverts to the previous decision point to try an alternative value, effectively undoing the invalid assignment. This recursive process continues until either a complete valid solution is found or all possibilities are exhausted, at which point the algorithm reports no solution exists. The state of the search is maintained through the recursion stack, which tracks the current partial solution and the choices made so far. A classic example of backtracking is the N-Queens problem, which seeks to place N queens on an N×N chessboard such that no two queens threaten each other (i.e., no two share the same row, column, or diagonal). The algorithm proceeds row by row: for each row, it tries placing a queen in each column, checking constraints against previously placed queens. If a placement violates a constraint, it backtracks to the previous row and tries the next column there. This process efficiently explores the solution space, often finding a solution for moderate N (e.g., N up to 15) by avoiding full enumeration of all N! permutations. Another representative application is solving Sudoku puzzles, where the goal is to fill a 9×9 grid with digits 1-9 such that each row, column, and 3×3 subgrid contains all digits without repetition. Backtracking treats empty cells as variables and digits as possible values, proceeding cell by cell in row-major order. For each cell, it tries digits 1 through 9, checking row, column, and subgrid constraints; invalid trials trigger backtracking to the previous cell. Constraint propagation—pre-checking used digits in affected rows, columns, and subgrids—enhances efficiency by reducing trial-and-error. This method can solve standard Sudoku instances rapidly, as the inherent constraints prune most branches early. Pruning is central to backtracking's efficiency, as it leverages problem-specific constraints to eliminate subtrees of the search space before full exploration. Without pruning, the worst-case time complexity is O(d^n) for a CSP with n variables each of domain size d, akin to brute-force but with systematic ordering; pruning can reduce this dramatically by detecting inconsistencies at shallow depths, making it feasible for instances where n is small (e.g., 20-30 variables). For instance, in N-Queens, diagonal and column checks prune invalid column choices immediately, avoiding deeper invalid placements. Implementation typically uses a recursive function that takes the current state—such as the partial assignment (e.g., an array of placed queens or filled cells) and the next position to fill—as parameters. The function first checks if the current partial solution is complete and valid; if so, it returns success. Otherwise, it iterates over possible values for the current position, temporarily assigns a value, and recursively calls itself on the updated state. If the recursive call succeeds, it propagates success upward; if it fails for all values, it backtracks by unassigning the current value and returns failure. Pseudocode for this structure, as in CSP solvers, ensures thread-safety via copying or undoing state changes.
pseudocode
function backtrack(state, position):
    if position == goal:
        return true  // Complete solution found
    for each candidate in choices_for(position):
        if is_valid(state, position, candidate):
            state[position] = candidate
            if backtrack(state, position + 1):
                return true
            state[position] = unset  // Backtrack
    return false

Branch-and-Bound

Branch-and-bound is an algorithmic paradigm employed to solve combinatorial optimization problems by systematically partitioning the solution space into subproblems represented as a search tree, while leveraging upper and lower bounds to prune branches that cannot produce solutions superior to the current best known feasible solution. This approach was pioneered by Land and Doig in 1960 as an extension of linear programming to address discrete programming problems where variables must take integer values, enabling the automatic enumeration and fathoming of suboptimal regions without exhaustive search. By focusing on bounding mechanisms, it guarantees exact optimal solutions for NP-hard problems, distinguishing it from heuristic methods that may only approximate. The core steps of branch-and-bound involve initializing the search with a root node representing the full problem and an initial incumbent solution (often obtained from a relaxation, with its objective value serving as a preliminary bound). The algorithm then selects a live node from the tree, computes its bound by solving a relaxation of the subproblem, and branches by dividing the node into child subproblems (e.g., by fixing a variable to possible integer values). If the subproblem's bound is worse than the incumbent (for minimization, a lower bound exceeding the incumbent's value leads to pruning), the branch is fathomed; otherwise, it is added to the active node list. This process repeats—selecting nodes, bounding, branching, and updating the incumbent upon finding better feasible solutions—until no live nodes remain, at which point the incumbent is the optimal solution. Branch-and-bound implementations vary by node selection strategy to balance computational efficiency and memory usage. FIFO (first-in, first-out), akin to , processes nodes in the order they are generated, exploring the tree level by level but potentially requiring exponential memory for large problems. LIFO (last-in, first-out), or , prioritizes the most recently created nodes, enabling deeper exploration with minimal memory overhead through recursive or stack-based implementation, though it may delay finding good incumbents. LC (least cost), also known as , selects the node with the smallest (tightest) bound, aiming to minimize the number of nodes processed by focusing on promising areas first, but it demands significant memory to maintain a priority queue of all active nodes. Illustrative applications include the traveling salesman problem (TSP), where the search tree branches on the sequence of cities in the tour, and lower bounds are computed using a reduced cost matrix derived from the relaxation; rows and columns are subtracted by their minima to yield a non-negative matrix, with the reduction cost plus the assignment solution providing a bound that prunes branches exceeding the best tour found so far. For instance, in a symmetric TSP with 50 cities, the paradigm efficiently navigates the (n-1)!/2 possible tours (approximately 3 × 10^{62}) by fathoming subtrees via these bounds. In (ILP), branch-and-bound fixes fractional variables from the LP relaxation to integer values (floor or ceiling), resolving subproblem LPs to obtain lower bounds; if the LP optimum exceeds the incumbent, the branch is pruned, effectively closing the integrality gap iteratively. Relaxations are central to generating effective bounds in branch-and-bound. Linear programming (LP) relaxations drop integrality constraints to solve continuous approximations at each node, yielding lower bounds for minimization problems that are often tight due to the polyhedral structure of ILPs. For TSP, the reduced matrix approach relaxes the tour structure into a linear assignment problem, while Lagrangian relaxation dualizes constraints (e.g., degree restrictions) to maximize a lower bound via subgradient optimization, sometimes outperforming standard LP by exploiting problem-specific structure. These techniques ensure pruning efficacy, as the bound quality directly impacts the algorithm's scalability.

Randomized Algorithms

Randomized algorithms incorporate elements of randomness into their execution to achieve improved efficiency, simplicity, or performance guarantees compared to purely deterministic approaches. These algorithms leverage probabilistic methods to solve problems where randomness helps avoid worst-case scenarios or simplifies complex computations. A key advantage is their ability to provide expected runtime bounds that are superior to deterministic counterparts in many cases, such as sorting or selection problems. Randomized algorithms are broadly classified into two types: Las Vegas and Monte Carlo. Las Vegas algorithms always produce the correct output but have randomized running times, ensuring no risk of error while benefiting from expected efficiency. A prominent example is randomized quicksort, which selects pivots uniformly at random to partition the array; its expected running time is O(n \log n) for n elements, analyzed using the linearity of expectation over pairwise comparisons. In contrast, Monte Carlo algorithms have fixed running times but may err with some probability, trading correctness for speed. The Miller-Rabin primality test exemplifies this: for an odd integer n > 2, it writes n-1 = 2^r d with d odd, picks a random base a \in \{2, \dots, n-2\}, and checks modular exponentiations a^{2^y d} \mod n for y = 0 to r; if n is prime, it always declares "prime," but for composites, the error probability per trial is at most $1/4, dropping to (1/4)^k after k independent runs. Common techniques in randomized algorithms include random sampling and hashing. Random sampling selects subsets probabilistically to estimate properties or solve subproblems, often simplifying analysis through independence assumptions. , introduced by Carter and Wegman, provides a family of hash functions where, for distinct keys x \neq y, the probability of collision h(x) = h(y) is at most $1/m for m buckets, mitigating worst-case adversarial inputs in data structures like hash tables. These techniques enable applications such as efficient testing, where random walks or sampling detect paths with high probability in logarithmic space. Analysis of randomized algorithms frequently relies on probabilistic tools like the linearity of expectation, which states that for indicator random variables X_i, the expected value E\left[\sum X_i\right] = \sum E[X_i], regardless of dependence; this simplifies bounding expected costs, as in quicksort's comparison count. Derandomization techniques, such as the method of conditional probabilities, convert these algorithms to deterministic ones by sequentially fixing random choices to preserve favorable expected values, often yielding polynomial-time deterministic alternatives. Unlike deterministic methods like dynamic programming, which guarantee exact solutions through exhaustive computation, randomized approaches offer probabilistic guarantees that can be derandomized when needed.

Specialized Paradigms

addresses NP-hard problems by incorporating an auxiliary parameter k, typically a small related to the problem instance, such as the solution size. A parameterized problem consists of an input instance paired with this parameter, and the goal is to determine if the problem admits an running in time O(f(k) \cdot n^c), where n is the input size, c is a constant independent of k, and f is a . This framework distinguishes tractable cases where k is fixed or small from intractable ones, providing a refined analysis beyond classical polynomial-time complexity. The central class in parameterized complexity is fixed-parameter tractable (FPT), comprising problems solvable in the desired time bound. A key technique for establishing FPT membership is kernelization, a preprocessing that reduces the instance to a whose size is bounded by a of k alone, after which standard algorithms suffice. For example, the k- problem—deciding if a has a vertex set of size at most k covering all edges—is FPT and admits a with at most $2k vertices via rounding and matching-based reductions. In contrast, problems in the W class, presumed not FPT under standard conjectures, include the k- problem of finding a of size k in a , which is W-hard. Core techniques for designing FPT algorithms include bounded search trees and color-coding. Bounded search trees involve recursively branching on choices (e.g., including or excluding a in a cover), ensuring the tree's and depth yield a total size bounded by $2^k or similar, leading to FPT time; for k-, a simple branching rule on high-degree vertices achieves O(2^k k n) time. Color-coding, a randomized method derandomizable via hashing, assigns colors to vertices to detect colorful substructures, useful for subgraph isomorphism problems like finding a k- in O(2^{O(k)} n) time, with extensions to other induced subgraphs. These approaches highlight how parameterization enables exact solutions for hard problems when k is modest.

Approximation Algorithms

Approximation algorithms provide polynomial-time methods for solving NP-hard optimization problems by producing solutions that are guaranteed to be within a specified of the optimal . For minimization problems, a ρ-approximation algorithm outputs a feasible solution whose is at most ρ times the optimal cost OPT, where ρ ≥ 1; for maximization problems, the value is at least OPT/ρ. This paradigm is essential for problems where exact solutions are computationally intractable, offering provable performance guarantees rather than heuristics without bounds. A classic example is the minimum vertex cover problem, where a 2-approximation is achieved by computing a maximal matching and selecting both endpoints of each matched edge, yielding a cover of size at most twice the optimum since the optimal cover must include at least one endpoint per matching edge. Techniques often extend greedy strategies, such as iteratively selecting elements to cover constraints while bounding the growth relative to OPT. Linear programming (LP) rounding is another core method: solve an LP relaxation of the integer program, then round the fractional solution to an integer one. For vertex cover, the LP relaxation minimizes ∑ x_v subject to x_u + x_v ≥ 1 for each edge (u,v) and 0 ≤ x_v ≤ 1; rounding vertices with x_v ≥ 1/2 yields a 2-approximation, as the rounded cost is at most twice the LP value, which lower-bounds OPT. Integrality gap analysis measures the worst-case ratio of OPT to the LP optimum, revealing inherent limits of the relaxation; for vertex cover, the gap is exactly 2, as shown by complete graphs where OPT = n-1 but LP ≤ n/2. Hardness results from the establish that certain problems resist better s unless P=. The states that every language in has a probabilistic verifier using O(log n) random bits and querying a constant number of proof bits, with perfect completeness and soundness at least 1/2. It implies no polynomial-time scheme (PTAS)—an achieving (1+ε)ρ for any ε>0—exists for MAX-SNP-hard problems like or maximum 3-SAT, as approximating them within a constant factor better than 7/8 for MAX-3SAT is NP-hard. A prominent application is the metric traveling salesman problem (TSP), where computes a , adds a minimum-weight on odd-degree vertices, and shortcuts the resulting Eulerian tour to form a Hamiltonian cycle, achieving a 1.5-approximation since the tree and matching costs are each at most OPT, and shortcutting preserves the bound in spaces. Randomized enhancements, such as derandomizing matchings, can integrate with these deterministic methods for robustness.

Online Algorithms

Online algorithms operate in an environment where input arrives incrementally over time, and the algorithm must make irrevocable decisions based solely on the data seen so far, without knowledge of future inputs. This paradigm contrasts with offline algorithms, which have access to the entire input in advance, and is particularly relevant for systems such as caching, scheduling, and where delays are not feasible. The performance of online algorithms is typically evaluated using competitive analysis, a framework introduced by Sleator and Tarjan, which measures how closely the algorithm's cost approximates that of an optimal offline algorithm on the same input sequence. Specifically, an algorithm is r-competitive if its cost ALG(I) satisfies ALG(I) ≤ r · OPT(I) for any input I, where OPT(I) denotes the cost of the optimal offline solution, often with an additive constant term for practicality. This ratio captures the inherent trade-off due to the lack of foresight, and lower bounds are established via adversary arguments that construct worst-case input sequences to force suboptimal choices. A classic example is the paging problem in management, where a limited of k pages must evict pages to accommodate new requests. The least recently used (LRU) eviction policy achieves a competitive of k, meaning its fault rate is at most k times that of the optimal offline paging algorithm, as proven through analysis that bounds the amortized cost per fault. Another foundational deterministic appears in the ski-rental problem, a rent-or-buy decision model where the algorithm rents at per day until the total rental cost reaches the buy cost B, then buys; this yields a 2-competitive , optimal for deterministic algorithms, as any can be forced to pay nearly twice the optimum by an adversary stopping just before or after the threshold. To mitigate the limitations of deterministic approaches, randomized online algorithms introduce randomness to hedge against adversarial inputs, analyzed via expected costs and , which equates the performance of the best randomized algorithm against an oblivious adversary to the minimax value over distributions on inputs and deterministic algorithms. For the ski-rental problem, randomization achieves an expected competitive ratio of e/(e-1) ≈ 1.582, by randomly selecting a rental threshold from a , outperforming deterministic methods while provides a matching lower bound by considering input distributions that minimize the maximum deterministic ratio. Practical examples include list scheduling for online load balancing on m identical machines, where jobs arrive sequentially and are greedily assigned to the machine with the current minimum load; this algorithm is (2 - 1/m)-competitive for minimizing makespan, as the maximum load is at most twice the optimum minus a small term, with the bound tight via adversarial job sequences of equal size. In online bin packing, the first-fit heuristic places each item into the lowest-indexed bin with sufficient space, achieving a competitive ratio of 1.7 for minimizing the number of bins used, derived from harmonic analysis of bin utilization and worst-case item size distributions constructed by an adversary to waste space. These examples highlight the paradigm's efficacy in dynamic settings, though inherent inefficiencies persist: without future knowledge, no online algorithm can match offline optima exactly, and adversary arguments reveal lower bounds like Ω(log n / log log n) for certain scheduling problems, underscoring the value of competitive ratios as benchmarks against offline approximations.

Applications in Domains

Computational Geometry

Computational geometry focuses on the design and for solving problems involving geometric objects, such as points, lines, and polygons in . Unlike general algorithmic paradigms that operate on abstract data structures, geometric paradigms exploit spatial relationships and to achieve efficiency, often achieving near-linear time complexities for low-dimensional problems through techniques like sweeping and duality transformations. These methods are particularly suited to handling the inherent in geometric intersections and arrangements, enabling practical solutions for real-world spatial computations. A cornerstone paradigm in computational geometry is the sweep line algorithm, which simulates a line moving across the plane to process events dynamically. For detecting all intersections among a set of n line segments, the Bentley-Ottmann algorithm employs a vertical sweep line that advances from left to right, maintaining an ordered status structure of active segments intersected by the line and a priority queue of event points, including segment endpoints and intersection candidates. This approach reports all k intersections in O((n + k) \log n) time, optimal in the algebraic decision tree model, by handling only O(n + k) events while using balanced trees for order maintenance. The paradigm extends to Voronoi diagrams, where Fortune's sweepline algorithm constructs the diagram for n sites in O(n \log n) time by transforming the problem into a parabolic wavefront propagation, tracking beach line boundaries with a kinetic data structure. Incremental construction paradigms build geometric structures by successively adding elements, repairing local inconsistencies to maintain global properties. In , which maximizes the minimum angle among all triangulations of a point set, Lawson's flipping algorithm starts with an arbitrary triangulation and iteratively replaces non-locally Delaunay edges—those violated by the empty circle property—with the alternative diagonal until convergence, yielding a triangulation in O(n^2) worst-case time but often faster in practice. An alternative lifting transformation maps points to a in 3D space, converting the 2D Delaunay triangulation to the lower of lifted points, computable via 3D in O(n \log n) expected time with randomized incremental methods. Randomized incremental construction further optimizes geometric paradigms by randomizing insertion order to bound expected conflicts. For trapezoidal maps, which decompose a planar subdivision into trapezoids for efficient point location, Seidel's algorithm inserts n non-intersecting segments randomly, resolving the visibility zone of each new segment in expected O(\log n) time per insertion via a search structure on prior trapezoids, achieving overall expected O(n \log n) construction time and supporting O(\log n) queries. This backward analysis technique, where conflict degrees decrease probabilistically, distinguishes geometric randomized methods from general ones by leveraging spatial locality. These paradigms find extensive applications in geographic information systems (GIS) for overlay analysis and spatial querying, where sweep line techniques efficiently compute map unions and intersections for large datasets. In , incremental and randomized constructions enable path planning, such as constructing graphs or trapezoidal decompositions for motion in polygonal environments, ensuring collision-free trajectories in O(n \log n) time. A key difference from general paradigms lies in tools like geometric ity, which interchanges points and lines (e.g., dual of a point (a,b) is line y = 2ax - b), transforming incidence problems into order queries to simplify arrangements and half-plane intersections. While divide-and-conquer provides a foundational base for many geometric algorithms, such as merge-based of slopes, specialized paradigms like sweeping offer dynamic event handling absent in static divisions.

Graph Algorithms

Graph algorithms represent a class of algorithmic paradigms tailored to problems on graph structures, where vertices and edges model relationships such as , distances, or capacities. These paradigms often leverage the graph's to achieve efficiency, transforming abstract problems into traversals, optimizations, or flows. Fundamental approaches include exhaustive search via traversals for basic and more sophisticated techniques like selection, dynamic programming, matching contractions, and network flows for problems involving paths, pairings, and bottlenecks. Breadth-first search (BFS) and depth-first search (DFS) exemplify brute-force traversal paradigms for exploring graph connectivity and finding shortest paths in unweighted graphs. BFS systematically explores vertices level by level from a starting source, using a queue to ensure all nearer vertices are visited before farther ones, achieving O(V + E) time complexity where V is the number of vertices and E the number of edges. This paradigm guarantees the shortest path in terms of edge count by maintaining FIFO order. DFS, conversely, delves deeply along each branch before backtracking, employing a stack or recursion, and also runs in O(V + E) time but prioritizes depth over breadth, useful for detecting cycles or topological orders. Both are foundational for graph exploration, with BFS originating in maze-solving contexts and DFS formalized in linear-time graph processing. The paradigm is prominently applied in for single-source shortest paths in non-negative weighted graphs. Starting from a source , it iteratively selects the unvisited with the smallest tentative distance, updating distances to its neighbors if a shorter path is found, ensuring optimality through the choice property. The original implementation uses a simple array for O(V^2) time, but with a , it achieves O((V + E) log V) efficiency by efficiently extracting minima and decreasing keys. This approach relies on the absence of negative weights to avoid revisiting vertices, making it a cornerstone for and problems. Dynamic programming paradigms exploit the acyclic structure of directed acyclic graphs (DAGs) to solve path optimization problems, such as finding the longest path, which is NP-hard in general graphs but tractable here. By first computing a topological sort in O(V + E) time, which orders vertices such that all predecessors precede successors, dynamic programming can then process vertices in this order. For each vertex u, the longest path length dp is computed as 1 plus the maximum dp over outgoing neighbors v, with base cases for sinks at 0; this ensures subproblems are solved bottom-up without cycles, yielding the global longest path in O(V + E) time. This paradigm highlights how DAGs enable optimal substructure for path maximization, contrasting with shortest paths where relaxation suffices. For matching problems, the by Edmonds introduces a paradigm of iterative augmentation and contraction to find maximum cardinality matchings in general (non-bipartite) graphs. It begins with an empty matching and repeatedly seeks augmenting paths relative to the current matching using DFS-like searches; upon discovering odd-length cycles (blossoms), it contracts them into supervertices to simplify the graph, then expands back after augmentation. This process runs in O(V^4) time in the original formulation, though optimized implementations achieve O(V^3), and guarantees a maximum matching by Berge's lemma, which states no larger matching exists without an augmenting path. The paradigm extends bipartite matching techniques like Hopcroft-Karp by handling non-bipartite complexities through structural reductions. Flow-based paradigms address connectivity and capacity constraints via the , which equates the maximum from to with the minimum cut separating them in a capacitated . The Ford-Fulkerson operationalizes this by iteratively finding augmenting paths in the residual —edges with remaining —and augmenting along them until no path exists, with each path increasing by its bottleneck . Using DFS or BFS for path discovery, it terminates in O(E f) time where f is the maximum value for integer capacities, but variants like Edmonds-Karp with BFS ensure O(V E^2) polynomial time by bounding augmentations. This paradigm models diverse applications, from bipartite matching (via unit capacities) to network reliability, with the theorem providing duality between and cuts.

Machine Learning

Machine learning relies on several algorithmic paradigms to optimize models, make decisions under , and combine predictions for improved accuracy. These paradigms draw from broader algorithmic traditions, adapting techniques like iterative optimization, dynamic programming, and ensemble methods to handle high-dimensional data and probabilistic outcomes. Central to many machine learning approaches is the paradigm of optimization, particularly , which iteratively refines model parameters to minimize a . Gradient descent operates as an iterative , updating parameters in the direction opposite to the of the objective function. The core update rule is given by \theta \leftarrow \theta - \alpha \nabla J(\theta), where \theta represents the parameters, \alpha is the , and \nabla J(\theta) is the of the loss J. This method, extended to variants for large-scale , enables efficient of models like and neural networks by approximating the full with mini-batches, reducing computational cost while converging to local minima. In practice, has become foundational for scalable , powering systems that process billions of examples. Another key paradigm in , especially , is dynamic programming, applied to solve Markov decision processes (MDPs). Value iteration, a dynamic programming technique, computes the optimal value function by iteratively applying the Bellman optimality equation: V(s) = \max_a \left[ R(s,a) + \gamma \sum_{s'} P(s'|s,a) V(s') \right], where V(s) is the value of state s, R(s,a) is the reward for action a in s, \gamma is the discount factor, P(s'|s,a) is the transition probability to s', and the maximization is over actions a. This equation, derived from the principle of optimality, allows systematic computation of expected future rewards, enabling policy derivation for sequential decision-making in environments like robotics and game playing. Introduced in the foundational work on dynamic programming, value iteration guarantees convergence to the optimal function under finite state-action spaces. Ensemble methods represent a of combining multiple learners to enhance performance, with and boosting serving as randomized and adaptive averaging techniques to reduce variance. , or , generates diverse predictors by training on bootstrap samples of the data and averaging their outputs, which stabilizes high-variance models like decision trees by mitigating through . This approach demonstrably improves accuracy on unstable classifiers, as evidenced in empirical studies on datasets like those from the UCI repository. Boosting, in contrast, builds ensembles sequentially by weighting misclassified examples more heavily in subsequent iterations, adaptively focusing on hard cases to minimize exponential loss; , a seminal boosting , achieves strong theoretical guarantees on error rates for weak learners. Together, these paradigms reduce variance in bagging via and bias in boosting via adaptation, yielding robust predictors in applications from to . In neural networks, an approximation paradigm underlies , which computes gradients via reverse-mode to efficiently train deep architectures. This method propagates errors backward through the network, adjusting weights layer by layer to approximate the true of the loss with respect to all parameters, enabling the scaling of to millions of weights. By leveraging the chain rule in a reverse pass, backpropagation avoids the exponential cost of forward-mode , making it indispensable for learning hierarchical representations in tasks like image recognition.

References

  1. [1]
    3.4 Algorithmic Paradigms - Introduction to Computer Science
    Nov 13, 2024 · Here, we will introduce the algorithmic paradigm, which is the common concepts and ideas behind algorithm design patterns.
  2. [2]
    Introduction to Algorithms | Electrical Engineering and Computer ...
    This course provides an introduction to mathematical modeling of computational problems. It covers the common algorithms, algorithmic paradigms, and data ...Lecture Notes · Assignments · Syllabus · Lecture 1: Algorithmic Thinking...
  3. [3]
    Algorithmic Patterns
    Definition. An algorithmic pattern, or algorithmic paradigm, is a method, strategy, or technique of solving a problem.
  4. [4]
    Differentiate between event driven paradigm and algorithmic ...
    Dec 1, 2021 · An algorithmic paradigm is a generic model or framework that underlies the design of a class of algorithms. It is an abstraction higher than the ...
  5. [5]
    [PDF] The seven-eleven problem - CS@Cornell
    Sep 15, 1983 · Saddleback Search is a useful algorithmic paradigm because of its generality-it is applicable to any matrix with ordered rows and columns. However we know ...
  6. [6]
    Euclidean algorithm | Algorithm, Division & GCD - Britannica
    Sep 16, 2025 · Euclidean algorithm, procedure for finding the greatest common divisor (GCD) of two numbers, described by the Greek mathematician Euclid in his Elements (c. ...
  7. [7]
    1.8: The Euclidean Algorithm - Mathematics LibreTexts
    Jan 22, 2022 · The Euclidean Algorithm is named after Euclid of Alexandria, who lived about 300 BCE. The algorithm 1 described in this chapter was recorded ...Missing: history origin
  8. [8]
    Analytical Engine | Description & Facts - Britannica
    Oct 9, 2025 · Analytical Engine, generally considered the first computer, designed and partly built by the English inventor Charles Babbage in the 19th century.Missing: decomposition | Show results with:decomposition
  9. [9]
    The Engines | Babbage Engine - Computer History Museum
    Babbage's calculating engines are decimal digital machines. They are decimal in that they use the familiar ten numbers '0' to '9' and they are digital in the ...Missing: problem decomposition
  10. [10]
    Classics in the History of Psychology -- Lovelace (1843)
    The Analytical Engine is an embodying of the science of operations, constructed with peculiar reference to abstract number as the subject of those operations.
  11. [11]
    Ada Lovelace: The World's First Computer Programmer Who ...
    Mar 22, 2023 · The idea that computing machines can go beyond just calculating numbers and be used to perform any abstract operation is the defining ...
  12. [12]
    The Algebra of Logic Tradition - Stanford Encyclopedia of Philosophy
    Mar 2, 2009 · The algebra of logic, as an explicit algebraic system showing the underlying mathematical structure of logic, was introduced by George Boole (1815–1864)Missing: groundwork | Show results with:groundwork
  13. [13]
    George Boole: A 200-Year View - Stephen Wolfram Writings
    Nov 2, 2015 · George Boole's melding of logic and mathematics into Boolean algebra set the stage for many developments including universal computation.
  14. [14]
    [PDF] ON COMPUTABLE NUMBERS, WITH AN APPLICATION TO THE ...
    The "computable" numbers may be described briefly as the real numbers whose expressions as a decimal are calculable by finite means.
  15. [15]
    The Theory of Dynamic Programming | RAND
    This paper is the text of an address by Richard Bellman before the annual summer meeting of the American Mathematical Society in Laramie, Wyoming, on September ...Missing: 1953 original
  16. [16]
    The Art of Computer Programming (TAOCP) - CS Stanford
    Volume 3. Sorting and Searching , Second Edition (Reading, Massachusetts: Addison-Wesley, 1998), xiv+780pp.+foldout.
  17. [17]
    [PDF] Cook 1971 - Department of Computer Science, University of Toronto
    A method of measuring the complexity of proof procedures for the predicate calculus is introduced and discussed. Throughout this paper, a set of strings means a ...
  18. [18]
    Do we teach the right algorithm design techniques?
    The paper points out these shortcomings, reevaluates some of the techniques, and proposes a new, hierarchical classification scheme by grouping techniques ...
  19. [19]
    Turing machine - Stanford Encyclopedia of Philosophy
    Sep 24, 2018 · A (real) number is Turing computable if there exists a Turing machine which computes an arbitrarily precise approximation to that number. All of ...
  20. [20]
    Parallelism in random access machines - ACM Digital Library
    A model of computation based on random access machines operating in parallel and sharing a common memory is presented.
  21. [21]
    Fixed-Parameter Tractability and Completeness I: Basic Results
    We establish the main results of a completeness program which addresses the apparent fixed-parameter intractability of many parameterized problems.Missing: paradigms | Show results with:paradigms
  22. [22]
    [PDF] Algorithms and Complexity What is the Brute Force Approach?
    What is the Brute Force Approach? • A straightforward approach, usually based directly on the problem's statement and definitions of the concepts involved. • ...
  23. [23]
    [PDF] An Evaluation of the Traveling Salesman Problem - ScholarWorks
    The brute force algorithm has exactly n! permutations that need to be checked. Thus the time complexity is O(n!) . Although the brute force algorithm is the ...
  24. [24]
    [PDF] TSP - Traveling Salesman Problem - Columbia CS
    That's this algorithm has exponential time complexity. ○ Bruteforce, calculate path distance in parallel. In the brute force approach, we first enumerate all ...
  25. [25]
    Brute Force Algorithms Explained - freeCodeCamp
    Jan 6, 2020 · Brute force algorithms are straightforward methods that rely on sheer computing power and trying every possibility, not advanced techniques.Missing: paradigm | Show results with:paradigm
  26. [26]
    [PDF] The Traveling Salesman Problem at Taylor University
    A brute force examination of a larger TSP, for example with 𝑛 = 70, would require an impossible amount of computational power. Thus, while there clearly is an ...
  27. [27]
    [PDF] STRINGS AND PATTERN MATCHING - CS@Purdue
    Strings and Pattern Matching. Brute Force Pseudo-Code. • Here's the pseudo-code do if (text letter == pattern letter) compare next letter of pattern to next.
  28. [28]
    [PDF] Lecture 4 Divide and Conquer Maximum/minimum Median finding
    These three basic steps – divide, conquer, and combine – lie behind most divide and conquer algorithms.
  29. [29]
    [PDF] Divide and Conquer 1 Paradigm 2 Sorting - cs.wisc.edu
    An instance of the given problem is divided into easier instances of the same problem, which are solved recursively and then combined to create a solution ...
  30. [30]
    [PDF] Lecture 4: Divide and Conquer
    Divide problem into subproblems. – Conquer problems by solving recursively. – Combine solutions to solve the original problem.
  31. [31]
    [PDF] Lecture 2: Merge Sort and Master Theorem
    Note : Usually, the time to divide the problem is negligible (python list slicing, etc.) ... time complexity of mergesort is O(n1 log n) = O(nlog n). We could have ...
  32. [32]
    Divide and Conquer :: CC 310 Textbook
    Jun 21, 2024 · One great example of a divide and conquer algorithm is the binary search algorithm. If we have a list of data that has already been sorted, as ...
  33. [33]
    [PDF] 1 Solving recurrences
    The master theorem is a formula for solving recurrences of the form T(n) = aT(n/b)+f(n), where a ≥ 1 and b > 1 and f(n) is asymptotically positive. ( ...
  34. [34]
    [PDF] Divide–and–Conquer Recurrences — The Master Theorem
    We assume a divide and conquer algorithm in which a problem with input size n is always divided into a subproblems, each with input size n/b. Here a and b are ...
  35. [35]
    [PDF] CS483 Analysis of Algorithms Lecture 03 – Divide-n-Conquer ∗
    Divide-n-conquer strategy. – Advantages of. ⊲. Make problems easier. ⊲. Easy parallelization. – Disadvantages of Divide-n-conquer strategy. ⊲. Recursion can ...
  36. [36]
    [PDF] A Method for the Construction of Minimum-Redundancy Codes*
    Minimum-Redundancy Codes*. DAVID A. HUFFMAN+, ASSOCIATE, IRE. September. Page 2. 1952. Huffman: A Method for the Construction of Minimum-Redundancy Codes. 1099 ...
  37. [37]
    [PDF] Dynamic programming - Algorithms - cs.Princeton
    Dec 22, 2020 · Algorithm design paradigm. ・Break up a problem into a series of overlapping subproblems. ・Build up solutions to larger and larger subproblems.
  38. [38]
    [PDF] THE THEORY OF DYNAMIC PROGRAMMING - Richard Bellman
    This paper is the text of an invited address before the annual summer meeting of the American Mathematical Society at. Laramie, Wyoming, September 2, 1954 ...
  39. [39]
    [PDF] Lecture 18: Dynamic Programming I: Memoization, Fibonacci, Crazy ...
    It's the Fibonacci sequence, described by the recursive formula: F0 := 0; F1 := 1;. Fn = Fn−1 + Fn−2, for all n ≥ 2.
  40. [40]
    [PDF] Lecture 13
    Step 1: Identify optimal substructure. • Step 2: Find a recursive formulation for the length of the longest common subsequence. • Step 3: Use dynamic ...<|separator|>
  41. [41]
    [PDF] Dynamic programming 0-1 Knapsack problem CSCE 310J Data ...
    Given some items, pack the knapsack to get the maximum total value. Each item has some weight and some value. Total weight that we can.
  42. [42]
    [PDF] 5 CONSTRAINT SATISFACTION PROBLEMS
    A simple backtracking algorithm for constraint satisfaction problems. The algorithm is modeled on the recursive depth-first search of Chapter 3. The ...Missing: lecture | Show results with:lecture
  43. [43]
    [PDF] Backtracking
    In the n-queens problem, the goal is a sequence of queen positions, one in each row, such that no two queens attack each other. For each row, the algorithm ...
  44. [44]
    The N-Queens Problem & Backtracking
    In a backtracking algorithm, one incrementally builds candidates for the solution (or solutions in some cases) for a particular problem by making a sequence of ...
  45. [45]
    [PDF] Admin Backtracking pseudocode Sudoku code
    Sudoku solver. Arrange 1 to 9 with no repeats in row, col, or block. • Solve by recursive backtracking. • Not much logic, just brute-force. Cast as decision ...
  46. [46]
    CS140 Lecture notes -- Recursion - Sudoku Solver - UTK-EECS
    There is no explicit backtracking really -- the important part is that if the recursive solver fails, it restores the state of the grid to the state when it was ...
  47. [47]
    [PDF] An Automatic Method of Solving Discrete Programming Problems ...
    Apr 4, 2007 · This paper presents a simple numerical algorithm for the solution of programming problems in which some or all of the variables can take only ...
  48. [48]
    [PDF] Branch and Bound Algorithms - Principles and Examples.
    Mar 12, 1999 · Branch and Bound (B&B) is by far the most widely used tool for solv- ing large scale NP-hard combinatorial optimization problems. B&B is,.Missing: "Land | Show results with:"Land
  49. [49]
    [PDF] Branch and Bound Methods for the Traveling Salesman Problem.
    6,3 = 3. The lower bound becomes 17 + 2 + 3 - 22. The new reduced cost matrix is shown in Table 3 and the corresponding admissible graph G in Fig. 4. Note ...<|separator|>
  50. [50]
    [PDF] Integer Programming: The Branch and Bound Method
    A linear programming model solu- tion with no integer restrictions is called a relaxed solution. The branch and bound method uses a tree diagram of nodes and ...
  51. [51]
    [PDF] Integer Programming: Lagrangian Relaxation - John Hooker
    Lagrangian relaxation is often used in this context, because it may provide better bounds than the standard linear programming (LP) relaxation. Lagrangian ...<|separator|>
  52. [52]
    [PDF] Randomized Algorithms - CMU School of Computer Science
    21.2 Las Vegas versus Monte Carlo​​ There are two types of randomized algorithms, which are actually quite different. A Las Vegas algorithm will always produce ...
  53. [53]
    [PDF] 23 Primality Testing - CMU School of Computer Science
    Unlike the Fermat Primality Test, the Miller–Rabin Primality Test works on every number n. Like the Fermat Primality Test, the Miller–Rabin test always returns.
  54. [54]
    Universal classes of hash functions - ScienceDirect.com
    G. MARKOWSKY, J. L. CARTER, AND M. N. WEGMAN, Analysis of a universal class of hash functions, in “Proceedings of the Seventh Mathematical Foundations of ...
  55. [55]
    [PDF] 44 RANDOMIZATION AND DERANDOMIZATION - CSUN
    METHOD OF CONDITIONAL PROBABILITIES. The method of conditional probabilities (also called the Raghavan-Spencer method) [Spe87, Rag88] implements a binary ...
  56. [56]
    Parameterized Complexity | SpringerLink
    This book presents an approach to complexity theory which offers a means of analyzing algorithms in terms of their tractability Downey considers problems in ...
  57. [57]
    Vertex Cover Kernelization Revisited | Theory of Computing Systems
    Mar 8, 2012 · Using the terminology of parameterized complexity we say that k-Vertex Cover has a kernel with 2k vertices. There is complexity-theoretic ...
  58. [58]
    Color-coding | Journal of the ACM
    ~ALON, N., YUSTER, R., AND ZWICK, U. Finding and counting given length cycles. SL4M J. Disc. ~Math. to appear. Google Scholar. [5]. ~BODLAENDER, H.L. 1993. On ...<|control11|><|separator|>
  59. [59]
    [PDF] Approximation Algorithms
    An overview of these results is presented in Chapter 29, assuming the main technical theo- rem, the PCP Theorem.
  60. [60]
    None
    ### Summary of LP Rounding for Vertex Cover and Integrality Gap Analysis
  61. [61]
    [PDF] Proof Verification and the Hardness of Approximation Problems
    This paper shows that every language in NP has a probabilistic verifier and that no MAX SNP-hard problem has a polynomial time approximation scheme unless NP=P.<|separator|>
  62. [62]
    [PDF] 1 Online Algorithms 2 Ski Rental - cs.Princeton
    The ski rental problem involves deciding whether to buy skis (one-time cost B) or rent (R per trip) without knowing how often you'll ski.
  63. [63]
  64. [64]
    Depth-First Search and Linear Graph Algorithms | SIAM Journal on ...
    References. 1. B. L. Fox, D. M. Landy, An algorithm for identifying the ergodic subchains and transient states of a stochastic matrix ...Missing: original | Show results with:original
  65. [65]
    [PDF] A Decentralized Algorithm for Finding the Shortest Paths in ... - DTIC
    Moore, E. F., “The shortest path through a maze,” Proc. m t. Svmp. on the Theory of Switching, Part II, April 2-5, 1957, t~e Annals of the Computation ...
  66. [66]
    [PDF] dijkstra-routing-1959.pdf
    Paper, P-923, 1956. [4] BERGE, C.: Théorie des graphes et ses applications, pp. 68-69. Paris: Dunod 1958. Mathematisch Centrum. 2e Boerhaavestraat 49.Missing: citation | Show results with:citation
  67. [67]
    [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 ...
  68. [68]
    [PDF] maximal flow through a network - lr ford, jr. and dr fulkerson
    399 Page 2 400 L. R. FORD, JR. AND D. R. FULKERSON a saturated arc. The value of a flow is the sum of the numbers of all the chain flows which compose it.Missing: citation | Show results with:citation