Fact-checked by Grok 2 weeks ago

Interior-point method

The interior-point method is a class of algorithms used to solve problems, particularly , by generating a sequence of iterates that remain strictly inside the and approach the optimal solution along a "central path" defined by barrier functions that penalize proximity to constraint boundaries. These methods employ self-concordant barrier functions, often logarithmic, combined with to compute search directions, ensuring polynomial-time convergence under appropriate conditions. The origins of interior-point methods trace back to the 1950s and 1960s with early barrier function approaches, such as Frisch's logarithmic barrier method for linear inequalities in 1955 and the more systematic development by Fiacco and McCormick in their 1968 book on nonlinear programming. However, these techniques saw limited practical adoption until 1984, when Narendra Karmarkar introduced a groundbreaking polynomial-time algorithm for linear programming at AT&T Bell Laboratories, which demonstrated up to 50 times faster performance than the simplex method on certain large-scale problems and sparked the "interior-point revolution." This projective scaling method was soon shown to be equivalent to classical barrier methods, leading to rapid theoretical advancements, including proofs of polynomial complexity by researchers like Renegar, Gonzaga, and Roos by 1988. Key theoretical foundations were solidified in the late 1980s and early 1990s through the work of and Nemirovski, who introduced the concept of self-concordant functions to guarantee efficient convergence via damped steps, achieving an iteration complexity of O(\sqrt{\nu} \log(1/\epsilon)), where \nu is the barrier parameter and \epsilon is the desired accuracy. Primal-dual variants emerged as particularly effective, solving both primal and dual problems simultaneously to exploit complementarity conditions, and were extended beyond to , conic, semidefinite, and nonlinear optimization by the mid-1990s. Influential contributors include Michael Todd, Yinyu Ye, and Tamás Terlaky, who refined algorithms for practical implementation. In terms of advantages, interior-point methods offer worst-case polynomial-time performance—contrasting with the simplex method's worst-case complexity—making them suitable for very large instances, though they can be computationally intensive due to solving large linear systems at each . They have profoundly impacted and , powering commercial solvers like CPLEX and Gurobi, and enabling applications in machine learning, control theory, , and approximations of NP-hard problems via relaxations. Over the four decades since Karmarkar's breakthrough, these methods have become the dominant paradigm for , with ongoing research addressing scalability and extensions to non-convex settings.

Overview

Basic Concept

Interior-point methods (IPMs) are a class of algorithms designed to solve problems, including linear and nonlinear programs, by generating a sequence of iterates that remain strictly inside the throughout the optimization process. This interior approach ensures that constraints are satisfied with positive slack, preventing direct contact with the boundaries where feasibility might be lost due to numerical issues. The fundamental mechanism in IPMs is the use of a , which modifies the objective to impose a penalty that increases rapidly as the iterate nears the boundary of the feasible set. A example is the logarithmic barrier for constraints f_i(x) \leq 0, i=1,\dots,m, given by \phi(x) = -\sum_{i=1}^m \log(-f_i(x)), defined on the strict interior where f_i(x) < 0 for all i, and tending to +\infty as any f_i(x) approaches 0 from below. More generally, self-concordant barrier functions, which satisfy specific convexity and higher-order smoothness conditions, provide a theoretical framework for analyzing convergence across broader classes of convex sets. In contrast to boundary methods like the , which move along the edges or vertices of the feasible polytope and can stall due to degeneracy—where multiple constraints bind at a point without progress—IPMs mitigate such issues by maintaining distance from boundaries, leading to more robust performance on ill-conditioned problems. At a high level, the algorithm begins with a strictly feasible initial point x^{(0)} in the interior of the feasible set. It then iteratively solves a sequence of barrier subproblems for decreasing positive parameters \mu_k \to 0: minimize the original objective f(x) plus the barrier term \mu_k \phi(x), subject to any equality constraints. Each solution x^{(\nu_k)} approximates the optimum better as \mu_k shrinks, with the path of these solutions forming the that converges to the optimal set. For illustration, consider the canonical linear program \min c^\top x subject to Ax = b, x \geq 0. The associated barrier subproblem is \min_{x > 0} \, c^\top x - \mu \sum_{j=1}^n \log x_j \quad \text{subject to} \quad Ax = b, where the logarithmic term penalizes small x_j values, enforcing positivity; as \mu \to 0, the minimizer approaches a optimum of the original problem.

Role in Optimization

Interior-point methods (IPMs) represent a significant advancement in optimization algorithms, particularly when compared to the method, which has long been the standard for . While the method traverses the vertices of the feasible and is often efficient in practice for small to medium-sized problems, it lacks theoretical polynomial-time guarantees and can exhibit due to or degeneracy. In contrast, IPMs approach the optimum from the interior of the , achieving polynomial-time complexity in theory, which makes them more reliable for proving the existence of efficient algorithms. However, for small instances, may require fewer iterations and is simpler to implement, whereas IPMs can demand more computational effort per iteration, especially in dense problems. The advantages of IPMs stem from their self-correcting nature, where deviations from the central path—a trajectory of points that converge to the optimum—are actively damped through mechanisms like predictor-corrector steps, preventing the stagnation issues common in vertex-based methods. This property, combined with their scalability to high-dimensional problems, allows IPMs to handle large-scale, sparse instances with hundreds of thousands of variables and constraints more effectively than , as they exploit structure through matrix factorizations that parallelize well on modern hardware. Furthermore, IPMs extend beyond to a broad class of problems, including quadratic, semidefinite, and conic programs, where the feasible region is defined by inequalities rather than equalities alone. Theoretically, IPMs offer polynomial complexity guarantees when using self-concordant barrier functions, which enforce interior feasibility while controlling the condition number of the problem; for instance, path-following variants require O(\sqrt{\vartheta} \log(1/\epsilon)) iterations to achieve an \epsilon-optimal solution, where \vartheta is the barrier parameter measuring the "complexity" of the feasible set. This ensures reliable performance across problem classes, such as linear programs with \vartheta = m (number of inequalities). In practice, IPMs dominate modern commercial solvers like Gurobi, which employs barrier methods as the primary engine for continuous optimization, and CVX (via interfaces to SDPT3 or SeDuMi), where they efficiently solve linear and semidefinite programs at scale, often outperforming alternatives for problems exceeding 10,000 variables.

Historical Development

Origins and Early Ideas

The origins of interior-point methods trace back to early efforts in during the mid-20th century, particularly through the development of barrier functions to handle inequalities without directly encountering boundaries. In 1955, introduced the logarithmic potential method, an early barrier approach for solving systems of linear inequalities by incorporating a logarithmic barrier term that penalizes proximity to the boundaries of the , thereby maintaining strict feasibility throughout the optimization process. This method laid foundational ideas for transforming constrained problems into unconstrained ones via barriers, focusing initially on contexts. Building on these ideas, the 1960s saw the emergence of penalty and barrier methods in , with significant advancements by Anthony V. Fiacco and Garth P. McCormick through their sequential unconstrained minimization technique (SUMT). Introduced in their 1968 book, SUMT employed logarithmic barrier functions to sequentially minimize a series of unconstrained problems, where a barrier parameter is gradually reduced to approach the original constrained optimum while ensuring iterates remain in the interior of the feasible set. This approach extended concepts to nonlinear constraints, emphasizing the use of barriers to navigate the interior and avoid the numerical instabilities associated with points in active-set methods. Key concepts from these early works included the notion of the analytic center, proposed by Pierre Huard in as a central point within the that maximizes the product of distances to the boundaries, serving as a robust starting point for iterative procedures. Barrier methods also introduced the idea of protecting against infeasibility by enforcing interior points, which prevented divergence in cases where the original problem might lack feasible solutions or exhibit ill-conditioning near boundaries. These innovations drew from broader influences in , where interior trajectories were explored to circumvent the challenges of Lagrange multipliers and direct constraint handling. Despite these contributions, early barrier methods suffered from notable limitations, including the absence of polynomial-time guarantees, which made their practical unpredictable for large-scale problems compared to emerging alternatives like the simplex method. Additionally, they were highly sensitive to the choice of barrier parameters, requiring careful tuning that often led to slow or numerical instability if not managed precisely. These shortcomings contributed to a decline in their use during the , as active-set and penalty methods gained prominence for their perceived reliability in nonlinear contexts.

Karmarkar's Breakthrough

In 1984, introduced a groundbreaking polynomial-time algorithm for solving problems, which relied on interior-point methods to navigate the from within rather than along its boundaries. The algorithm first transforms the linear program into a standard form: minimize c^T y subject to A y = 0, e^T y = 1, and y > 0, where e is the vector of ones. It employs a projective transformation T(y) = D^{-1} y / (e^T D^{-1} y), with D = \diag(y), to rescale the space and center the current iterate at the simplex center, preserving the problem's structure under this metric. Iterative steps then compute a direction of steepest descent for the projected gradient and update the point by reducing a , ensuring progress toward the optimum while staying in the interior. The key innovation of Karmarkar's projective was its demonstration of polynomial-time complexity for , achieving O(n^{3.5} L) arithmetic operations in the worst case, where n is the number of variables and L is the of the input. This bound arose from the potential function's consistent reduction by a fixed amount per , combined with efficient handling of the projective rescalings. At its mathematical core, the algorithm uses a homogeneous projective formulation to handle feasibility issues. A central element is the potential function \phi(y) = n \log(c^T y) - \sum_{i=1}^n \log y_i, which balances objective progress and interiority in the projective space and is reduced iteratively to drive convergence. Karmarkar's work ignited the "interior-point revolution" in optimization, revitalizing barrier methods and inspiring thousands of subsequent research papers that extended and refined these techniques for broader applications.

Evolution Post-1984

Following Karmarkar's 1984 breakthrough, interior-point methods rapidly evolved through variants emphasizing affine scaling and path-following approaches, which improved theoretical guarantees and practical efficiency. In 1988, James Renegar introduced a polynomial-time based on that followed the central path of the logarithmic , providing a foundational framework for path-following methods and highlighting the role of barrier functions in ensuring . This work laid the groundwork for analyzing barrier properties, paving the way for subsequent theoretical advancements in handling convex constraints. The development of primal-dual methods marked a significant refinement in the late 1980s and early 1990s, shifting focus toward simultaneous optimization of and problems for enhanced . A pivotal contribution came in 1992 with Sanjay Mehrotra's predictor-corrector algorithm, which iteratively predicts a centering and corrects toward feasibility, achieving superlinear and becoming a for practical implementations due to its efficiency on large-scale linear programs. This method addressed early limitations in pure or approaches by balancing and progress, and it remains widely adopted in software for its robustness. Extensions to broader classes of optimization problems accelerated in the mid-1990s, particularly through theoretical tools for nonlinear and conic forms. In 1994, and Arkadii Nemirovski developed the concept of universal barrier functions and self-concordant barriers, enabling polynomial-time interior-point algorithms for general convex programming by associating any proper cone with a suitable barrier that satisfies self-concordance properties for . Their framework unified prior variants and extended applicability to conic quadratic and semidefinite programs, demonstrating that most convex problems admit such barriers with controlled complexity parameters. Key milestones in the 1990s included the adaptation of interior-point methods to (), where polynomial-time solvability was established for problems involving matrix constraints. Farid Alizadeh's 1995 primal-dual path-following algorithm for extended techniques using trace-based barriers, proving convergence in O(\sqrt{n} \log(1/\epsilon)) iterations for n \times n matrices and accuracy \epsilon, which facilitated applications in and . By the late 1990s and early 2000s, these methods were integrated into commercial solvers such as CPLEX and MOSEK, where interior-point implementations outperformed methods on dense, large-scale instances, driven by advances in sparse linear algebra. Ongoing challenges, such as warm-starting to reuse solutions from nearby problems and managing ill-conditioned instances, were systematically addressed during this period to enhance real-world applicability. Warm-starting techniques, refined in the late 1990s, involved perturbing barrier parameters and recovering central points from prior iterates, reducing iteration counts by up to 50% in sequential optimizations like , as shown in analyses of homogeneous self-dual embeddings. For ill-conditioned problems, regularization strategies—such as adding scaled slacks or modifying centrality measures—were developed to mitigate numerical instability near boundaries, ensuring reliable convergence even when condition numbers exceed 10^{10}, particularly in and nonlinear extensions.

Mathematical Foundations

Barrier Methods and Functions

Barrier methods form a cornerstone of interior-point techniques for , relying on to enforce feasibility by penalizing approaches to the boundary of the feasible set. A for a closed G \subseteq \mathbb{R}^n is a B: \operatorname{int} G \to \mathbb{R} that tends to +\infty as x approaches the boundary of G, ensuring that minimization occurs strictly within the interior while approximating the original constrained problem as the barrier's influence diminishes. This property allows iterates to remain feasible throughout the optimization process, avoiding explicit projections onto the boundary. For linear inequality constraints defining a polytope G = \{ x \in \mathbb{R}^n \mid A x \leq b \}, where A \in \mathbb{R}^{m \times n} and b \in \mathbb{R}^m, the standard logarithmic barrier function is given by B(x) = -\sum_{i=1}^m \log(b_i - a_i^T x), defined on the interior where A x < b. This function is strictly convex and diverges to +\infty as any slack b_i - a_i^T x approaches zero, effectively repelling solutions from inactive constraints. The logarithmic form arises naturally from the need for a smooth, differentiable penalty that scales appropriately with the distance to the boundary. A crucial property enabling efficient computation in interior-point methods is the self-concordance of the barrier function, which ensures that Newton steps remain affine-invariant and provide reliable damping for convergence analysis. A thrice-differentiable convex function B on an open convex set is self-concordant if it satisfies |D^3 B(x)[h, h, h]| \leq 2 (h^T \nabla^2 B(x) h)^{3/2} for all x \in \operatorname{int} G and h \in \mathbb{R}^n. For a \nu-self-concordant barrier, it additionally holds that \nabla B(x)^T [\nabla^2 B(x)]^{-1} \nabla B(x) \leq \nu for all x, where the left-hand side is the square of the . For the standard logarithmic barrier over m inequalities, \nu = m, which bounds the complexity of Newton iterations and guarantees polynomial-time solvability independent of conditioning. This property underpins the damped 's quadratic convergence locally and global linear reduction in the barrier parameter. To extend barrier methods beyond polytopes to arbitrary convex sets, universal barrier functions provide a general framework with controlled self-concordance parameters. For any n-dimensional closed convex domain G, there exists a universal \mathcal{O}(n)-self-concordant barrier B(x) = -\log \int_{G^*(x)} e^{-(s, x)} \, ds, where G^*(x) is an affine image of the domain, ensuring applicability to diverse problems like semidefinite programming. The self-concordance parameter \nu for such barriers satisfies \nu = \mathcal{O}(n), leading to iteration complexity bounds of \mathcal{O}(\sqrt{\nu} \log(1/\epsilon)) Newton steps to achieve \epsilon-accuracy in the objective, where the logarithm accounts for the initial distance to optimality. These bounds hold for path-following schemes and highlight the scalability with problem dimension. The core optimization subproblem in barrier methods involves solving a sequence of unconstrained minimizations of the form \min_{x \in \operatorname{int} G} f(x) + \mu B(x), where f is the original convex objective and \mu > 0 is a decreased geometrically to zero (e.g., \mu_{k+1} = \beta \mu_k with $0 < \beta < 1). As \mu \to 0, the minimizers x(\mu) approach the optimum of the constrained problem, with the duality gap scaling as \mathcal{O}(\mu \nu). Each subproblem is typically solved approximately using self-concordant Newton iterations until proximity to the central path is achieved.

Central Path and Duality

The interior-point method for linear programming revolves around solving a primal-dual pair of optimization problems. The primal problem is formulated as minimizing c^T x subject to Ax = b and x \geq 0, while the dual problem maximizes b^T y subject to A^T y + s = c and s \geq 0, where A is an m \times n matrix, b \in \mathbb{R}^m, c \in \mathbb{R}^n, and the variables x, s \in \mathbb{R}^n, y \in \mathbb{R}^m. Under strict feasibility assumptions, strong duality holds, and the optimal value is attained when the duality gap c^T x - b^T y = x^T s = 0, satisfying complementary slackness x_i s_i = 0 for all i. To navigate the interior of the feasible region toward optimality, the central path is defined as the trajectory \{(x(\mu), y(\mu), s(\mu)) : \mu > 0\}, where for each barrier parameter \mu > 0, the triplet minimizes the primal-dual logarithmic barrier problem: minimize c^T x - b^T y + \mu \sum_{i=1}^n (\log x_i + \log s_i) subject to Ax = b and A^T y + s = c, with x > 0 and s > 0. As \mu \to 0^+, points on the central path converge to a primal-dual optimal pair, approaching the optimal face of the . The centrality condition characterizing points on this path is x_i s_i = \mu for all i = 1, \dots, n, or equivalently in vector form, Xs = \mu e, where e is the all-ones vector and X = \operatorname{diag}(x). This condition enforces a balanced distribution of the duality gap x^T s = n\mu, preventing iterates from drifting toward the boundary prematurely. In the limit as \mu \to 0, it recovers complementary slackness x_i s_i = 0, linking the path directly to primal-dual optimality. Geometrically, the central path traces a smooth curve through the interior of the and feasible regions, approximating the optimal face while damping oscillations toward the boundary; it can be viewed as a sequence of analytic centers weighted by the objective, providing a that avoids ill-conditioning near vertices. This path-following approach, first formalized in the context of weighted centers, ensures that perturbations remain controlled as the algorithm progresses. To maintain centrality, a damped Newton step is computed by solving the perturbed Karush-Kuhn-Tucker (KKT) system for increments \Delta x and \Delta s: \begin{align*} A \Delta x &= 0, \\ A^T \Delta y + \Delta s &= 0, \\ S \Delta x + X \Delta s &= \sigma \mu e - X s, \end{align*} where \sigma \in (0,1] is the centering parameter that targets a fraction \sigma of the current , and the full step is scaled by a to preserve positivity. This step recenters the iterate toward the path, with \sigma = 1 yielding pure centering and smaller \sigma allowing progress toward \mu = 0.

Feasibility and Optimality Conditions

In problems, the Karush-Kuhn-Tucker (KKT) conditions provide necessary and sufficient criteria for optimality under appropriate constraint qualifications, such as . These conditions consist of four components: stationarity, which requires that the of the equals zero, i.e., \nabla f_0(x) + \sum_{i=1}^m \lambda_i \nabla f_i(x) + A^T \nu = 0; primal feasibility, ensuring f_i(x) \leq 0 for i=1,\dots,m and Ax = b; dual feasibility, requiring \lambda \geq 0; and complementary slackness, stating \lambda_i f_i(x) = 0 for all i. From the perspective of interior-point methods, strict feasibility—where there exists a point x > 0 (or more generally, in the strict interior of the feasible set) and a dual slack s > 0 satisfying the equality constraints—plays a crucial role. This assumption, akin to , guarantees , the attainment of optimal primal and dual values, and the existence of a unique central path that approaches the optimal solution as the barrier parameter tends to zero. The , defined as the difference between primal and dual objective values c^T x - b^T y, equals \sum_i x_i s_i in standard formulations under the equality constraints Ax = b and A^T y + s = c. Interior-point methods initialize with a large duality gap and iteratively reduce it toward zero by following the central path, typically achieving a gap below a tolerance \epsilon > 0. To initiate the main phase of interior-point methods, a phase-I procedure is employed to compute an initial strictly feasible point, such as one satisfying Ax = b and x \geq \delta e for some \delta > 0 and vector of ones e, often via an auxiliary linear program or big-M formulation. A point is deemed \epsilon-optimal when the is less than \epsilon and the and residuals (measuring violations) are sufficiently small, ensuring the objective value is within \epsilon of the true optimum.

Core Algorithms

Path-Following Methods

Path-following methods form a class of interior-point algorithms that explicitly trace the central by iteratively solving perturbed optimality conditions with a decreasing barrier parameter μ, ensuring iterates remain in a neighborhood of the path while approaching the optimal solution. These methods originated with early analyses demonstrating polynomial-time convergence for , building on the theoretical foundation of self-concordant barrier functions. In the general framework, the algorithm begins with an initial strictly feasible point and a positive μ₀ larger than the initial . At each k, given a current point (x^k, s^k) in the interior, the method computes a search direction by applying to the barrier subproblem defined by the logarithmic with parameter μ_k. This involves solving for the Newton step Δx and Δs that approximately satisfy the perturbed complementarity condition X S e ≈ μ_k (σ e), where σ ∈ [0,1] controls the centering: σ=1 fully recenters on the central path, while σ<1 allows progress toward the optimum. The step is then taken as (x^{k+1}, s^{k+1}) = (x^k, s^k) + α (Δx, Δs), where the step length α is chosen to maintain strict feasibility and keep the new point within a defined neighborhood of the central path, such as the negative infinity norm neighborhood N_{-∞}(β) for β <1. Following the step, μ is reduced multiplicatively, typically by a factor θ = 1 - δ/√n with δ >0 fixed, ensuring the decreases steadily. This process continues until μ falls below a tolerance related to the desired accuracy ε. A prominent variant is the predictor-corrector method, which improves practical efficiency by separating the computation of the search direction into two phases. In the predictor step, σ=0 is used to compute an affine-scaling direction that estimates the path to the , maximizing in reducing the while respecting feasibility. This direction may deviate from the central , so a subsequent corrector step applies σ close to 1 (often computed adaptively, e.g., σ = (ν^{aff}/ν)^3 where ν^{aff} is the predicted duality measure) to recenter the iterate toward the central for the next μ. The combined step length is determined by the minimum of the predictor and corrector feasibility bounds, often allowing larger steps than pure path-following. This variant achieves high practical performance and maintains theoretical guarantees when step sizes are controlled appropriately. Affine scaling emerges as a special case of path-following when σ=0 throughout, yielding a pure direction for feasibility without centering toward μ e. In this setup, the search direction solves the system that linearizes the primal and feasibility conditions along with the unperturbed complementarity (ignoring μ), effectively the variables inversely to their current values to enlarge the toward the boundary. While conceptually simple and useful for initialization or pure feasibility problems, affine scaling alone often requires smaller steps and more iterations in practice due to potential large deviations from , making it less efficient for optimization compared to centered variants. In - implementations of path-following, the search directions satisfy the following system (assuming residuals r_p and r_c for and feasibility): \begin{cases} A \Delta x = -r_p, \\ A^T \Delta \lambda + \Delta s = -r_c, \\ S \Delta x + X \Delta s = \sigma \mu e - X S e, \end{cases} where A is the constraint matrix, X = \operatorname{diag}(x), S = \operatorname{diag}(s), e is the vector of ones, \mu is the current duality measure, and \sigma \in [0,1] is the centering parameter. Under strict feasibility (r_p = r_c = 0), this simplifies accordingly. This structure enables symmetric updates and efficient computation via factorization. For linear programs, short-step path-following methods exhibit worst-case iteration complexity of O(√n log(1/ε)), where n is the number of variables and ε is the desired relative accuracy, assuming a self-concordant barrier of complexity O(n). Long-step variants, including predictor-corrector, match this bound under adaptive step-size rules while allowing larger practical steps.

Potential-Reduction Methods

Potential-reduction methods constitute a fundamental class of interior-point algorithms that guarantee progress toward optimality by systematically decreasing a specially designed potential function at each iteration, distinguishing them from approaches that closely track the central path. Developed in the wake of Karmarkar's 1984 breakthrough, these methods emphasize theoretical robustness and have been extended to various convex optimization settings, including linear and some nonlinear programs. The core of these methods revolves around a primal-dual defined as \Phi(x, s) = (n + \rho) \log(x^T s) - \sum_{i=1}^n \log x_i - \sum_{i=1}^n \log s_i, where x > 0 and s > 0 represent the and variables, respectively, n is the , and the \rho > 0 (with effective q = n + \rho > n) ensures the function is coercive and bounded below while balancing reduction against barrier penalties. Choices such as \rho = \sqrt{n} provide the necessary margin for potential decrease guarantees. The algorithm proceeds iteratively from a strictly feasible starting point (x, s): at each step, a search direction is computed via Newton's method applied to the quadratic approximation of \Phi, subject to linearized primal and dual feasibility constraints, yielding affine-scaling-like directions \Delta x and \Delta s. A line search is then performed along this direction to select the largest step size \alpha_k that ensures a sufficient reduction in \Phi (typically at least a fixed fraction like 0.1 of the predicted decrease) while preserving strict positivity of x + \alpha_k \Delta x and s + \alpha_k \Delta s. This adaptive stepping maximizes progress in the potential without requiring proximity to the central path. Under standard assumptions, such as a bounded initial duality gap and appropriate parameter selection, each iteration achieves a reduction in \Phi of at least a positive constant (e.g., \delta = 1/(2(n + \sqrt{n}))), independent of the problem instance. Given that \Phi is initially finite and bounded below by a term scaling with n and the logarithm of the initial gap, this yields a total of O(n \log(1/\epsilon)) iterations to attain an \epsilon-optimal solution, where \epsilon measures the relative duality gap. This complexity bound holds for linear programming and extends to self-concordant barrier functions in broader convex settings. These methods trace their lineage to Karmarkar's original , which employed a projective potential \phi(d) = -\sum \log d_i in a homogeneous, self-dual of the linear , achieving potential reductions via projective transformations that inspired the general framework. Subsequent primal-dual variants generalized this by symmetrizing treatment of primal and dual variables, improving iteration complexity from Karmarkar's O(n^{3.5} L) (with L the input size) to the more efficient form above. Potential-reduction methods offer advantages in analytical simplicity, particularly for nonlinear problems where barrier parameters may vary, allowing straightforward proofs of convergence without explicit centrality conditions. In practice, however, they tend to be less efficient than path-following methods, as the line searches can be computationally demanding and step sizes more conservative, leading to fewer adoptions in modern solvers.

Primal-Dual Methods

Primal-dual interior-point methods represent a class of symmetric algorithms that simultaneously update the variables x, variables y, and slack variables s to follow the central path toward optimality in problems, particularly . These methods leverage the duality between and problems to maintain feasibility and complementarity throughout the iterations, offering improved and efficiency compared to purely or approaches. The foundational framework was introduced by Kojima, , and Yoshise, who demonstrated that joint updates preserve the interior-point property and converge to the optimal solution under mild assumptions. In these methods, the search directions \Delta x, \Delta y, and \Delta s are computed by solving an augmented derived from the Newton equations for the perturbed Karush-Kuhn-Tucker (KKT) conditions. Assuming strict feasibility for simplicity (with residuals incorporated in general implementations), the takes the form: \begin{cases} A \Delta x = 0, \\ A^T \Delta y + \Delta s = 0, \\ S \Delta x + X \Delta s = \sigma \mu e - X S e, \end{cases} where A is the matrix, X = \operatorname{diag}(x), S = \operatorname{diag}(s), e is the vector of ones, \mu is the current duality measure, and \sigma \in [0,1] is the centering that controls the proximity to the central . This can be reformulated as a symmetric indefinite of size m + n + n, where m and n are the dimensions of the and variables, respectively, and solved efficiently using specialized factorizations. The resulting directions ensure that the updated iterates remain in the interior of the while reducing the . A prominent variant is Mehrotra's predictor-corrector method, which enhances efficiency through adaptive stepping without explicit line searches. The algorithm first computes a predictor direction by setting \sigma = 0, yielding an affine-scaling step that estimates the direction toward the optimal solution while potentially violating centrality. A corrector direction is then obtained using a centering parameter \sigma computed from the predicted duality gap \mu^{\text{aff}} as \sigma = \left( \frac{\mu^{\text{aff}}}{\mu} \right)^3, where \mu is the current centrality measure; this step recenters the iterate toward the path. The final step combines these directions with a heuristic line search to maximize the step length while maintaining a safe proximity to the central path, typically ensuring a reduction in the duality gap by a fixed factor. This approach, introduced by Mehrotra, significantly reduces the number of iterations in practice by allowing larger adaptive steps. Primal-dual methods distinguish between short-step and long-step strategies for damping the directions. Short-step methods employ a fixed step size \alpha = \Theta(1/\sqrt{n}), guaranteeing polynomial-time by staying within a narrow neighborhood of the central , but requiring up to O(\sqrt{n} \log(1/\epsilon)) iterations for \epsilon-accuracy, where n is the problem . Long-step variants, such as Mehrotra's, use adaptive mechanisms like line searches or ratio tests to select larger \alpha, potentially achieving faster practical performance at the cost of more complex proximity control. These methods dominate modern solvers due to fewer required factorizations per iteration and robust handling of ill-conditioned problems. The iteration complexity of primal-dual path-following methods is O(\sqrt{n} \log(1/\epsilon)), matching the best theoretical bounds for interior-point algorithms and enabling efficient solution of large-scale problems. Detailed analyses, including large-update and predictor-corrector variants, are covered in subsequent sections on aspects.

Implementation Aspects

Initialization and Phase-I Techniques

Interior-point methods require an initial strictly feasible point in the interior of the to initiate the main optimization phase, as the algorithms rely on barrier functions that become on the boundary. One standard approach to obtain such a point is the phase-I problem, which introduces artificial slack variables to measure and minimize infeasibilities. For a linear program in standard form \min \{ c^T x \mid Ax = b, x \geq 0 \}, the phase-I formulation minimizes the auxiliary objective e^T r^+ + e^T r^-, where r = Ax - b = r^+ - r^- with r^+, r^- \geq 0, subject to Ax + r^+ - r^- = b, x, r^+, r^- \geq 0. This drives the residual r to zero while ensuring nonnegativity, yielding a feasible starting point if the minimum is zero; otherwise, the problem is detected as infeasible. To address both feasibility and optimality in a unified , the homogeneous self-dual transforms the original problem into a single homogeneous system where the origin is the unique optimal solution if the problem is feasible and bounded, or it reveals infeasibility otherwise. This constructs an extended linear program with additional variables and constraints that incorporate the , , and a homogenizing component, allowing the interior-point to start from a known interior point like the all-ones scaled appropriately. The ensures polynomial-time without separate phase-I, as introduced in the seminal work establishing O(\sqrt{n} L)- . For a robust starting point, the analytic center of the phase-I can be computed by solving a barrier subproblem, minimizing -\sum \log s_i - \sum \log x_j subject to the relaxed constraints with slacks s \geq \delta e and x \geq \delta e, where \delta > 0 is a small and e is the all-ones . This yields an initial point sufficiently distant from the boundary, facilitating damped steps in the main algorithm. In practice, initialization often employs perturbations, such as adding a small \epsilon > 0 to lower bounds (x \geq \epsilon e) or regularizing the Hessian with a multiple of the , to ensure strict feasibility and , particularly for ill-conditioned problems. For problems with bounded variables, big-M methods modify the objective with a large penalty M on artificial variables exceeding bounds, aiding without full reformulation. These techniques are commonly implemented in solvers to handle real-world data inconsistencies. The phase-I process incurs additional iteration complexity of O(\sqrt{n} \log(1/\delta)) Newton steps, where n is the problem , matching the order of the main and preserving overall polynomial time.

Convergence Analysis

The convergence analysis of interior-point methods relies on the properties of self-concordant barrier functions, which enable polynomial-time guarantees for achieving ε-optimality in problems. For a self-concordant barrier with ν (the complexity , equal to the n for with n variables), the number of Newton iterations required to reduce the error to ε is bounded by O(√ν log(ν/ε)). This bound arises from path-following or potential-reduction schemes that systematically decrease a measure of suboptimality, such as the or a barrier potential, while maintaining interior feasibility. Proximity to the central path is quantified using measures that assess how closely iterates align with central points, ensuring local efficiency of steps. A common proximity measure is the to , defined as ||w - w^c|| where w = (x, s) represents primal-dual iterates and w^c denotes the central path point at the current barrier parameter; this controls the in updates to guarantee progress. When this proximity is sufficiently small (e.g., below a threshold like 0.2), steps exhibit locally near the path, as the self-concordance property bounds the third of the barrier, leading to second-order model accuracy. In primal-dual methods, each iteration reduces the —the difference between primal and dual objectives—by a controlled factor, facilitating global . Specifically, the normalized duality gap μ (scaled by ν) is multiplied by (1 - δ/√ν) per step, where δ is a positive constant (typically around 0.1–0.3 depending on the neighborhood); this ensures the gap decreases geometrically until reaching ε. The bound derives from centering steps that keep iterates within a wide or narrow neighborhood of the central path, with the reduction factor balancing progress and feasibility preservation. Theoretical worst-case bounds, such as O(√ν log(ν/ε)) iterations, often overestimate the number required in practice due to adaptive step-sizing and the use of large-step strategies that exploit problem beyond the conservative assumptions of the . Empirical observations on linear and semidefinite programs show iteration counts scaling more like O(log(1/ε)) or far fewer than √ν, attributed to superlinear local behavior and efficient preconditioning. Extensions to predictor-corrector variants achieve superlinear convergence under good initialization, where the predictor step explores the path direction and the corrector recenters, yielding q-superlinear rates (error reduced faster than linear) when proximity to the optimum is O(1/√ν). This enhancement, analyzed in frameworks for conic programs, improves practical efficiency without altering the global polynomial bound.

Practical Considerations and Software

One major numerical challenge in interior-point methods arises from the ill-conditioning of the Karush-Kuhn-Tucker (KKT) as iterates approach the optimal solution, where components of the scaling become large, leading to potential breakdown in direct methods like . To mitigate this, implementations often employ for symmetric indefinite systems or iterative solvers such as preconditioned conjugate gradients (PCG), which are particularly effective for large-scale sparse problems by avoiding explicit formation. Heuristics play a crucial role in enhancing the robustness and of interior-point methods, with Mehrotra's predictor-corrector being a widely adopted approach that uses an affine-scaling predictor step to estimate centering directions and an adaptive corrector step based on the predicted to dynamically adjust the barrier , often reducing the need for line searches. Warm-starting techniques further improve by initializing subsequent solves with approximate solutions from prior related problems, such as perturbed instances or sequential optimizations, thereby reducing iteration counts; for example, strategies that adjust dual variables and barrier parameters based on the previous central path can achieve near-quadratic in practice for linear programs. For large sparse problems, and preconditioning are essential to address ill-conditioning and improve solver , including equilibration to balance row norms in the and block-diagonal preconditioners that exploit the of the KKT , such as separating and blocks to accelerate iterative methods like GMRES. These techniques ensure that the remains manageable, enabling efficient handling of problems with thousands of variables and constraints. Prominent software implementations include , an open-source solver for large-scale nonlinear optimization that employs a primal-dual interior-point framework with filter-line search and supports sparse linear algebra via interfaces to HSL or for factorization. MOSEK provides a high-performance interior-point optimizer for conic problems, utilizing a homogeneous self-dual that handles linear, quadratic, and semidefinite programs with built-in scaling and parametric optimization features. For , commercial solvers like Gurobi and CPLEX incorporate barrier (interior-point) methods as options, often with automatic selection based on problem characteristics, achieving superior speed on dense or network-structured instances compared to alternatives. As of 2025, open-source options like HiGHS, featuring the IPX solver, have gained prominence for large-scale with efficient iterative methods and preconditioning. Emerging GPU-accelerated solvers, such as MadIPM, enable handling of very large-scale problems by leveraging for Newton system solves. Performance can be further optimized through crossover routines that transition from an interior-point solution to a solution via pivoting, providing an exact for post-processing or . Degeneracy, which can stall progress by causing multiple optimal bases, is commonly handled with small perturbations to the right-hand side or coefficients, restoring strict feasibility without altering the optimal value significantly.

Applications to Convex Programs

Linear Programming

The interior-point method (IPM) for linear programming (LP) addresses the standard primal problem of minimizing \mathbf{c}^T \mathbf{x} subject to A \mathbf{x} = \mathbf{b} and \mathbf{x} \geq \mathbf{0}, where A is an m \times n matrix, \mathbf{b} \in \mathbb{R}^m, \mathbf{c} \in \mathbb{R}^n, and n denotes the number of variables. The corresponding dual is maximizing \mathbf{b}^T \boldsymbol{\lambda} subject to A^T \boldsymbol{\lambda} + \mathbf{s} = \mathbf{c} and \mathbf{s} \geq \mathbf{0}. To handle the non-negativity constraints, IPMs incorporate a logarithmic barrier function B(\mathbf{x}) = -\sum_{i=1}^n \log x_i, which penalizes iterates approaching the boundary of the feasible region while keeping them strictly interior. In the primal-dual framework, IPMs follow the central path defined by the perturbed complementarity condition x_i s_i = \mu for all i, where \mu > 0 is the barrier parameter, and the duality gap is approximately n \mu. For the standard LP barrier, the parameter \nu = n, reflecting the self-concordance parameter of the logarithmic barrier. The algorithm generates iterates by solving Newton systems on the primal-dual optimality conditions, typically using a predictor-corrector scheme to approximate the central path and reduce \mu toward zero. This approach ensures polynomial-time convergence, requiring O(\sqrt{n} L) arithmetic operations, where L is the total bit length of the input data. The core linear algebra in each iteration involves solving the normal equations derived from the step, often in the form A X^2 A^T \Delta \boldsymbol{\lambda} = \mathbf{r}, where X = \operatorname{diag}(\mathbf{x}), and \mathbf{r} incorporates residuals from the predictor direction; variants use a scaled A D^2 A^T with D = S^{-1/2} X^{1/2} to symmetrize the system. These systems are efficiently solved via sparse Cholesky factorization when A is sparse, enabling . In practice, IPMs dominate for large-scale LPs, solving problems with millions of variables in seconds on modern hardware by exploiting sparsity in the constraint matrix and using preconditioned iterative solvers for equations. For instance, in —formulated as minimizing risk subject to return constraints and budget limits—IPMs efficiently handle sparse networks with thousands of assets, yielding optimal allocations rapidly.

Semidefinite Programming

Semidefinite programming (SDP) extends to optimization over the cone of matrices, where interior-point methods adapt the self-concordant barrier approach to handle matrix variables. The standard primal SDP form is to minimize the trace inner product \mathbf{C} \bullet \mathbf{X} subject to linear equality constraints \mathbf{A}_i \bullet \mathbf{X} = b_i for i = 1, \dots, m, and \mathbf{X} \succeq 0, with \mathbf{X} an n \times n and \bullet denoting the \operatorname{trace}(\mathbf{X}^\top \mathbf{Y}). The associated self-concordant for the cone is -\log \det \mathbf{X}, defined on the interior where \mathbf{X} \succ 0; this barrier has self-concordance parameter \nu = n. In primal-dual interior-point methods for , the central path is parameterized by the barrier parameter \mu > 0, satisfying the centrality condition \mathbf{X} \mathbf{S} = \mu \mathbf{I}, where \mathbf{S} \succeq 0 is the from the \max \mathbf{b}^\top \mathbf{y} subject to \sum y_i \mathbf{A}_i + \mathbf{S} = \mathbf{C}, \mathbf{S} \succeq 0. Algorithms follow this path by solving systems to compute search directions that reduce the while maintaining approximate centrality, typically using predictor-corrector or Mehrotra-style steps. The theoretical iteration complexity is O(\sqrt{n} \log(1/[\epsilon](/page/Epsilon))) to achieve an \epsilon-optimal , though the per-iteration cost is high, reaching O(n^6) naively due to the need to form and factorize the m \times m involving n^3-cost operations. Applications of interior-point methods for are prominent in , where they solve linear matrix inequalities (LMIs) for robust controller design, such as stabilizing uncertain systems via Lyapunov functions. In , SDP relaxations solved by these methods provide tight approximations for NP-hard problems; for instance, the Goemans-Williamson uses an SDP relaxation for the max-cut problem, achieving a 0.878-approximation by the solution. These methods enable polynomial-time solvability for large-scale instances in practice, despite the theoretical costs.

Quadratic and Geometric Programs

Interior-point methods (IPMs) extend naturally to quadratically constrained programs (QCQPs), which take the form \min_x x^T Q x + q^T x subject to quadratic inequalities x^T A_i x + a_i^T x + b_i \leq 0 for i=1,\dots,m, where Q \succeq 0 and A_i \succeq 0 ensure . For QCQPs, these problems can be reformulated as semidefinite programs (SDPs) when the quadratic forms are low-rank, by lifting the variables via the substitution X = xx^T, yielding linear objectives and constraints over the semidefinite cone. This SDP reformulation enables the application of established IPM frameworks for SDPs, such as primal-dual path-following methods that achieve polynomial-time convergence. The barrier function for QCQPs combines the quadratic objective with logarithmic barriers on the constraints, forming a self-concordant barrier like -\sum_i \log(-x^T A_i x - a_i^T x - b_i), which guides Newton steps along the central path toward optimality. IPM adaptations for QCQPs involve solving the resulting KKT systems via Newton iterations, often exploiting sparsity in the Hessian for efficiency in large-scale instances; for the convex case, this yields complexity bounds similar to those of SDPs, typically O(\sqrt{n} \log(1/\epsilon)) iterations for n-dimensional problems to reach \epsilon-accuracy. Geometric programs (GPs), characterized by posynomial objectives and constraints (sums of monomials c_k \prod_j x_j^{a_{kj}} with c_k > 0), are nonconvex in standard form but can be transformed into problems via the y_j = \log x_j, converting the problem to \min_y \log \sum_k \exp(f_k(y)) subject to similar exponential inequalities, where f_k(y) are affine. IPMs for these reformulated GPs employ barriers tailored to the cone or use self-concordant barriers on the log-sum-exp terms, enabling efficient Newton-based solves with the same polynomial complexity as linear programs in the transformed space. These methods find applications in , where QCQPs model uncertainty sets for worst-case performance guarantees in control systems, and in , where GPs optimize parameters like widths under posynomial delay and power constraints. In both domains, IPMs provide robust, high-precision solutions, outperforming active-set methods for medium-to-large instances due to their global convergence properties.

Emerging Applications

Interior-point methods (IPMs) have found emerging applications in approximating \ell_p-norm minimization problems, particularly for p=1 and p=\infty, where the objective is to solve \min \|Ax - b\|_p. For p=1, this reduces to a linear program (LP) that IPMs solve efficiently using logarithmic barrier functions on the non-negativity constraints of the reformulated variables, enabling exact recovery in sparse approximation contexts. Similarly, for p=\infty, the problem is formulated as an LP minimizing \epsilon subject to -\epsilon \mathbf{1} \leq Ax - b \leq \epsilon \mathbf{1}, with IPMs applying central path-following techniques to handle the constraints robustly. In cases requiring smoothness, smoothed barrier functions approximate the non-differentiable norms, allowing Newton-based iterations to converge polynomially while avoiding direct handling of the kink in the objective. These approaches outperform simplex methods in large-scale instances, as demonstrated in discrete approximation problems where IPMs achieve convergence in fewer iterations for high-dimensional data. In , IPMs facilitate solving support vector machines (SVMs) by treating them as programs (QPs), where the dual form minimizes a subject to linear inequalities. Primal-dual IPMs, such as the Mehrotra predictor-corrector variant, exploit low-rank structure in the kernel matrix to scale to massive datasets with millions of variables, solving problems in hours that other methods fail to handle due to memory constraints. For instance, on datasets with 60 million observations, these methods reduce computation time by factors of 1.6 to over 1,000 compared to decomposition-based solvers like SVMTorch, while maintaining through out-of-core data processing. Additionally, in spectral methods, IPMs solve semidefinite programs () for learning optimal kernel matrices, optimizing kernels to maximize margins in SVMs or minimize relative entropy in Gaussian processes. This SDP formulation, solved via path-following IPMs, enables kernel alignment for tasks, achieving polynomial-time complexity O((m + n)^2 n^{2.5}) where n is the data size, outperforming kernel selections in accuracy on datasets. Signal processing leverages IPMs for sparse recovery through \ell_1-norm minimization, known as basis pursuit, which solves \min \|x\|_1 subject to Ax = b to reconstruct signals from underdetermined measurements. IPMs reformulate this using auxiliary variables and apply primal-dual steps along the central path, with conjugate gradient solvers for the Newton systems to manage high dimensionality. The l1-magic implements this for compressive sensing, recovering sparse signals (e.g., 20 spikes in 512 dimensions from 120 measurements) with under restricted conditions. These methods provide optimality guarantees absent in alternatives like orthogonal , and scale to problems with thousands of variables via efficient barrier parameter updates. In control systems, IPMs address robust (MPC) involving constraints, formulating the problem as a robust that minimizes worst-case costs over uncertain linear models subject to state and input bounds. Primal-dual IPMs solve the Karush-Kuhn-Tucker conditions iteratively, using Riccati to reduce complexity linearly with the horizon, enabling implementation for systems with over 1,000 variables and 5,000 constraints in minutes. For a double-tank with five models and horizon 50, these methods yield controllers that outperform nominal MPC by reducing peak deviations under disturbances by up to 30%. This approach handles the sets via semidefinite relaxations when needed, providing robust guarantees. Compared to alternatives like active-set or methods, IPMs excel in handling large-scale relaxations in these domains due to their worst-case complexity and ability to exploit problem structure, such as sparsity or low-rank updates, for faster in practice—often requiring fewer iterations (10–50) than subgradient methods while avoiding local optima.

Recent Advances

Quantum Interior-Point Methods

Quantum interior-point methods (QIPMs) adapt classical interior-point methods (IPMs) to by leveraging quantum linear system solvers, such as the Harrow-Hassidim-Lloyd (, to solve the Karush-Kuhn-Tucker (KKT) systems that arise in primal-dual steps. These systems, which involve large-scale inversions for computing search directions, are encoded into quantum states, enabling solutions in superposition and potentially exponential speedups for high-dimensional problems. In (SDP), recent QIPMs formulate the optimization by representing primal and dual variables as quantum-accessible matrices, where the step reduces to solving a of the form A \Delta x = b, with A derived from the of the . The inverts A with query complexity scaling as O(\kappa \log n / \epsilon), where \kappa is the , n the problem , and \epsilon the desired , contrasting with classical direct solvers' O(n^3) time per iteration. Two prominent approaches include an inexact that computes approximate directions without feasibility guarantees and a feasible variant using nullspace reductions to maintain , both converging to optimal solutions under standard IPM assumptions like self-concordance. These methods achieve speedups in n over classical IPMs, which typically require O(n^{3.5} L) total time for with bit length L, though quantum versions exhibit worse dependence on \kappa and \epsilon. Advancements from 2023 to 2025 have refined QIPMs for practical settings, such as preconditioned inexact infeasible variants that reduce the from quadratic to linear dependence on the , improving accuracy and dimension scaling for linear and semidefinite programs. For instance, in , iterative refinement techniques integrated with QIPMs yield exponential improvements in worst-case runtime by mitigating HHL's precision limitations. However, challenges persist, including high qubit demands—often O(n^2 \log n) for n \times n SDP matrices—sensitivity to , and the necessity for error-corrected qubits to achieve , limiting current implementations to noisy intermediate-scale quantum devices. QIPMs hold promise for applications like in , where semidefinite relaxations compute reduced density matrices for molecular ground states, bypassing full wavefunction calculations. Accelerating these SDPs quantumly could enable simulations of larger systems, though realizations remain theoretical pending scalable .

and Optimization

Interior-point methods (IPMs) have been adapted for and optimization to meet the stringent computational constraints of resource-limited hardware in and autonomous systems. Recent developments, particularly from 2024 to 2025, focus on low-memory implementations that enable efficient solving of problems, such as quadratic programs (QPs) and quadratically constrained quadratic programs (QCQPs), on processors with limited floating-point operations per second () and memory. These customized solvers prioritize predictability and speed over exhaustive precision, making them suitable for applications requiring decisions in milliseconds. Key techniques in these embedded IPMs include truncated Newton methods, which approximate the Newton step by solving linear systems iteratively up to a tolerance rather than exactly, thereby reducing computational overhead while maintaining convergence. Approximate preconditioners, such as incomplete Cholesky factorizations tailored to sparse problem structures, further accelerate iterative solvers by improving conditioning without full matrix inversions. Fixed-point iterations are employed to minimize floating-point operations, often by precomputing sparsity patterns offline for static memory allocation and using predictor-corrector primal-dual frameworks with homogeneous self-dual embedding to handle infeasibility detection efficiently. These adaptations exploit the repetitive nature of real-time problems, allowing for code generation in C for direct deployment on microcontrollers. In applications like (MPC) for drones, these IPMs solve QCQPs online to optimize trajectories under constraints such as obstacle avoidance and energy limits. For instance, in control, the solver computes control inputs by minimizing a quadratic cost over a prediction horizon, achieving solution times of 10-100 milliseconds on embedded hardware, which is critical for maintaining at high frequencies (e.g., 100 Hz). Similar techniques apply to autonomous systems like planetary landers, where real-time ensures safe descent amid uncertainties. Trade-offs in these methods involve leveraging warm-starts from previous iterations to reduce the number of steps, often limiting to 5-10 iterations per solve, at the expense of relaxed accuracy tolerances (e.g., feasibility around $10^{-4} instead of $10^{-8}). This enables deterministic execution but may require periodic full-precision recomputation offline. A notable example is a 2025 solver for problems with nonlinear constraints, which integrates IPM with successive and achieves up to a 10x over the general-purpose solver on benchmarks like quadcopter MPC, demonstrating reliability in both efficiency and .

References

  1. [1]
  2. [2]
    [PDF] Interior-point methods for optimization
    Interior-point methods (IPMs) are used for convex, conic, and nonlinear optimization, and have revolutionized the field of optimization.
  3. [3]
    Twenty-Five Years of Interior Point Methods - PubsOnLine
    The interior-point revolution in optimization: History, recent developments, and lasting consequences. Bulletin of the American Mathematical Society 42(1): ...
  4. [4]
    [PDF] Interior-point methods - Convex Optimization
    ▷ log barrier function for constraints f1(x) ≤ 0,..., fm(x) ≤ 0. 𝜙(x) ... ▷ phase II uses barrier method starting from strictly feasible point found in phase I.
  5. [5]
    [PDF] Lecture 17: Interior Point Methods - cs.Princeton
    Nov 19, 2018 · Interior point methods are based on a very different strategy than the Ellipsoid method. The idea is to start inside the convex set K – i.e. to ...
  6. [6]
    [PDF] Practical Guidelines for Solving Difficult Linear Programs
    dual simplex algorithm might effectively solve a highly primal degenerate problem, and vice versa. 717. Interior point algorithms are not prone to degeneracy ...
  7. [7]
    [PDF] Comparisionofinterior-Point Methods Versus Simplex Algorithms
    Nov 6, 2020 · We have concentrated on the theory and application in structured convex programming of interior- point methods since the polynomial-time ...
  8. [8]
    [PDF] INTERIOR POINT POLYNOMIAL TIME METHODS IN CONVEX ...
    given a self-concordant barrier F for a solid G, one can associate with this barrier interior- point methods for minimizing linear objectives over G in ...
  9. [9]
    [PDF] RECENT DEVELOPMENTS IN INTERIOR- POINT METHODS
    Interior-point methodology has been used as part of the solution strategy in many other optimization contexts as well, including ana- lytic center methods and ...<|control11|><|separator|>
  10. [10]
  11. [11]
    [PDF] The CVX Users' Guide
    CVX provides a special SDP mode that allows this LMI notation to be employed inside CVX models using. Matlab's standard inequality operators >=, <=. ... Special ...
  12. [12]
    (PDF) Ragnar Frisch and interior-point methods - ResearchGate
    Aug 7, 2025 · Frisch's logarithmic potential. 3 The problem and barrier methods. To review the originality of Frisch's contributions ...
  13. [13]
    Interior Point Methods for Linear Programming - PubsOnLine
    Logarithmic Barrier Methods. Logarithmic barrier methods were introduced by Frisch[16] and developed by Fiacco and McCormick.[15] Initially, the.Missing: origins | Show results with:origins
  14. [14]
    Nonlinear Programming - SIAM Publications Library
    Nonlinear Programming: Sequential Unconstrained Minimization Techniques. Author(s):. Anthony V. Fiacco and; Garth P. McCormick ... 1968. Although out of print ...
  15. [15]
    The Sequential Unconstrained Minimization Technique (SUMT ...
    Fiacco, Garth P. McCormick, (1967) The Sequential Unconstrained Minimization Technique (SUMT) Without Parameters. Operations Research 15(5):820-827. https ...
  16. [16]
    Interior Point Methods for Nondifferentiable Optimization | SpringerLink
    We describe the analytic center cutting plane method and its relationship to classical methods of nondifferentiable optimization and column generation.Missing: origins | Show results with:origins
  17. [17]
    [PDF] Introducing Interior-Point Methods for Introductory Operations ...
    Sep 1, 2008 · The logarithmic barrier method was first introduced by Frisch [39] in 1955. The method of analytic centers was suggested by Huard [40] in ...
  18. [18]
    THE INTERIOR-POINT REVOLUTION IN OPTIMIZATION
    Sep 21, 2004 · During the 1970s, barrier methods were superseded, nearly to the point of oblivion, by newly emerging and seemingly more efficient alternatives ...Missing: limitations | Show results with:limitations
  19. [19]
    Interior-point methods - ScienceDirect.com
    The methods that have been proposed to date contain many ingredients, including primal–dual steps, barrier and merit functions, and scaled trust regions. For ...
  20. [20]
    [PDF] Karmarkar's Linear Programming Algorithm - John Hooker
    Jul 19, 2020 · 373-395. Karmarkar, N. 1984c, "A new polynomial-time algorithm for linear programming," presen tation at Carnegie-Mellon University, Pitts.
  21. [21]
    A new polynomial-time algorithm for linear programming
    A new polynomial-time algorithm for linear programming. Author: N. Karmarkar ... Published: 01 December 1984 Publication History. 664citation9,184Downloads.Missing: original paper
  22. [22]
    A polynomial-time algorithm, based on Newton's method, for linear ...
    Jun 15, 1987 · Renegar, J. A polynomial-time algorithm, based on Newton's method, for linear programming. Mathematical Programming 40, 59–93 (1988). https ...
  23. [23]
    Warm-Start Strategies in Interior-Point Methods for Linear ... - SIAM.org
    We describe strategies for recovering a "warm-start" point for the perturbed problem instance from the iterates of the original problem instance. We obtain ...
  24. [24]
    Regularization techniques in interior point methods - ScienceDirect
    It was shown that by regularization, free variables can be handled in a numerically stable way by avoiding column splitting that makes the set of optimal ...Missing: post- | Show results with:post-<|control11|><|separator|>
  25. [25]
    [PDF] Interior-Point Methods - cs.wisc.edu
    Feb 10, 2000 · Abstract. The modern era of interior-point methods dates to 1984, when Karmarkar proposed his algorithm for linear programming.
  26. [26]
    [PDF] Interior Point Algorithms I: Geometric Explanation - Stanford University
    Central Path for Linear Programming. The path. C = {(x(µ),y(µ),s(µ)) ∈ intF : Xs = µe, 0 <µ< ∞}; is called the (primal and dual) central path of linear ...
  27. [27]
    Pathways to the Optimal Set in Linear Programming - SpringerLink
    This chapter presents continuous paths leading to the set of optimal solutions of a linear programming problem. These paths are derived from the weighted ...
  28. [28]
    [PDF] 1 RLS and the Representer Theorem - MIT
    Feb 27, 2012 · The KKT conditions can be grouped into four categories: stationarity, primal feasibility, dual feasibility and complementary slackness.
  29. [29]
    [PDF] 12. Interior-point methods
    • we assume problem is strictly feasible; hence strong duality holds and dual optimum is attained examples of greatest interest: SOCP, SDP. Interior-point ...
  30. [30]
    [PDF] Lecture 26 1 Interior-Point Methods for Linear Programming
    Dec 4, 2008 · xT s = xT (c − AT y) = cT x − bT y. Thus this is the difference between the primal and dual objective functions; we call this difference.Missing: x_i s_i
  31. [31]
    [PDF] Interior Point Method
    Dense problems, however, are not in general degenerate so simplex-type methods are the better choice. Example. As an example, consider the problem that ...Missing: avoid | Show results with:avoid
  32. [32]
    HTML - Stanford Engineering Everywhere
    You just check the duality gap. And so they're actually returning essentially the suboptimal point and the certificate proving its epsilon suboptimal, okay? So ...<|separator|>
  33. [33]
    [PDF] Linear Programming: Interior-Point Methods - cs.wisc.edu
    Interior-point methods approach the boundary of the feasible set only in the limit.
  34. [34]
    [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-.Missing: seminal | Show results with:seminal
  35. [35]
    A Primal-Dual Interior Point Algorithm for Linear Programming
    This chapter presents an algorithm that works simultaneously on primal and dual linear programming problems and generates a sequence of pairs of their interior ...
  36. [36]
    Interior Point Methods for Linear Optimization - SpringerLink
    Book Title: Interior Point Methods for Linear Optimization. Authors: Cornelis Roos, Tamás Terlaky, Jean-Philiipe Vial. DOI: https://doi.org/10.1007/b100325.
  37. [37]
    On the Implementation of a Primal-Dual Interior Point Method
    This paper gives an approach to implementing a second-order primal-dual interior point method. It uses a Taylor polynomial of second order to approximate a ...
  38. [38]
    [PDF] 12. Interior-point methods
    = 1, duality gap 100). • terminates when t = 10. 8. (gap 10−6). • centering ... (t) = −(1/t)F(x. ⋆. (t))−1 is feasible for maximize tr(GZ) subject to tr ...Missing: x_i s_i
  39. [39]
  40. [40]
    [PDF] A Mathematical View of Interior-Point Methods for Convex Optimization
    Jul 28, 1998 · The primordial self-concordant barrier functional is the \logarithmic barrier function for the non-negative orthant" having domain Df := <n.Missing: 1988 | Show results with:1988
  41. [41]
    [PDF] Lecture 15 Central path and interior-point methods - MIT
    Apr 23, 2024 · Definition 1.3 (Barrier function). A strongly nondegenerate self-concordant barrier (for us, simply barrier) is a strongly nondegenerate ...
  42. [42]
    [PDF] Interior Point Methods 25 Years Later∗ - School of Mathematics
    Oct 14, 2010 · Interior point methods for optimization have been around for more than 25 years now. Their presence has shaken up the field of optimization.Missing: adoption | Show results with:adoption
  43. [43]
    [PDF] Initialization: The Big-M Formulation
    The idea behind this approach, which is naturally called the big-M method, is that although the value of A1 may be positive initially, but with this added term ...
  44. [44]
  45. [45]
    [PDF] Interior-Point Polynomial Algorithms in Convex Programming
    The latter scheme originating from Huard (see, e.g., [BH 66]) leads to what is called methods of centers. Note that the above schemes possess two main ...<|separator|>
  46. [46]
    [PDF] Interior-point methods for optimization - Cornell University
    In Section 2, we discuss self-concordant barriers and their properties, and then describe interior-point methods for both general convex optimization.
  47. [47]
    On Central-Path Proximity Measures in Interior-Point Methods
    Sep 2, 2025 · Measuring how close the iterates are to the central path is an important aspect of such methods and it is accomplished by using proximity ...Missing: distance c
  48. [48]
    [PDF] Primal-Dual Interior-Point Methods
    Primal-dual interior-point methods solve the same problems as barrier methods, taking one Newton step, and are often more efficient, but less intuitive.
  49. [49]
    Superlinear Convergence of Infeasible Predictor-Corrector Path ...
    Jan 24, 2010 · In this paper, we give a weak sufficient condition using these off-central paths that guarantees superlinear convergence of a predictor- ...
  50. [50]
    Superlinear Convergence of an Infeasible Predictor-Corrector Path ...
    In this paper, we give a sufficient condition using these off-central paths that guarantees superlinear convergence of a predictor-corrector path-following ...
  51. [51]
    [PDF] Constrained Optimization PART 4: Introduction to Interior Point
    The KKT system becomes very ill-conditioned as we approach the solution, because some components of X−1 get very large. Even so, the ill-conditioning does ...
  52. [52]
    A Preconditioner for Linear Systems Arising From Interior Point ...
    We explore a preconditioning technique applied to the problem of solving linear systems arising from primal-dual interior point algorithms in linear and ...
  53. [53]
    An Interior-Point Method for Large-Scale l1-Regularized Least ...
    In this paper we describe a specialized interior-point method for solving large-scale l1-regularized LSPs that uses the preconditioned conjugate gradients (PCG) ...
  54. [54]
    On Implementing Mehrotra's Predictor–Corrector Interior-Point ...
    This paper describes a full implementation of this algorithm, with extensions for solving problems with free variables and problems with bounds on primal ...
  55. [55]
    [PDF] warm-start strategies in interior-point methods for linear programming
    This paper describes and analyzes warm-start strategies for interior-point methods applied to linear programming (LP) problems. We consider the situation in ...Missing: 1984 | Show results with:1984
  56. [56]
    Block preconditioners for linear systems in interior point methods for ...
    Aug 18, 2022 · In this paper, we address the preconditioned iterative solution of the saddle-point linear systems arising from the (regularized) Interior Point method.
  57. [57]
    Ipopt - COIN-OR Documentation
    This document is a guide to using Ipopt. It includes instructions on how to obtain and compile Ipopt, a description of the interface, user options, etc., ...Installing Ipopt · Ipopt Options · Ipopt Output · Ipopt Namespace Reference
  58. [58]
    13.3 Conic Optimization - Interior-point optimizer - Mosek
    The interior-point optimizer is an implementation of the so-called homogeneous and self-dual algorithm. For a detailed description of the algorithm, please see ...
  59. [59]
    [PDF] Advanced Gurobi Algorithms
    • Now we have an interior point method, but what makes it a barrier method? ... MIP solvers find new feasible solutions in two ways. • Branching. • Primal ...
  60. [60]
    Combining Interior-Point and Pivoting Algorithms for Linear ...
    We propose a new approach to combine linear programming (LP) interior-point and simplex pivoting algorithms. In any iteration of an interior-point algorithm ...
  61. [61]
    [PDF] On the -perturbation Method for Avoiding Degeneracy
    Our aim here is to point out that for theoretical purposes degeneracy can easily be dispensed with in polynomial time. The basic idea is an old one and is ...
  62. [62]
    Interior-Point Polynomial Algorithms in Convex Programming | SIAM Publications Library
    ### Summary of Key Contributions to SDP in Interior-Point Methods, Barrier Functions, and Complexity
  63. [63]
    [PDF] interior point methods in semidefinite programming - UF CISE
    Next we present an interior point algorithm which converges to the optimal solution in polynomial time. The approach is a direct extension of Ye,s projective ...
  64. [64]
    [PDF] Semidefinite Programming - Stanford University
    A worst-case analysis of interior-point methods forsemidef- inite programming shows that the effort required to solve a semidefinite program to a given accuracy ...
  65. [65]
    [PDF] Applications of Semidefinite Programming
    Aug 25, 1998 · It has also been recognized in combinatorial optimization as a valuable technique for obtaining bounds on the solution of NP-hard problems. The ...
  66. [66]
    [PDF] Semidefinite Programming and Combinatorial Optimization
    In this paper, we describe a few applications of semidefinite programming in combinatorial optimization. Because of space limitations, we restrict our attention.
  67. [67]
    Interior Point Methods in Semidefinite Programming with ...
    This paper studies the semidefinite programming SDP problem, ie, the optimization problem of a linear function of a symmetric matrix subject to linear equality ...
  68. [68]
    [PDF] A tutorial on geometric programming - Stanford University
    Apr 10, 2007 · In addition to being fast, interior-point methods for GPs are also very robust. They require essentially no algorithm parameter tuning, and they ...
  69. [69]
    [PDF] Interior-point Methods, Cone Programming, and Applications
    control, combinatorial optimization, signal processing, circuit design, communications, . . . • robust optimization robust versions of LP, LS, other problems.
  70. [70]
    [PDF] LOQO: AN INTERIOR POINT CODE FOR QUADRATIC ...
    Oct 6, 1998 · This paper describes a software package, called LOQO, which implements a primal- dual interior-point method for general nonlinear programming.
  71. [71]
    Solving the discrete lp-approximation problem by a method of centers
    We report our experience in solving the discrete lp-approximation problem by an interior point method. We discuss the convergence property of the algorithm and ...
  72. [72]
    Interior-Point Methods for Massive Support Vector Machines
    We investigate the use of interior-point methods for solving quadratic programming problems with a small number of linear constraints, where the quadratic ...
  73. [73]
    [PDF] Learning the Kernel Matrix with Semidefinite Programming
    Kernel-based learning algorithms work by embedding the data into a Euclidean space, and then searching for linear relations among the embedded data points.
  74. [74]
    [PDF] l1-magic : Recovery of Sparse Signals via Convex Programming
    In the next section, we describe how to solve linear and second-order cone programs using modern interior point methods. 2 Interior point methods. Advances in ...
  75. [75]
    Interior-Point Methods in l 1 Optimal Sparse Representation ...
    In this paper, we consider the basis pursuit principle to find the representation (frequency) coefficients of a harmonic signal by minimizing the l 1 norm. For ...Missing: l1 | Show results with:l1
  76. [76]
    [PDF] Robust Optimal Control of Linear Discrete-Time Systems Using ...
    In this paper is described how to efficiently solve a robust op- timal control problem using recently developed primal-dual interior-point methods.
  77. [77]
    [PDF] Application of Interior-Point Methods to Model Predictive Control
    We start with a general description of the interior-point method of choice for linear and convex quadratic programming: Mehrotra's predictor-corrector algorithm ...
  78. [78]
    Quantum Interior Point Methods for Semidefinite Optimization
    Sep 11, 2023 · We present two quantum interior point methods for semidefinite optimization problems, building on recent advances in quantum linear system algorithms.
  79. [79]
    A preconditioned inexact infeasible quantum interior point method ...
    Dec 15, 2024 · Quantum Interior Point Methods (QIPMs) have been attracting significant interests recently due to their potential of solving optimization ...
  80. [80]
    Improvements to Quantum Interior Point Method for Linear ...
    Jan 14, 2025 · Another class of quantum algorithms includes the Quantum Interior Point Methods ... Generally, the HHL algorithm starts from quantum state ...
  81. [81]
    Quantum Computing Inspired Iterative Refinement for Semidefinite ...
    Dec 18, 2023 · We also show that using IR with Quantum Interior Point Methods (QIPMs) leads to exponential improvements in the worst-case overall running ...
  82. [82]
    Realization of Quantum Chemistry without Wave Functions through ...
    Nov 15, 2004 · Through efficient first-order semidefinite programming, the variational 2-RDM method practically realizes the dream of a quantum chemistry ...
  83. [83]
    [PDF] An Interior-Point LPV MPC Solver for Real-time Systems - HAL
    Apr 9, 2025 · Therefore, in this work, we develop and validate a novel library- independent embedded optimisation based on the LPV MPC approach. In particular ...