Fact-checked by Grok 2 weeks ago

Karmarkar's algorithm

Karmarkar's algorithm is an for solving problems, developed by mathematician at Bell Laboratories and published in 1984, which achieves a proven polynomial-time complexity of O(n3.5 L) arithmetic operations, where n denotes the problem dimension and L the input precision in bits. The approach transforms the standard-form linear program—maximizing cTx subject to Ax = b and x ≥ 0—into a sequence of approximate optimization steps within a transformed , using projective rescaling to maintain strict interior feasibility and a logarithmic barrier potential to guide progress toward the optimum. This marked a pivotal theoretical and practical advance over prior polynomial-time methods like the ellipsoid algorithm, which, despite proving resolvability in polynomial time in 1979, suffered from poor constant factors rendering it inefficient for real-world use. The algorithm's core innovation lies in its potential-reduction framework, where each reduces a by a fixed factor (typically via parameter γ ≈ 0.99), ensuring in O(L log(1/ε)) steps for ε-optimal solutions, combined with per- costs dominated by solving a Newtonian via Cholesky . Initial implementations and analyses suggested substantial speedups over the prevalent method for large-scale problems, prompting widespread interest and the evolution of interior-point techniques into cornerstone solvers in . While practical performance varied—often trailing revised on sparse instances due to denser linear algebra—its guaranteed worst-case efficiency underscored linear programming's membership in P and influenced broader paradigms.

Historical Context

Pre-Karmarkar Linear Programming

The simplex algorithm, developed by George Dantzig in 1947, served as the primary practical method for solving linear programming problems for nearly four decades. Originating from Dantzig's work on resource allocation for the U.S. Air Force during and after World War II, the algorithm iteratively moves along the edges of the feasible polytope toward an optimal vertex, leveraging a tableau representation to perform pivot operations. In practice, it exhibited average-case polynomial performance and enabled solutions to problems with thousands of variables on early computers, finding widespread adoption in operations research for applications such as production planning and transportation scheduling. Despite its empirical efficiency, the simplex method lacked a proven polynomial-time worst-case guarantee, with counterexamples like the Klee-Minty hypercube demonstrating exponential iteration counts in contrived instances. This theoretical limitation spurred efforts in the 1970s to develop algorithms with rigorous polynomial complexity, amid growing computational demands from industrial sectors. Linear programming models were increasingly applied to optimize logistics networks, supply chain distribution, and early telecommunications routing, where problem sizes often exceeded the reliable performance bounds of simplex implementations on available hardware. These applications highlighted the need for methods that could handle larger-scale, sparse constraint systems without degenerate pivoting delays. In 1979, introduced the , marking the first algorithm proven to solve in time, specifically with O(n²L) iterations where n is the number of variables and L relates to the input size in bits. The method constructs a sequence of shrinking ellipsoids containing the , using subgradient information from violated constraints to update the bounding set until convergence to an optimal point. However, its high constant factors, sensitivity to numerical precision due to ellipsoid volume computations, and dependence on a deep degree rendered it unsuitable for practical large-scale problems, outperforming only in very small or specially structured cases. Thus, remained the dominant solver, underscoring a gap between theoretical advances and deployable efficiency for real-world optimization in fields like and .

Development at Bell Labs

Narendra Karmarkar, an Indian mathematician who joined Bell Laboratories in 1983, developed the algorithm while employed there as part of the organization's mathematical research efforts. The work was carried out internally at the labs in , focusing on advancing computational methods for optimization problems relevant to telecommunications planning and . In late 1984, publicly announced the algorithm through press releases and media coverage, emphasizing its potential for dramatically accelerating solutions to large instances. The company claimed it achieved up to a 50-fold on benchmark test problems compared to the prevailing simplex method, particularly for problems with thousands of variables and constraints previously deemed computationally prohibitive. Karmarkar's initial formal , titled "A new polynomial-time for ," appeared in the December issue of Combinatorica, providing a rigorous proof of its polynomial-time complexity bounded by O(n3.5 L), where n denotes the number of variables and L the total of the input data. Empirical validations accompanying the announcement demonstrated the method's efficacy on substantial real-world-scale problems, such as those involving over 10,000 constraints, which the could not resolve in feasible timeframes on contemporary hardware.

Algorithm Fundamentals

Core Principles

Karmarkar's algorithm represents a paradigm shift in linear programming solvers by employing interior-point methods, which navigate through the interior of the feasible polytope rather than traversing its boundary vertices as in the simplex method. The simplex algorithm, developed by George Dantzig in 1947, pivots from one basic feasible solution (vertex) to adjacent vertices, potentially requiring exponentially many steps in the worst case for problems with n variables and m constraints. In contrast, Karmarkar's approach, introduced in 1984, initiates from a strictly feasible interior point and follows a trajectory that asymptotically approaches the optimal face without touching the boundary until convergence, enabling polynomial-time complexity guarantees. Central to the method is the projective scaling transformation, which homogenizes the and maps it into a standard unit , facilitating symmetric treatment of and promoting movement toward the "center" of the transformed space. This transformation, applied iteratively, distorts the geometry to recenter the current iterate near the analytic center of the , allowing a deep feasible step that reduces while maintaining strict feasibility. The projective framework draws from geometric insights into polytopes, ensuring that each iteration scales variables inversely to residual slacks, akin to a barrier against violations without explicit penalty terms. Progress is rigorously controlled via a primal potential function, typically of logarithmic form \phi(x) = - \sum_{i=1}^m \log v_i^k + \rho \|c\|^T x, where v^k denotes slack residuals and \rho is a large parameter scaling the objective. Each iteration guarantees a constant fractional reduction in this potential—approximately 0.02 to 0.2 depending on step size \alpha (often set near 0.25)—leading to logarithmic convergence in the number of steps, bounded by O(n L \log(1/\epsilon)) to achieve \epsilon-optimality, with L the bit length of input data. This potential embodies barrier-like principles, penalizing proximity to boundaries through logarithmic terms that grow unbounded as slacks approach zero, and its consistent decrease provides a Lyapunov-style measure of global convergence independent of local geometry. The algorithm's foundations prefigure broader interior-point theory, including self-concordant barrier functions for , where the negative log-barrier induces a central path of Newton steps tracing minimizers of perturbed objectives. Karmarkar's projective variant approximates this path-following via scaled Newton directions in transformed coordinates, establishing causal links between interior geometry, potential reduction, and polynomial solvability that underpin subsequent primal-dual and barrier methods.

Projective Scaling and Potential Reduction

Projective rescaling in Karmarkar's algorithm applies a nonlinear to the current feasible point, mapping it to the barycenter of the transformed —typically the uniform point (1/n, \dots, 1/n) in a homogenized space where variables sum to unity—thus centering the iterate and symmetrizing the constraints geometrically. This step exploits to rescale the problem instance, ensuring subsequent computations occur relative to a canonical position that equalizes variable scales and facilitates isotropic analysis of the . The algorithm then minimizes a logarithmic barrier , \phi(x) = n \log(c^T x) - \sum_{j=1}^n \log x_j, which encodes objective progress via the first term while the barrier term -\sum \log x_j enforces strict interiority by diverging as any x_j approaches zero. In the rescaled space, a direction is computed as the projected minimizing this potential subject to linearized constraints, approximating the central and yielding a search direction that reduces while respecting feasibility. A , often around 0.1 to 0.5, scales the step to guarantee a fixed relative decrease in \phi, such as at least 0.25 per in analyzed variants. This interior-point strategy causally evades degeneracy by maintaining iterates away from boundaries via the barrier's penalization, preventing the algorithm from stalling on low-dimensional faces or requiring pivots through ill-conditioned vertices as in boundary-following methods; instead, the smooth on the potential traces a approximating the analytic , inherently robust to rank deficiencies. Iteration complexity derives from the potential's bounded decrease: starting from an \phi(x_0) scaling with input L (the of coefficients), and terminating when c^T x_k < 2^{-2L} c^T x_0, yields at most O(n L) steps, where the factor n arises from the potential's dimension dependence and L captures the problem's condition number through logarithmic requirements.

Detailed Procedure

Mathematical Formulation

Karmarkar's algorithm solves linear programming problems, typically formulated in inequality standard form as maximizing \mathbf{c}^T \mathbf{x} subject to A \mathbf{x} \leq \mathbf{b} and \mathbf{x} \geq \mathbf{0}, where A is an m \times n matrix with m < n, \mathbf{c} \in \mathbb{R}^n, and \mathbf{b} \in \mathbb{R}^m_+. To convert inequalities to equalities, nonnegative slack variables \mathbf{s} \geq \mathbf{0} are introduced, yielding A \mathbf{x} + I \mathbf{s} = \mathbf{b} with \mathbf{x}, \mathbf{s} \geq \mathbf{0}. The algorithm requires transforming the problem into a canonical form suitable for projective transformations, where the feasible region lies within the standard simplex \{\mathbf{y} \geq \mathbf{0} \mid \mathbf{e}^T \mathbf{y} = 1\}, with \mathbf{e} the all-ones vector. This is achieved through homogenization by adjoining a scaling variable t \geq 0: the original problem is equivalent to maximizing t subject to A (\mathbf{x}/t) \leq \mathbf{b}/t, \mathbf{e}^T (\mathbf{x}/t) = 1, and \mathbf{x}/t \geq \mathbf{0}, assuming the original optimum is bounded. Scaling ensures the right-hand side becomes the all-ones vector, embedding the scaled feasible set \{\mathbf{z} \geq \mathbf{0} \mid A' \mathbf{z} \leq \mathbf{1}, \mathbf{e}^T \mathbf{z} = 1\} inside the simplex, where A' incorporates the adjusted constraints. Theoretical guarantees rely on strict feasibility, meaning there exists \mathbf{x}^0 > \mathbf{0} such that A \mathbf{x}^0 < \mathbf{b} in the original form (or equivalently A' \mathbf{z}^0 < \mathbf{1}, \mathbf{z}^0 > \mathbf{0}, \mathbf{e}^T \mathbf{z}^0 = 1 in the ), ensuring the relative interior of the is nonempty and allowing barrier-like potential functions to drive . Boundedness of the optimal value is also assumed, preventing unbounded rays that could hinder polynomial-time termination. Affine methods, which linearly transform the to center iterates at the origin while preserving affine structure, conceptually preceded projective approaches by emphasizing interior-point trajectories but differ in using linear rather than nonlinear (projective) maps.

Iterative Steps

The iterative procedure of Karmarkar's algorithm requires an initial strictly feasible point x^0 \geq 0 such that Ax^0 < b, ensuring positive slacks v^0 = b - Ax^0 > 0, typically found via a preliminary phase I solver or of the problem instance. This point is mapped via a projective to the interior of a standard \{z \mid e^T z = 1, z > 0\}, with the image under transformation centered at e/n = (1/n, \dots, 1/n). The iteration index is initialized as k \leftarrow 0. Each iteration computes an improving direction within the transformed space using a scaled Newton-like step that approximates the steepest descent for reducing a logarithmic \Phi(x) = -\sum \log v_i(x) - R \log(c^T x + \delta), where R > n controls the trade-off and \delta offsets for boundedness. Specifically, the residual slacks are updated as v^k \leftarrow b - A x^k, followed by forming the diagonal scaling D_v \leftarrow \diag(v_1^k, \dots, v_m^k). The direction solves the positive definite system h_x \leftarrow (A^T D_v^{-2} A)^{-1} c for maximizing local objective progress under the metric induced by the barrier D_v^{-2}. The corresponding slack direction is then h_v \leftarrow -A h_x. If h_v \geq 0 componentwise, no improving feasible direction exists, indicating optimality. Otherwise, a line search determines the maximum step preserving interiority: \alpha \leftarrow \gamma \cdot \min \{ -v_i^k / (h_v)_i \mid (h_v)_i < 0, \, i=1,\dots,m \}, with damping factor \gamma \approx 0.1 to $0.25 ensuring potential reduction by at least $1 - 1/(2R) per step and avoiding numerical issues. The update follows as x^{k+1} \leftarrow x^k + \alpha h_x, after which the point is remapped via the inverse projective transformation T^{-1}(z) = D z / (e^T D z) with D = \diag(z), and k \leftarrow k+1. Termination occurs when the computed \alpha exceeds a large threshold (indicating near-optimality), the duality gap c^T x^k + b^T y^k - (A x^k)^T y^k falls below a precision \epsilon > 0, or potential stagnation is detected after O(n \log(1/\epsilon)) steps. In implementations, heuristics such as row/column equilibration of A before solving the direction system, with rank-one updates for , or adaptive \gamma based on trial steps enhance against ill-conditioning from disparate slack magnitudes.

Illustrative Example

Consider the problem of maximizing z = 3x_1 + 2x_2 subject to $2x_1 + x_2 \leq 100, x_1 + x_2 \leq 80, x_1 \leq 40, and x_1, x_2 \geq 0. This forms a bounded in the first quadrant, with vertices including (0,0), (40,0), (40,20), (20,60), and (0,80). Karmarkar's algorithm transforms the problem into a standard form with non-negative variables summing to a constant under linear equalities, starting from an interior point like the normalized all-ones vector scaled appropriately. In each iteration, a projective transformation centers the current point at the origin of the transformed space, preserving the feasible set's structure. The search direction is then found by minimizing a approximation of the objective over an inscribed within the transformed , yielding a descent direction h that reduces the Karmarkar \Phi(x) = \sum \log x_i - \mu \log(-\nabla f(x)), where \mu controls barrier strength. The step length \alpha is selected to keep the new point interior, typically \alpha = \frac{1}{R+1} where R is the transformed polytope's "radius," ensuring a constant potential decrease per . Updating x^{k+1} = x^k + \alpha h, the process repeats, converging in O(n L \log(1/\epsilon)) iterations to an \epsilon-optimal solution, where n is variables, m constraints, and L bit length. For the example problem, after slack variable introduction and transformation, iterations move interior points toward the optimum at approximately (20,60), improving the objective from an initial ~150 to near 200.

Performance Analysis

Theoretical Complexity

Karmarkar's algorithm solves problems in time, with a total arithmetic of O(n^{3.5} [L](/page/L')), where [n](/page/N+) is the number of variables and [L](/page/L') is the total encoding the input coefficients, right-hand sides, and objective . This bound arises from O(n L) iterations, each requiring O(n^{2.5} L) operations primarily for computing the search direction via of the projected A^T D_v^{-2} A, where D_v is a diagonal derived from the current residual. The iteration count follows from the potential reduction mechanism, which employs a projective potential function \Phi(x) = - \sum_{i=1}^m \log v_i^k + n \log(n+1), decreasing by at least a fixed \delta \approx 0.02 (or improved to \approx 0.25 with \alpha = 0.5) per step through affine scaling and a damped Newton update with step length \alpha \gamma, where \gamma < 1/4 ensures descent while staying within the feasible region. Starting from a scaled interior point with initial potential O(n \log R), where R bounds the polytope size, the algorithm requires O(n \log(R / \epsilon)) steps to drive the duality gap below \epsilon, with \log(R / \epsilon) = O(L) under standard input assumptions that coefficients and bounds are polynomially representable. The per-iteration cost depends on problem data through L, which captures the Lipschitz-like conditioning via bit for matrix inversions and eigenvalue bounds in the projective , necessitating higher arithmetic to avoid numerical in the barrier term. While the framework leverages the geometry of the standard for bounded , akin to a self-concordant with scaling as O(n), the resulting linear dependence on n in iterations yields larger constants than later path-following interior-point methods achieving O(\sqrt{n} \log(1/\epsilon)) steps. Subsequent refinements, such as affine-scaling variants, have reduced the exponent to O(n^3 L) in some analyses, highlighting that Karmarkar's original bound, though groundbreaking, incorporates pessimistic factors from the homogeneous and unoptimized updates not always realized in refined implementations.

Empirical Comparisons with Simplex Method

Early implementations of Karmarkar's algorithm, particularly variants in the KORBX system developed in the late 1980s, demonstrated substantial speedups over -based solvers on large-scale problems. For instance, on NETLIB test problems with thousands of rows and columns, KORBX achieved up to 553-fold reductions in computation time compared to the solver, such as solving the ken9t instance in 59.8 seconds versus 33,147.8 seconds. These gains were most pronounced on dense problems, like scheduling models with around 50 nonzeros per column, where KORBX on eight processors yielded up to 195-fold improvements; on a Pacific Basin problem with 15,425 rows, it was 20 times faster than IBM's MPSX code. Independent tests by Adler et al. in 1986 on 30 real-world linear programs (27 to 1,151 rows, 51 to 5,533 columns) reported Karmarkar variants as 3 times faster than on average, with speedups scaling up to 8.6-fold on larger instances. However, these advantages were context-dependent, with Karmarkar's projective scaling underperforming on sparse or small problems where excels due to its exploitation of sparsity. On certain NETLIB instances like ship04l, KORBX took longer (29.9 seconds) than (25.9 seconds), as the algorithm's core operations, such as inverting dense matrices like A^T D_v^{-2} A, generate fill-in that negates sparsity benefits inherent in pivoting. Early codes also suffered numerical , particularly near degenerate where least-squares subproblems became ill-conditioned, leading to in transformed matrices and risks of infeasible paths without accurate solutions. Post-1980s evaluations confirmed these limitations, spurring approaches that combine interior-point potential reduction with simplex-like pivoting for sparsity, though pure Karmarkar implementations are rarely used today in favor of refined primal-dual interior-point methods. Verifiable studies, such as those comparing affine variants to projective methods, showed consistent superiority of interior points on dense, large problems but no universal dominance, with performance hinging on problem and rather than theoretical promises.

Patent Dispute

Filing and Granting Process

Bell Laboratories filed patent application serial number 06/725,342 on April 19, 1985, covering methods and apparatus for based on implementations of the projective scaling algorithm developed by . The application emphasized novel processes involving interior-point transformations, diagonal scaling of constraints, and reductions to achieve polynomial-time convergence in optimization. These claims distinguished the patented implementations from by specifying computational steps for handling constraints through mappings and barrier-like potential minimization. The United States Patent and Trademark Office examined the application under pre-1995 rules, issuing US Patent 4,744,028 on May 10, 1988, to inventor Narendra K. Karmarkar and assignee American Telephone and Telegraph Company. The granted patent comprised 20 claims, primarily method claims (1-15) detailing iterative scaling via diagonal matrices derived from residual vectors, computation of search directions through pseudoinverse approximations of transformed Hessians, and line searches bounded by positivity constraints on affine steps. Apparatus claims (16-20) extended coverage to hardware systems programmed for these operations, enabling efficient solving of large-scale transportation and network problems. While the theoretical core of the algorithm was disclosed in Karmarkar's December 1984 publication, rendering it non-patentable as , AT&T commercialized the specific implementations via licensing to software vendors and cross-licensing arrangements, such as with for overlapping optimization techniques. These deals facilitated revenue from proprietary software incorporating the patented steps, distinct from freely available academic variants. Under the 17-year term applicable to patents granted before , 1995, the patent expired on , 2005, after which all claimed processes entered the without maintenance fees or extensions. Post-expiration, implementations previously restricted by the patent became freely available for unrestricted use in optimization software.

Mathematical Community Response

The granting of U.S. Patent 4,744,028 for Karmarkar's algorithm in May 1988 elicited significant unease within the mathematical community, as reported in a March 12, 1989, New York Times article, where mathematicians expressed concern that patenting a pure akin to a "recipe" for optimization blurred the line between and unpatentable of natural principles. This sentiment stemmed from a longstanding view that abstract , like algorithms, constitutes rather than proprietary technology eligible for protection, a reinforced by prior rulings equating mathematical formulas to laws of . The controversy intensified because , Karmarkar's employer, had filed the patent application in April 1985 while withholding full algorithmic details from , limiting immediate academic scrutiny and implementation. Proponents of the , primarily from perspectives, argued it safeguarded 's substantial investments in refining into practical software, such as the KORBX system, thereby incentivizing corporate funding for theoretical breakthroughs that might otherwise lack commercial motivation. This protection enabled AT&T to license implementations, fostering tools that integrated the method into applications and demonstrating how patents could bridge academia and by recouping costs for high-risk mathematical R&D. Critics countered that the patent initially hindered open academic by restricting access to the precise projective scaling , prompting some researchers to develop alternative interior-point variants to circumvent licensing requirements, and raising questions about the algorithm's novelty given perceived parallels to earlier nonlinear approaches from the . Despite these barriers, the patent did not ultimately suppress the method's proliferation, as generalized interior-point techniques proliferated in academic literature post-1988, leading to polynomial-time solvers that supplanted the method in many contexts. The episode amplified broader debates on software and algorithm patentability, highlighting tensions between intellectual property incentives and the collaborative ethos of mathematics, where secrecy via patents was seen as antithetical to cumulative progress, though empirical evidence showed the innovation's core ideas disseminating freely after patent expiration in 2006.

Practical Applications

Operations Research and Optimization

Karmarkar's algorithm enabled practical solutions to large-scale problems in telecommunications network design at , where the KORBX system implemented variants of the projective scaling method on vector multi-processor hardware tailored for such computations. Developed in the mid-1980s, KORBX addressed real-world tasks, including calls across the Pacific Basin, involving thousands of variables and constraints that challenged earlier simplex-based approaches. These implementations achieved unprecedented efficiency for operational optimization, processing models in significantly reduced time compared to prior methods. The algorithm's variants extended to airline crew scheduling through adoption of the KORBX system by Delta Airlines, optimizing in high-dimensional problems typical of in transportation. Such applications demonstrated the method's viability for industrial-scale linear programs, where rapid convergence facilitated decisions in time-sensitive environments like flight operations and network capacity planning. In and contexts, early deployments at firms like leveraged the algorithm for integrated optimization of and , influencing subsequent solver developments. solvers such as CPLEX later incorporated interior-point methods evolved from Karmarkar's projective to resolve linear relaxations in mixed-integer programs, supporting extensions to in and with thousands of variables. These integrations enabled case studies from the onward, where problems solvable only after hours via were handled in minutes, marking a shift in operational scalability.

Extensions to Nonlinear Problems

The interior-point methods inspired by Karmarkar's algorithm for linear programming were extended to convex nonlinear programming through the development of self-concordant barrier functions, which provide a universal framework for encoding convex feasible sets. In 1994, Nesterov and Nemirovski introduced this theory, enabling polynomial-time algorithms for minimizing convex functions over convex domains by leveraging barriers that satisfy specific third-order differentiability and curvature conditions, thus generalizing the logarithmic barrier used in linear programming. These methods maintain the central path-following approach but adapt Newton steps to handle nonlinear objectives and constraints, achieving complexity bounds dependent on the barrier's parameters rather than problem dimensions alone. A key application of these extensions appears in (), where decision variables are matrices under linear constraints, introducing nonlinearity via the semidefiniteness requirement. Interior-point algorithms for , evolved from Karmarkar's primal-dual , solve problems with high-dimensional variables efficiently in practice and theory, with self-concordant barriers like the over the positive semidefinite cone. Such methods underpin applications in , including the formulation and solution of linear inequalities for analysis and controller , as well as in for tasks like and semidefinite relaxations of combinatorial problems. Despite these advances, extensions remain confined to problems, as the reliance on self-concordant barriers and central path ensures global only under assumptions; nonconvex nonlinear cases lack polynomial-time guarantees and are generally NP-hard, with interior-point adaptations prone to local optima without additional heuristics like regularization or multistart strategies.

Legacy and Evolutions

Influence on Interior-Point Methods

Karmarkar's 1984 polynomial-time algorithm for revitalized interest in interior-point approaches, which had previously been overshadowed by the simplex method since the , by demonstrating a theoretically efficient path through the interior of the using projective transformations and a logarithmic barrier . This breakthrough prompted an explosion of research in the late and , shifting focus from boundary-tracing methods to interior trajectories that avoid degeneracy issues inherent in vertex-based algorithms. Subsequent developments rapidly extended Karmarkar's framework, yielding primal-dual interior-point methods that simultaneously optimize primal and dual problems for enhanced and predictor-corrector variants that approximate the central path more accurately through short-step or long-step iterations. These innovations, pioneered by researchers like Renegar, , and others, improved upon the original O(n^{3.5} L) arithmetic complexity by achieving O(\sqrt{n} L) iteration bounds in path-following schemes, as refined in works by Anstreicher and contemporaries analyzing projective and potential-reduction classes. The causal linkage from Karmarkar's theoretical guarantee of time to practical efficacy enabled interior-point methods to scale to massive datasets with sparse structure, where each iteration solves Newtonian systems via factorizations that exploit problem sparsity, often converging faster than revised on structured instances like flows or economic models. This permeated education, with textbooks post-1990 integrating interior-point algorithms as core alternatives to , reflecting their empirical competitiveness on high-dimensional problems exceeding thousands of variables.

Recent Theoretical and Computational Advances

Since the early , refinements to interior-point methods (IPMs) inspired by Karmarkar's projective algorithm have emphasized warm-starting techniques to accelerate in sequential or perturbed (LP) problems. These strategies recover feasible starting points from prior solutions, reducing iteration counts by 50-60% compared to cold starts, particularly beneficial for iterative applications like or online optimization. For instance, combined perturbation and centrality correction approaches outperform individual methods across diverse LP instances, maintaining while handling problem modifications efficiently. Parallelization efforts have scaled IPMs to massive datasets, with block-structured solvers like OOPS enabling solutions for LPs with $10^9 variables by exploiting sparsity and distributing Newton-Schur computations across processors. In the , such advancements underpin barrier methods in commercial solvers like Gurobi and MOSEK, which integrate IPM cores for large-scale LP subproblems in training (e.g., support vector machines) and financial , often hybridized with for sparsity but retaining IPM dominance in dense, high-dimensional cases. These evolutions affirm the original projective framework's robustness, as modern primal-dual variants achieve O(\sqrt{n} \log(1/\epsilon)) iteration bounds without paradigm shifts beyond incremental preconditioning and GPU acceleration. Theoretical extensions have incorporated stochastic elements, with randomized IPMs for tall/wide LPs reducing per-iteration costs via sketching and sampling, suitable for under uncertainty in and . Recent stochastic IPM frameworks for conic programs () adapt barrier functions to noisy gradients, extending Karmarkar-style potential reduction to distributionally robust settings while preserving near-polynomial guarantees. Despite hybrids prevailing in niche scenarios, pure IPM implementations persist in high-performance codes for their reliable scaling on , underscoring the enduring validity of the 1984 first-principles approach amid no fundamental theoretical upheavals.

References

  1. [1]
    A new polynomial-time algorithm for linear programming
    Nov 9, 1984 · We present a new polynomial-time algorithm for linear programming. In the worst case, the algorithm requiresO(n 3.5 L) arithmetic operations onO(L) bit numbers.
  2. [2]
    [PDF] The New Linear Programming Method of Karmarkar
    KHACHIYAN published the first polynomial-time method for linear programming, the ellipsoid method. This method, though theoretically efficient, turned out to ...
  3. [3]
    A new polynomial-time algorithm for linear programming
    We present a new polynomial-time algorithm for linear programming. The running-time of this algorithm is O(n3-5L2), as compared to O(n6L2) for the ellipsoid ...
  4. [4]
    Karmarkar's Linear Programming Algorithm | Interfaces - PubsOnLine
    N. Karmarkar's new projective scaling algorithm for linear programming has caused quite a stir in the press, mainly because of reports that it is 50 times ...
  5. [5]
    [PDF] Karmarkar's Linear Programming Algorithm - John Hooker
    Jul 19, 2020 · Karmarkar's Linear Programming Algorithm. Author(s): J. N. Hooker. Source: Interfaces , Jul. - Aug., 1986, Vol. 16, No. 4 (Jul. - Aug., 1986) ...
  6. [6]
    A history of scientific computing: Origins of the simplex method
    Introduction. In the summer of 1947, when I began to work on the simplex method for solving linear programs, the first idea that occurred to me is one.
  7. [7]
    [PDF] THE (DANTZIG) SIMPLEX METHOD FOR LINEAR PROGRAMMING
    George Dantzig created a simplex algorithm to solve linear programs for planning and decision-making in large-scale enterprises. The algorithm's success led to ...
  8. [8]
    [PDF] George B. Dantzig and Systems Optimization - Stanford University
    His Air Force work led to the simplex method, which succeeded in solving what would be seen today as very small linear programs, such as a 77-variable diet ...
  9. [9]
    [PDF] Origins of the Simplex Method - DTIC
    Accordingly, I developed a variant of the algorithm without the convexity constraint (9) and arranged in the fall of 1947 to have the Bureau of Standards test ...
  10. [10]
    [PDF] Large-Scale Linear Programming - IIASA PURE
    The final chapter, 9, contains a number of applications of large-scale LP techniques to practical problems in industrial and agricultural production and ...
  11. [11]
    The Ellipsoid Method: A Survey - PubsOnLine
    In February 1979 a note by L. G. Khachiyan indicated how an ellipsoid method for linear programming can be implemented in polynomial time. This.
  12. [12]
    [PDF] The Ellipsoid (Kachiyan) Method - Stanford University
    The significant contribution of Khachiyan was to demonstrate in two papers—published in 1979 and 1980—that under certain assumptions, the ellipsoid method ...
  13. [13]
    [PDF] THE ELLIPSOID METHOD
    In February 1979 a note by L.G. Khachiyan indicated how an ellipsoid method for linear programming can be implemented in poly- nomial time. This result has ...
  14. [14]
    [PDF] Solving Real-World Linear Programs: A Decade and More of Progress
    At first these algorithms did not include a simplex algorithm; eventually, however,. I concluded that it would be useful to incorporate compu- tation in the LP ...
  15. [15]
    Karmarkar's Projective Scaling Algorithm: The Practical Revolution ...
    Jun 2, 2025 · Narendra Karmarkar arrived at Bell Laboratories in 1983 with a background in both pure mathematics and practical algorithm development. Bell ...
  16. [16]
    BREAKTHROUGH IN PROBLEM SOLVING - The New York Times
    Nov 19, 1984 · The Bell Labs mathematician, Dr. Narendra Karmarkar, has devised a radically new procedure that may speed the routine handling of such problems ...Missing: history | Show results with:history
  17. [17]
    Interview with Margaret Wright - Celebratio Mathematica
    Then in 1984, Narendra Karmarkar of Bell Labs announced a polynomial-time algorithm for linear programming that was supposedly 50 times faster than the ...
  18. [18]
    On the Performance of Karmarkar's Algorithm - jstor
    A new polynomial-time algorithm for linear programming was announced by Narendra Karmarkar of. Bell Laboratories in 1984. This algorithm is claimed by Bell ...
  19. [19]
    [PDF] Linear Programming: Interior-Point Methods - cs.wisc.edu
    Karmarkar's projective algorithm [10], announced in 1984, also has the poly- nomial complexity property , but it came with the added inducement of good.
  20. [20]
    [PDF] Interior-point methods for optimization
    Almost twenty-five years ago, Karmarkar (1984) proposed his projective method to solve linear programming problems: from a theoretical point of view, this was a ...
  21. [21]
    [PDF] Potential Reduction Algorithms - biz.uiowa.edu
    Potential reduction algorithms have a distinguished role in the area of in- terior point methods for mathematical programming. Karmarkar's [44] al-.
  22. [22]
    None
    ### Summary of Karmarkar’s Method from Appendix E
  23. [23]
    [PDF] Sec. 6.1 Basic Ideas of Karmarkar's Algorithm - NC State ISE
    The algorithm starts with an interior solution in the original space, maps the solution to the center of A by a projective transformation, applies Equation ( ...Missing: core principles
  24. [24]
    [PDF] Karmarkar's Algorithm [Ka] - cs.Princeton
    We are now ready to prove the main result that the point b produced in. Karmarkar's algorithm also provides a good decrease in the potential function.
  25. [25]
    Primal-Dual Interior-Point Methods | 4. Potential-Reduction Methods
    ... algorithms. Karmarkar's algorithm [57], when adapted for the standard-form linear program (1.1), makes use of a logarithmic potential function of the form. ρ ...
  26. [26]
    [PDF] Interior-Point Polynomial Algorithms in Convex Programming
    our main notions of self-concordant functions and barriers. Chapter 3 is devoted ... Karmarkar's algorithm for lin- ear programming,Report No. ORC 86-8 ...
  27. [27]
    (PDF) Karmarkar's Approach for Solving Linear Programming ...
    May 19, 2017 · Karmarkar's projective transformation algorithm for solving a linear program (LP) is the first polynomial time algorithm which was published in ...
  28. [28]
    [PDF] A Variant of Karmarkar's Linear Programming Algorithm for ...
    Apr 18, 1988 · The linear programs are assumed to be in standard form. The variant we develop is a primal projective algorithm; we show that it takes O(ng) ...<|separator|>
  29. [29]
    The iterative step in the linear programming algorithm of N. Karmarkar
    Abstract. We simplify and strengthen the analysis of the improvement obtained in one step of Karmarkar's algorithm.Missing: pseudocode | Show results with:pseudocode
  30. [30]
    Karmarkar's Algorithm - agill.xyz
    In this post I intend to explain what a Linear Program (LP) is, and how to solve an LP problem using Karmarkar's Algorithm implemented in Python.
  31. [31]
    Convergence in Karmarkar's Algorithm for Linear Programming
    Karmarkar's algorithm is formulated so as to avoid the possibility of failure because of unbounded solutions. A general inequality gives an easy proof of ...Missing: pseudocode | Show results with:pseudocode
  32. [32]
    An O(n 3 L) potential reduction algorithm for linear programming
    Feb 13, 1989 · Veiga, “An implementation of Karmarkar's algorithm for linear ... Sun, “An algorithm for convex quadratic programming that requires O(n 3.5 L) ...
  33. [33]
    [PDF] The AT&T Korbx System - Robert Vanderbei
    This paper describes the AT&T KORBX system, which implements certain variants ofthe Karmarkar algorithm on a computer that has multiple proces- sors, each ...<|separator|>
  34. [34]
    US4744028A - Methods and apparatus for efficient resource allocation
    Application filed by AMERICAN TELEPHONE AND TELEGRAPH COMPANY AT&T BELL LABORATORIES. 1985-04-19. Priority to US06/725,342. 1988-05-10. Publication of ...
  35. [35]
    [PDF] ON ABSTRACTION AND EQUIVALENCE - Andrew Chin
    Karmarkar filed a United States patent application on April 19, 1985 titled ... 225–26. 154. See U.S. Patent No. 4,744,028 cols. 7–8 (filed Apr. 19, 1985) ...
  36. [36]
    Karmarkar patent - software patents wiki (ESP Wiki)
    Jul 20, 2010 · According to Wikipedia, this patent expired in April 2006 and is presently in the public domain. See Karmarkar's algorithm patent controversy.
  37. [37]
    IDEAS & TRENDS; Mathematicians Are Troubled by Claims on Their ...
    Mar 12, 1989 · The scheduling algorithm, discovered by Narendra K. Karmarkar, a mathematician for A.T. & T., is a revolutionary way to solve what are known as ...Missing: response | Show results with:response
  38. [38]
    BUSINESS TECHNOLOGY; Patents on Equations: Some See a ...
    Feb 15, 1989 · In the past, the courts have ruled that mathematical equations could not be patented because they were similar to the laws of nature and to ...Missing: Karmarkar | Show results with:Karmarkar
  39. [39]
    Did Karmarkar deserve a patent? - CMU School of Computer Science
    Peter Treloar suggested the linear programming patent developed at AT&T > Bell Labs by Karmarkar. ... Karmarkar and AT&T for their implementation in 1984 were ...Missing: press | Show results with:press
  40. [40]
    Linear Programming Problem - an overview | ScienceDirect Topics
    The algorithm was developed at AT&T to solve very large linear programming problems concerned with routing telephone calls in the Pacific Basin. This algorithm ...<|control11|><|separator|>
  41. [41]
    Flight Crew Scheduling - jstor
    widely-known crew scheduling programs. Korbx is. AT&T's system based on the Karmarkar algorithm. As indicated earlier Korbx, is used by Delta Airlines. TPACS.
  42. [42]
    [PDF] Computational Progress in Linear and Mixed Integer Programming
    Separation of logical and physical allocation of data. ERP systems introduced. ◦ Karmarkar's 1984 paper on interior-point methods. 6. © 2017 Gurobi Optimization ...
  43. [43]
    [PDF] Self-Scaled Barriers and Interior-Point Methods for Convex ...
    Sep 3, 1996 · Abstract: This paper provides a theoretical foundation for efficient interior-point algorithms for convex programming problems expressed in ...
  44. [44]
    [PDF] INTERIOR POINT POLYNOMIAL TIME METHODS IN CONVEX ...
    The point, anyhow, is that actual behaviour of the method of. Karmarkar ... rate of convergence always becomes worse and worse as these parameters become worse.
  45. [45]
    [PDF] Semidefinite Programming - Stanford University
    Karmarkar's algorithm and the interior-point methods developed afterwards ... The central path passes through the analytic center of the constraint. F(x) ...
  46. [46]
    [PDF] Interior Point Methods for Semidefinite Programming
    May 20, 2025 · Newton barrier methods for linear programming and an equivalence to Karmarkar's projective method. Mathematical Programming, 36:183–209, 1986.
  47. [47]
    [PDF] Interior-Point Methods for Nonconvex Nonlinear Programming ...
    ABSTRACT. The paper extends prior work by the authors on LOQO, an interior point al- gorithm for nonconvex nonlinear programming.
  48. [48]
    [PDF] Complexity Analysis of Interior Point Algorithms for Non-Lipschitz ...
    Abstract We propose a first order interior point algorithm for a class of non-. Lipschitz and nonconvex minimization problems with box constraints, which.Missing: limitations | Show results with:limitations
  49. [49]
    Interior-point methods - ScienceDirect.com
    For an appropriately chosen starting point (y0,z0) with μ0=(y0)Tz0/n, we obtain convergence to a point with μ⩽ε in O n τ log μ 0 ε iterations , where τ= 1 2 , 1 ...
  50. [50]
    Twenty-Five Years of Interior Point Methods | Tutorials in OR
    Oct 14, 2014 · Twenty-five years after the publication of Karmarkar's path-breaking paper, this tutorial attempts to give a glimpse of the seminal results ...
  51. [51]
    Interior point methods, a decade after Karmarkar—a survey, with ...
    The introduction of Karmarkar's polynomial algorithm for linear programming (LP) in 1984 has influenced wide areas in the field of optimization.
  52. [52]
    A Path-Following Projective Interior Point Method for Linear ...
    In this paper the authors present a projective interior point method which follows the “central trajectory” and finds an optimal solution in at most O ( n L ) ...
  53. [53]
    On the Complexity of a Class of Projective Interior Point Methods - jstor
    We show that this class of algorithm has a worst case iteration bound of O(-nL) and is closely related to the class of potentia reduction interior point methods ...
  54. [54]
    Interior point methods in the year 2025 - ScienceDirect
    In this survey we will discuss several issues related to interior point methods.
  55. [55]
    Warm-Start Strategies in Interior-Point Methods for Linear ... - SIAM.org
    We find that the resulting best combined warmstarting strategy manages to save between 50 and 60% of interior point iterations, consistently outperforming ...Missing: parallelization | Show results with:parallelization
  56. [56]
    [PDF] Interior Point Methods 25 Years Later∗ - School of Mathematics
    Oct 14, 2010 · Below we briefly review a recent development of Gondzio [43] to redesign interior point methods to allow the solution of very large LPs and ...
  57. [57]
    [PDF] Advanced Gurobi Algorithms
    • Can be warm-started. • Effectively handles problem modifications. • Barrier warm-start still an open research topic. • Primal or Dual degeneracy can be ...
  58. [58]
    None
    Summary of each segment:
  59. [59]
    [PDF] Faster Randomized Interior Point Methods for Tall/Wide Linear ...
    Path-following IPMs (also called central-path algorithms) and, in particular, long-step path following IPMs, are among the most practical approaches for solving ...
  60. [60]
    Stochastic interior-point methods for smooth conic optimization with ...
    Dec 17, 2024 · We introduce a stochastic interior-point method (SIPM) framework for general conic optimization, along with four novel SIPM variants leveraging distinct ...