Fact-checked by Grok 2 weeks ago

Time complexity

Time complexity is a fundamental concept in that quantifies the amount of computational time required by an to process an input of a given size, typically expressed as a function of the input length n. It focuses on the worst-case scenario, where the running time T(n) represents the maximum number of steps a deterministic takes to halt on any input of length n. The measurement of time complexity involves counting basic operations, such as comparisons or assignments, performed during execution, rather than actual clock time, to ensure hardware-independent analysis. Different cases are considered: ** captures the maximum time over all inputs of size n, average-case evaluates the expected time assuming a over inputs, and best-case denotes the minimum time for favorable inputs. For instance, algorithms like exhibit O(n) worst-case time, while more efficient structures like binary search achieve O(\log n). Asymptotic notation, particularly (O), provides an upper bound on the growth rate of the running time, ignoring constant factors and lower-order terms to highlight scalability for large n. Complementary notations include Big Omega (\Omega) for lower bounds and Big Theta (\Theta) for tight bounds, enabling precise comparisons of algorithm efficiency. Common complexities range from constant O(1) for fixed-time operations to exponential O(2^n) for intractable problems. Time complexity analysis is essential for algorithm design and selection, influencing the classification of problems into complexity classes such as , the set of decision problems solvable in polynomial time (i.e., O(n^k) for some constant k). It underpins by revealing resource trade-offs and proving hierarchies, like the time hierarchy theorem stating that more time allows solving harder problems. In practice, it guides optimizations in to ensure for real-world data sizes.

Fundamentals

Definition and Measurement

Time complexity refers to the amount of computational resources, specifically the number of basic operations, required by an to process an input of size n. It is formally defined as a function T(n), which quantifies the number of steps or operations needed to complete the . This measure captures the algorithm's efficiency in terms of time, independent of hardware specifics, by focusing on the growth rate relative to input size. Analysis typically considers the worst case (maximum resources needed), average case (expected resources over random inputs), or best case (minimum resources). To measure time complexity, one counts the operations executed by , such as assignments, comparisons, operations, or control transfers, each assumed to take constant time. For instance, in analysis, each line or statement involving these primitives contributes to the total count, with the overall T(n) derived from the dominant terms in loops or recursions. Constant factors and lower-order terms are disregarded to emphasize for large n, as these details vary by but do not affect the asymptotic behavior. The concept originated in the 1960s through foundational works on resource-bounded . Alan Cobham's 1965 paper introduced the idea of intrinsic computational difficulty, proposing polynomial-time as a machine-independent measure of feasible . Independently, Juris Hartmanis and Richard E. Stearns formalized time complexity in their 1965 paper, defining it via steps to classify computable functions by resource demands. For example, accessing an element in an by requires a fixed number of operations regardless of array size, while an array of n elements involves a variable number of comparisons and swaps that grows with n. These illustrate how T(n) can remain bounded or scale polynomially.

Asymptotic Notation

Asymptotic notation provides a mathematical framework for describing the growth rates of functions, particularly the running time T(n) of algorithms as the input size n tends to . By focusing on dominant terms and ignoring constants and lower-order factors, these notations enable comparisons of algorithm efficiency without precise measurements. The most common notations—, Big Θ, and Big Ω—offer upper, tight, and lower bounds, respectively, while their "little" counterparts provide stricter inequalities. These tools originated in but were popularized in for analyzing . Big O notation, denoted as T(n) = O(f(n)), specifies an asymptotic upper bound on the growth rate of T(n). Formally, T(n) = O(f(n)) if there exist constants c > 0 and n_0 > 0 such that $0 \leq T(n) \leq c \cdot f(n) for all n \geq n_0. This definition captures the worst-case scenario, ensuring that T(n) grows no faster than a constant multiple of f(n) for sufficiently large n. An equivalent formulation uses the limit superior: T(n) = O(f(n)) if \limsup_{n \to \infty} \frac{T(n)}{f(n)} < \infty. For instance, constant-time operations like array access are analyzed as O(1). The Theta notation, T(n) = \Theta(f(n)), provides a tight bound by combining upper and lower estimates: T(n) = O(f(n)) and T(n) = \Omega(f(n)). Formally, there exist constants c_1 > 0, c_2 > 0, and n_0 > 0 such that c_1 \cdot f(n) \leq T(n) \leq c_2 \cdot f(n) for all n \geq n_0. This indicates that T(n) grows at the same rate as f(n), up to constant factors. The form is \lim_{n \to \infty} \frac{T(n)}{f(n)} = L where $0 < L < \infty. Theta is particularly useful for average- or best-case analyses where bounds coincide. Big Ω notation, T(n) = \Omega(f(n)), describes a lower bound on growth: T(n) = \Omega(f(n)) if there exist constants c > 0 and n_0 > 0 such that T(n) \geq c \cdot f(n) for all n \geq n_0. Equivalently, f(n) = O(T(n)), or \liminf_{n \to \infty} \frac{T(n)}{f(n)} > 0. This ensures T(n) grows at least as fast as f(n), highlighting minimum resource requirements. For linear scans, such as traversing an of size n, the time is \Omega(n). Little-o and little-ω notations introduce stricter bounds without constant factors. Little-o, T(n) = o(f(n)), means \lim_{n \to \infty} \frac{T(n)}{f(n)} = 0, indicating T(n) grows strictly slower than f(n). For example, n = o(n \log n) since \lim_{n \to \infty} \frac{n}{n \log n} = 0, showing linear growth lags behind . Conversely, little-ω, T(n) = \omega(f(n)), satisfies \lim_{n \to \infty} \frac{f(n)}{T(n)} = 0, or f(n) = o(T(n)), for strict lower bounds. These are less common in basic analyses but refine comparisons of subdominant terms. To illustrate derivation, consider an with two nested s, each from 1 to n, executing a constant-time inside:
for i in 1 to n:
    for j in 1 to n:
        # constant time [operation](/page/Operation)
The inner runs n times per outer , yielding \sum_{i=1}^n n = n^2 total s. Since n^2 \leq 1 \cdot n^2 for n ≥ 1, this proves T(n) = O(n^2). More generally, summations like \sum_{i=1}^n i = \frac{n(n+1)}{2} = O(n^2) bound growth.

Sublinear Time Complexities

Constant Time

Constant time complexity, denoted as O(1), refers to algorithms whose execution time remains fixed and independent of the input size n. In such cases, the number of basic operations performed is bounded by a constant, regardless of how large the input grows. This is formally expressed by the time function T(n) = c, where c is a positive constant representing the fixed number of steps. Representative examples of O(1) operations include direct array indexing, where accessing an element at a given requires a constant number of steps to compute the . Similarly, hash table lookups achieve O(1) average-case performance under ideal hashing conditions, as the key is mapped directly to a location with a fixed number of comparisons. Simple arithmetic operations, such as adding two numbers, also fall into this category, executing in a bounded time irrespective of size in implementations. In practice, constant time algorithms offer excellent for handling large datasets, as their does not degrade with increasing input volume, making them preferable in systems where predictable execution is essential. However, hardware constraints, such as CPU behavior, can introduce variations in the actual runtime even for O(1) operations; for instance, cache misses may cause timing discrepancies that affect security-sensitive applications requiring strict constant-time guarantees. A common misconception is that O(1) implies uniformly fast execution across all implementations, overlooking how factors like effects or optimizations can alter the hidden constant c.

Logarithmic Time

Logarithmic time complexity describes algorithms whose worst-case running time grows proportionally to the logarithm of the input size n, formally expressed as O(\log n). This class of complexity arises in scenarios where the problem size is repeatedly divided by a constant factor, typically 2 in binary-based operations, leading to a number of steps that scales logarithmically rather than linearly with n. In computer science, the logarithm is conventionally base-2 (\log_2 n), reflecting the binary nature of data representation and operations like bit shifts or halvings. A prominent example of logarithmic time is binary search, which efficiently locates an in a sorted by repeatedly dividing the search interval in half. Starting with an of size n, each iteration compares the target to the middle and discards half the remaining elements, reducing the problem size exponentially until the target is found or confirmed absent. This process requires at most \lceil \log_2 (n+1) \rceil comparisons in the worst case. Another key example involves operations on balanced binary search trees (BSTs), such as search, insertion, or deletion, which take time proportional to the tree's height; in a balanced BST with n nodes, the height is O(\log n), ensuring logarithmic performance for these operations. The logarithmic bound in these algorithms derives from a modeling the divide-and-conquer process. For binary search, the running time T(n) satisfies the recurrence T(n) = T\left(\frac{n}{2}\right) + \Theta(1), where \Theta(1) accounts for the constant-time midpoint computation and comparison. Unrolling this recurrence yields T(n) = \Theta(\log n), as the depth of is the number of halvings needed to reduce n to 1, which is \log_2 n. Similarly, for balanced BST operations, the height h follows h = 1 + h/2 on average, solving to h = O(\log n). The of logarithmic time is independent of the logarithm's base, as changing from base b to base k (both greater than 1) introduces only a constant factor: \log_b n = \frac{\log_k n}{\log_k b}. Thus, O(\log_b n) = O(\log_k n) for any valid bases, allowing flexibility in notation without altering the . This property underscores why \log n without a specified base is standard in big-O notation for such algorithms.

Linear and Quasilinear Time

Linear Time

Linear time complexity, denoted as O(n), refers to algorithms whose running time is directly proportional to the input size n, meaning the number of operations grows linearly as n increases. This makes linear time a fundamental baseline for efficiency in design, as it ensures predictable scaling without excessive overhead for small to moderate inputs. The time complexity function for such algorithms satisfies T(n) = O(n), often realized through a single loop that iterates exactly n times or a constant multiple thereof. For instance, a simple for-loop from 1 to n performs n iterations, each taking constant time, yielding overall linear time. Representative examples include computing the sum of elements in an of size n, which requires one pass to accumulate the total, and , which scans the array sequentially until finding a target element or reaching the end. These operations are straightforward and commonly used in practice for tasks like or element lookup in unsorted collections. Linear time algorithms are feasible and efficient for moderate input sizes, such as n up to millions on modern hardware, but they become a performance bottleneck for very large datasets, like those in processing, where even linear growth can lead to prohibitive execution times. They often pair with O(1) extra , relying on in-place modifications or fixed auxiliary variables without requiring additional storage proportional to n.

Quasilinear Time

Quasilinear time complexity refers to algorithms with a running time of O(n \log n), where n is the input size and the logarithm is typically base 2, though the base affects only a constant factor. This bound arises in scenarios where operations scale linearly with the input but include an additional logarithmic multiplier, often due to divide-and-conquer strategies or tree-based structures. The term "" highlights its proximity to linear time while acknowledging the superlinear growth from the \log n term. This complexity class is particularly prevalent in comparison-based sorting algorithms, where the information-theoretic lower bound establishes that any such algorithm requires at least \Omega(n \log n) comparisons in the worst case. The bound derives from the need to distinguish among n! possible input permutations using binary decisions from comparisons; since \log_2(n!) = \Theta(n \log n) by (n! \approx \sqrt{2\pi n} (n/e)^n), at least this many comparisons are necessary to identify the sorted order. Key examples include , which runs in O(n \log n) time in all cases via a divide-and-conquer approach that recursively splits the and merges sorted halves; , which achieves O(n \log n) on average by partitioning around a ; and , which maintains a structure to extract elements in sorted order, also in O(n \log n) worst-case time. For , the recurrence T(n) = 2T(n/2) + \Theta(n) falls under case 2 of the , where the work per level is \Theta(n) across \log n levels, yielding T(n) = \Theta(n \log n). Similar recurrences govern the other algorithms, confirming their bounds. In practice, quasilinear algorithms handle inputs up to millions of elements efficiently, as the \log n factor grows slowly—for instance, \log_2 10^6 \approx 20, resulting in roughly 20 linear passes. Their performance benefits from hardware optimizations like cache locality in or in-place operations in , making them suitable for real-world where n reaches $10^6 to $10^7 without excessive slowdowns compared to linear time methods.

Polynomial Time Complexities

General Polynomial Time

In , polynomial time describes algorithms whose worst-case running time is bounded by a function of the input size n, formally expressed as T(n) = O(n^k) for some constant integer k \geq 0 . This bound ensures that the computation grows at a rate that remains practical for large inputs, distinguishing it from slower-growing sublinear complexities or faster-growing superpolynomial ones. The constant k captures various subcases, such as constant time when k=0, linear time when k=1 (e.g., scanning a list once) or quadratic time when k=2 (e.g., checking all pairs in a set). The growth rate of polynomial-time algorithms is considered reasonable for small values of k, such as 2 or 3, where even inputs of size n=1000 yield manageable computation times on contemporary hardware— for instance, n^3 \approx 10^9 operations, executable in seconds. However, for larger k, such as 10 or more, the time explodes rapidly; n=10 might be feasible, but n=100 could require $10^{20} operations, far beyond current capabilities. This scalability underscores why polynomial time is viewed as a for tractability in design. A classic example is the naive algorithm for multiplying two n \times n matrices, which performs O(n^3) scalar multiplications and additions by iterating over all triples of indices. Historically, Alan Cobham's 1965 thesis proposed that polynomial-time computations align with what is feasibly computable across different machine models, independent of specific hardware details, establishing a foundational notion of efficient solvability. This perspective, known as Cobham's thesis, has influenced the definition of the P for decision problems resolvable in polynomial time.

Complexity Classes

In , complexity classes provide a formal framework for classifying decision problems based on the computational resources required to solve them, particularly time bounds on . The class , standing for "polynomial time," consists of all decision problems that can be solved by a deterministic in a number of steps bounded by a in the length of the input. This class captures the notion of efficient or feasible computation, as polynomial-time algorithms scale reasonably with input size compared to superpolynomial alternatives. Formally, a language L \subseteq \{0,1\}^* is in P if there exists a deterministic M and a p such that for every input string x, the machine M halts after at most p(|x|) steps and accepts x if and only if x \in L. This definition ensures that membership in P can be decided efficiently for all inputs, with the polynomial bound guaranteeing that the running time grows manageably, such as O(n^k) for some k. Representative examples include determining graph connectivity via , which runs in O(V + E) time where V is the number of vertices and E the number of edges, and computing shortest paths in a with non-negative weights using , which achieves O((V + E) \log V) time with a . The class P relates to the broader time hierarchy defined by the deterministic time complexity classes DTIME(t(n)), which contain languages decidable in at most t(n) steps. The time hierarchy theorem establishes strict inclusions among these classes; for time-constructible functions f and g where f(n) \log f(n) = o(g(n)), DTIME(f(n)) is a proper of DTIME(g(n)). This implies hierarchies like DTIME(n) \subset DTIME(n \log n) leading into P, which itself is properly contained in larger classes like exponential time. A key structural property of P is its closure under composition: if problems L_1 and L_2 are in P, solvable by machines running in polynomials p_1 and p_2 respectively, then the composed problem—where an instance of L_1 is reduced to L_2 via a polynomial-time —is also in P, with running time bounded by a polynomial composition of p_1 and p_2.

Superpolynomial Time

Quasi-Polynomial Time

Quasi-polynomial time, often denoted as , describes algorithms whose running time is bounded by $2^{O((\log n)^k)} for some constant k \geq 1, where n is the input size. This class equivalently encompasses running times of the form n^{O(\log n)}, which includes polylogarithmic factors in the exponent. Such complexities arise in , bridging polynomial-time solvability and exponential-time barriers. The growth rate of quasi-polynomial functions exceeds any polynomial n^{O(1)}, as the exponent (\log n)^k grows without bound albeit slowly, but remains subexponential, falling short of $2^{\Omega(n^\epsilon)} for any \epsilon > 0. Formally, a representative bound is T(n) = 2^{(\log n)^c} for constant c, capturing algorithms that scale efficiently for practical input sizes yet theoretically surpass polynomial limits. Prominent examples include László Babai's algorithm for graph isomorphism, which solves the problem in quasi-polynomial time using group-theoretic techniques on combinatorial structures. Another is the computation of discrete logarithms in finite fields of small characteristic, achieved via index calculus methods refined to quasi-polynomial complexity. In algorithmic game theory, approximate Nash equilibria in certain games can be found in quasi-polynomial time, leveraging randomization derandomization. In terms of hardness, problems like reside in QP but are not known to be in , with conditional lower bounds suggesting they require superpolynomial time under assumptions like the . This positions QP as a class where some computationally hard problems admit practical algorithms, though distinguishing QP from remains open, with implications for problems.

Subexponential and Exponential Time

Subexponential Time

In , subexponential time describes algorithms whose running time T(n) satisfies T(n) = 2^{o(n)}, where n is the input size. This bound ensures that the running time grows asymptotically slower than $2^{c n} for any constant c > 0, placing it between and complexities. Unlike quasi-polynomial time, which requires $2^{O((\log n)^k)} for some constant k and represents even slower growth, the $2^{o(n)} bound is broader, encompassing a wider range of super-quasi-polynomial but sub-exponential functions. A related definition, commonly employed in analyses tied to the (), characterizes subexponential time as $2^{O(n^\epsilon)} for some fixed \epsilon with $0 < \epsilon < 1. This form captures algorithms where the exponent grows sublinearly but polynomially, often arising in fine-grained complexity studies. The $2^{o(n)} definition permits slower overall growth in certain cases, as the exponent o(n) can approach linear rates more closely than the fixed sublinear polynomial O(n^\epsilon) in the second bound, though both remain asymptotically below full exponential $2^{\Theta(n)}. For instance, a 3-SAT solver operating in $2^{O(n^{0.4})} time would exemplify subexponential performance under the second definition, though no such algorithm is known, and the posits that 3-SAT requires \Omega(2^{c n}) time for some c > 0. The , as a consequence, implies that many NP-complete problems lack subexponential-time solutions. Subexponential time plays a key role in cryptography, where hardness assumptions for problems like (LWE) rely on resistance to subexponential attacks, enabling constructions of secure such as public-key under the assumption that LWE cannot be solved in $2^{o(n)} time. These assumptions ensure that even advanced algorithms running in $2^{O(n^\epsilon)} for \epsilon < 1 remain computationally infeasible for practical parameter sizes.

Exponential Time

In computational complexity theory, exponential time describes algorithms whose worst-case running time grows as $2^{O(n)}, where n is the input size, or more precisely, T(n) = 2^{\Theta(n)}. This class includes problems solvable by deterministic Turing machines in time bounded by c \cdot 2^{k n} for constants c > 0 and k \geq 1. Such growth distinguishes exponential time from polynomial time, marking a threshold for computational intractability in practice. Classic examples illustrate this complexity. The Held-Karp dynamic programming algorithm for the exact Traveling Salesman Problem (TSP), which finds the shortest tour visiting n cities, requires O(2^n n^2) time by enumerating subsets of cities and computing minimum paths. Similarly, the meet-in-the-middle approach for the exact —determining if a of n integers sums to a target—achieves O(2^{n/2} n) time, still exponential since $2^{n/2} = (\sqrt{2})^n. The rapid doubling of computation with each additional input element limits feasibility to small n, typically under 40 on standard hardware, beyond which even optimized implementations become prohibitive due to memory and time constraints. The Exponential Time Hypothesis (ETH), proposed by Impagliazzo and Paturi, strengthens this by conjecturing that 3-SAT on n variables requires $2^{\Omega(n)} time, with no $2^{o(n)} algorithm existing; this implies exponential hardness for numerous NP-complete problems under ETH. Time hierarchy theorems further delineate these classes, proving that for time-constructible functions t(n) with t(n) \log t(n) = o(s(n)), DTIME(t(n)) \subsetneq DTIME(s(n)), thus separating exponential time from both and stricter exponential bounds like $2^{2n}. Subexponential algorithms, running in $2^{o(n)}, occasionally yield approximations but rarely exact solutions for these problems.

Super-Exponential Time Complexities

Factorial Time

Factorial time complexity describes algorithms whose worst-case running time T(n) satisfies T(n) = O(n!), where n is the input size and n! denotes the function. This class is super-exponential, as n! grows faster than any pure c^n for constant c > 1, formally expressed as n! = \omega(c^n). Algorithms in this category are rarely used except in theoretical contexts or for exhaustive enumeration problems. Common examples include generating all permutations of n distinct elements, which inherently requires examining \Theta(n!) arrangements. Another prominent case is the brute-force approach to solving the exact traveling salesman problem (TSP), where all possible tours among n cities must be evaluated, leading to a \Theta(n!) time requirement (or \Theta((n-1)!/2) for undirected graphs, asymptotically equivalent). To understand the explosive growth of factorial time, Stirling's approximation provides a useful asymptotic : n! \approx \sqrt{2 \pi n} \left( \frac{n}{e} \right)^n This reveals that n! is dominated by the term (n/e)^n, highlighting its super-exponential nature. More precisely, the running time satisfies T(n) = \Theta(n!), and asymptotically, n! \sim 2^{n \log n} in the sense that \log(n!) = \Theta(n \log n). Due to this rapid escalation—for instance, $12! \approx 4.79 \times 10^8— time algorithms are practical only for tiny inputs, such as n < 12 in the TSP, where might take seconds on modern but becomes infeasible beyond that. Double exponential time, involving iterated exponentials like $2^{2^n}, represents an even more prohibitive growth rate.

Double Exponential Time

Double exponential time complexity refers to the running time of algorithms bounded above by a function of the form $2^{2^{p(n)}}, where p(n) is a polynomial in the input size n. This class, often denoted EEXP or 2-EXP, captures computations that require time exponential in an exponential quantity. A related subclass, EE, uses a linear exponent in the outer exponential, corresponding to DTIME($2^{2^{O(n)}}). Such rapid growth arises in problems involving iterated exponentials or towers of height two, exemplified by the naive recursive evaluation of the at levels where the recursion depth leads to double call counts. For instance, computing values in the Ackermann hierarchy up to A(3, n) yields time, but extensions toward higher fixed levels approach double behavior through nested recursions. In , proof search strategies like cylindrical algebraic decomposition (CAD) for over real closed fields exhibit double time complexity in the number of variables, as the algorithm constructs cell decompositions that explode in size with each projection step. The growth of double exponential functions vastly outpaces factorial time, as $2^{2^n} involves iterated exponentiation while n! is merely a repeated product; for example, even modest n like 5 yields $2^{2^5} = 2^{32} \approx 4 \times 10^9, exceeding $5! = 120, with the gap widening dramatically thereafter. Formally, a typical bound is T(n) = 2^{\Theta(2^n)}, reflecting the primitive recursive nature of functions within ELEMENTARY, where double exponential time forms a foundational level. In , double exponential time plays a key role in separating levels of the ELEMENTARY , which is the union over k of classes requiring time bounded by k-fold exponential towers of polynomials; EEXP marks the second level, distinguishing problems solvable with bounded recursion depth from those needing taller towers.

References

  1. [1]
    [PDF] Lecture 13: Time Complexity - People | MIT CSAIL
    (We will only consider decidable languages now!) Definition: The running time or time complexity of M is the function T : ℕ → ℕ such that. T(n) = maximum ...
  2. [2]
    [PDF] Time-Complexity.pdf
    The running time (or time complexity) of M is the function / : o -→ o such that /(n) is the maximum number of steps that M uses on any input string of length n.
  3. [3]
    [PDF] Three time complexity functions Three time complexity functions
    The average-case time complexity of the algorithm is the function defined by an average number of operations performed, taken across all instances of size n. '.
  4. [4]
    [PDF] CS 311 1 Time Complexity 2 Notation - CSUSM
    The Time Complexity of an Algorithm is the running time needed by an algo- rithm expressed as a function of the input size. Input Size can be many elements to ...
  5. [5]
    Time Complexity - UMSL
    A brief review of basic time complexity and asymptotic bounds. The exact running time of an algorithm is usually a complex expression.
  6. [6]
    [PDF] Asymptotic Time Complexity and Intro to Abstract Data Types
    Page 3. The Analysis Framework. • Time efficiency (time complexity): indicates how fast an algorithm. runs. • Space efficiency (space complexity): refers to ...
  7. [7]
    ON THE COMPUTATIONAL COMPLEXITY OF ALGORITHMS
    I. Introduction. In his celebrated paper [1], A. M. Turing investigated the computability of sequences (functions) by mechanical procedures and showed.
  8. [8]
    [PDF] Analysis of Algorithms - Computer Science, UWO
    Time Complexity. • The time complexity of an algorithm is denoted as t(n) or f(n), where n is the size of the input. • The time complexity is non-decreasing ...
  9. [9]
    [PDF] the intrinsic computational difficulty of functions 25
    THE INTRINSIC COMPUTATIONAL DIFFICULTY OF FUNCTIONS. ALAN COBHAM. I.B.M. Research Center, Yorktown Heights, N. Y., U.S.A.. The subject of my talk is perhaps ...
  10. [10]
    [PDF] The Origins of Computational Complexity
    Mar 31, 2010 · How did the work of Blum, Cobham, Hartmanis, and Rabin in the late 1950's and early 1960's influence the existence of the field of computational ...
  11. [11]
    [PDF] 13.7 Asymptotic Notation - MIT OpenCourseWare
    Big O is the most frequently used asymptotic notation. It is used to give an ... The lower bound can also be described with another special notation “big Omega.
  12. [12]
    [PDF] UNDERSTANDING PROGRAM EFFICIENCY - MIT OpenCourseWare
    ◦ O(1 + 3n + 1) = O(3n + 2) = O(n). ▫ overall just O(n). 6.0001 LECTURE 10. 31. Page 32. NESTED LOOPS. ▫ simple loops are linear in complexity. ▫ what about ...
  13. [13]
    Complexity and Big-O Notation - cs.wisc.edu
    one read; one write (of a primitive type). Some methods perform the same number of operations every time they are called. For example, the size ...Missing: source | Show results with:source
  14. [14]
    Complexity
    ### Summary of Constant Time: O(1)
  15. [15]
    [PDF] Some time-complexity classes - CS@Cornell
    O(1). Constant time.​​ Algorithms whose running time does not depend on the size of the input; there is an upper bound on the number of basic steps executed. ...
  16. [16]
    Worst-Case Complexity Analysis
    Mar 15, 2022 · Array indexing a[i] is O(1). That doesn't mean that everything we do with array elements will be O(1). For example, if I write int array[100]; ...Missing: citation | Show results with:citation
  17. [17]
    Hash Tables—Theory and Practice
    The number of steps is O(1). The worst-case search time for a hash table is O(n), which can happen when all keys are stored in the same bucket. Nevertheless ...
  18. [18]
    [PDF] CacheBleed: A Timing Attack on OpenSSL Constant Time RSA
    In this paper we show that scatter-gather is not constant-time. We implement a cache timing attack against the scatter-gather implementation used in the modular ...
  19. [19]
    [PDF] Hardware Support for Constant-Time Programming
    Nov 1, 2023 · Constant-time programming have two main rules to mitigate cache-based side channels: i) no branch on secrets (control-flow secu- rity) and ii) ...Missing: variations | Show results with:variations
  20. [20]
    [PDF] CS231 - Fall 2017 Algorithm Analysis (also called Asymptotic ...
    Time Complexity: O (1)​​ The most common base for the logarithm function in computer science is 2 as computers store integers in binary. That is, for us, log n = ...
  21. [21]
    [PDF] UNIT 5B Binary Search
    We say Binary Search has complexity O(log n). 22. “floor”. Page 42. O(log n) (“logarithmic time”). 24 n. (amount of data). Number of. Operations log. 2 n log.
  22. [22]
    [PDF] 22C:21 Lecture Notes Running time of Binary Search
    Sep 3, 2008 · Define the running time of an algorithm (or a function or a program) as the number of basic operations executed by the algorithm, when provided ...
  23. [23]
    Basic Algorithms: Lecture #15
    In a sense that we will not make precise, binary search trees have logarithmic performance since `most' trees have logarithmic height. Nonetheless we know that ...
  24. [24]
    [PDF] CSE 3500 Algorithms and Complexity – Fall 2016 Lecture 7
    The example of binary search: Let T(n) be the run time of binary search on any input of size n. Then, we can express T(n) as follows: T(n) = T(n/2) + 1 ...
  25. [25]
    Big-Oh for Recursive Functions: Recurrence Relations
    Apr 17, 2003 · This web page gives an introduction to how recurrence relations can be used to help determine the big-Oh running time of recursive functions.
  26. [26]
    [PDF] EXAMPLES and CALCULATIONS
    This means that their run time increases slowly, with the log of the size of the input. For example: • Binary search of a sorted list or binary search tree (BST).Missing: definition | Show results with:definition
  27. [27]
    [PDF] Asymptotic Running Time of Algorithms - Cornell: Computer Science
    For asymptotic complexity, base of logarithms does not matter. • Let us show that log2(n) = O(logb(n)) for any b > 1. • Need to find a witness pair (c, n0) ...
  28. [28]
    Algorithm Efficiency
    The order of log n is independent of the base taken loga n = O(logb n) for all a, b > 1; Logarithms grow more slowly than any power of n log n = O(na) for any a ...
  29. [29]
    [PDF] Class Notes, CS 3137 1 Measuring the Running Time of Programs
    – If T(N) = logN, then we can say T(N) is a logarithmic time algorithm. ... can always convert a logarithm in one base to any other base using a constant.
  30. [30]
    Algorithmic Complexity - UMSL
    Constant Time: O(1). An algorithm is said to run in constant time if it requires the same amount of time regardless of the input size. Examples: array ...
  31. [31]
    Reasoning about Algorithmic Complexity
    The function describing how much time the algorithm needs is called the time complexity ... O(1) means constant time . Polynomial time means ... examples:.Asymptotic Analysis · Order Of Growth & ``big-O''... · Analyzing Running Times Of...
  32. [32]
  33. [33]
    1.4 Analysis of Algorithms
    Mar 18, 2020 · Your algorithm should run in linear time, use O(1) extra space, and may not modify the original array. Hint: pointer doubling. Finding common ...
  34. [34]
    5. Analyzing Complexity | CS 2110
    Sep 9, 2025 · Let's combine the concept of loop invariants from last lecture with the concept of time complexity to develop and analyze two search algorithms.Missing: implications | Show results with:implications
  35. [35]
    [PDF] On Quasilinear Time Complexity Theory
    This is also enough to give BQL[BQL[C]] = BQL[C]. Definition 2.3. The quasilinear time hierarchy is defined by: Pql. 0 = Qql. 0 = ∆ ql. 0 = DQL, and for k ...
  36. [36]
    [PDF] Comparison-based Lower Bounds for Sorting
    In this lecture we discuss the notion of lower bounds, in particular for the problem of sorting. We show that any deterministic comparison-based sorting ...
  37. [37]
    Introduction to Algorithms
    ### Summary of References to Quasilinear Time, O(n log n), Sorting Algorithms, and Master Theorem
  38. [38]
    [PDF] 1 Solving recurrences
    Last class we introduced recurrence relations, such as T(n) = 2T(bn/2c) + n. Typically these reflect the runtime of recursive algorithms.
  39. [39]
    Analysis of Algorithms
    O(n log n)​​ Increases more rapidly than linear algorithms but not as fast as the quadratic ones.
  40. [40]
    Computational Complexity Theory
    Jul 27, 2015 · A complexity class can now be defined to be the set of problems for which there exists a decision procedure with a given running time or running ...
  41. [41]
    [PDF] 1 Algorithms 2 Polynomial time
    2 Polynomial time. The complexity class P consists of all problems solvable in polynomial time. ( Technically, it is a. class of languages)
  42. [42]
    [PDF] AN OVERVIEW OF COMPUTATIONAL COMPLEXITY
    Cobham pointed out that the class was well de- fined, independent of which computer model was chosen, and gave it a characterization in the spirit of recursive ...
  43. [43]
    Complexity Zoo:P
    Jul 18, 2025 · P: Polynomial-Time. The class that started it all. The class of decision problems solvable in polynomial time by a Turing machine.P/poly · PCTC · k {\displaystyle k} {\displaystyle k} · PFCHK(t(n))
  44. [44]
    Complexity Classes - UMBC
    Nov 30, 2003 · Complexity Class Brief Definitions. Language Classes. Class: Brief description P: the set of languages accepted by deterministic Turing machines ...
  45. [45]
    Graph Traversal: BFS and DFS
    The time complexity for BFS is O ( V + E ) (linear in the size of the graph) because you need to visit each edge once and only once, and each node is added to ...
  46. [46]
    Running Time of Dijkstra's algorithm - Virtual Labs
    The inner loop has decreaseKey() operation which takes O(LogV) time. So overall time complexity is O(E+V)*O(LogV) which is O((E+V)*LogV) = O(ELogV) ...
  47. [47]
    [PDF] Lecture 5: Hierarchy Theorems
    Jan 26, 2022 · Theorem 3 (Time Hierarchy). If r, t are time-constructible functions satisfying r(n) log r(n) = o(t(n)), then DTIME(r(n)) ( DTIME(t(n)).
  48. [48]
    [PDF] Computational Complexity 1 Introduction - Duke Computer Science
    Speci cally, the class is closed under addition, multiplication and composition. The growth-rate of polynomials allows to consider as e cient all procedures.
  49. [49]
    [1512.03547] Graph Isomorphism in Quasipolynomial Time - arXiv
    Dec 11, 2015 · Authors:László Babai. View a PDF of the paper titled Graph Isomorphism in Quasipolynomial Time, by L\'aszl\'o Babai. View PDF. Abstract:We show ...
  50. [50]
    An Improved Quasi-Polynomial Algorithm for Approximate Well ...
    Nash equilibrium in time nO(log n/ε2), for any ε > 0, in. games with n pure strategies per player. Such a running. time is referred to as quasi-polynomial. ...
  51. [51]
    [PDF] Parity Games: Zielonka's Algorithm in Quasi-Polynomial Time - LaBRI
    Apr 29, 2019 · [8] exists. The running time of our algorithm is quasi-polynomial, and the space complexity is quadratic (more precisely, O(n · h), where n is ...
  52. [52]
    [1710.04574] Graph isomorphisms in quasi-polynomial time - arXiv
    Oct 12, 2017 · Babai has recently shown how to solve these problems -- and others linked to them -- in quasi-polynomial time, i.e. in time \exp\left(O(\log n) ...
  53. [53]
    [PDF] A Framework for ETH-Tight Algorithms and Lower Bounds in ... - arXiv
    Dec 17, 2020 · Abstract. We give an algorithmic and lower-bound framework that facilitates the construction of subexponential algorithms and matching ...
  54. [54]
    [PDF] On the Complexity of k- SAT - UCSD CSE
    Define ETH (Exponential-Time Hypothesis) for k-SAT as follows: for k 3, sk>0. In other words, for k 3, k-SAT does not have a subexponential-time algorithm. In.
  55. [55]
    [PDF] Continuous LWE is as Hard as LWE - arXiv
    Nov 2, 2022 · For one, assuming the sub-exponential hardness of LWE, we show that density estimation of g = (log n)1+ǫ Gaussians cannot be done in polynomial ...
  56. [56]
    [PDF] Statistical mechanics of the vertex-cover problem
    The most interesting problems lie on the interface between polynomial and exponential running time. ... time-complexity of this approach is O(2N ). Early ... N=40.
  57. [57]
    [PDF] Which Problems Have Strongly Exponential Complexity? - UCSD CSE
    We can make the complexity implications of sub-exponential algorithms for the above problems more precise 516 IMPAGLIAZZO, PATURI, AND ZANE Page 6 using ...
  58. [58]
    The tight deterministic time hierarchy - ACM Digital Library
    In order to get a strong hierarchy, where the lower bounds hold for almost all input lengths, we first strengthen. Theorem 2. Definition. DSPACETINEk(S, t) is ...
  59. [59]
    Complexity Zoo:E
    Nov 6, 2024 · EEXP: Double-Exponential Time ... Equals DTIME(22p(n)) for p a polynomial. Also known as 2-EXP. Contains EE, and is contained in NEEXP.
  60. [60]
    An inherently iterative computation of ackermann's function
    An iterative algorithm for computing A(i,n) is presented. It has O(i) space complexity and O(iA(i,n)) time complexity, both of which are much smaller than the ...Missing: evaluation | Show results with:evaluation
  61. [61]
    The complexity of quantifier elimination and cylindrical algebraic ...
    This paper has two parts. In the first part we give a simple and constructive proof that quantifier elimination in real algebra is doubly exponential.
  62. [62]
    Exponential vs. Factorial
    We are interested in comparing repeated applications of an exponential function, with any base, and the factorial func- ... When are two iterated exponential se-.<|control11|><|separator|>