Fact-checked by Grok 2 weeks ago

Cost function

A cost function, also known as an objective function in optimization contexts, is a mathematical that assigns a numerical value representing the "cost" or penalty to each possible decision or configuration within a problem, which is typically minimized to identify the optimal solution subject to constraints. In the field of optimization and , cost functions are central to formulating problems such as , where they quantify objectives like production costs or penalties; for instance, in linear optimization, the cost function may take an affine form, such as c(x) = a \cdot x + b, where x represents decision variables, and minimization occurs over a feasible set defined by linear inequalities. These functions are often linear or to ensure tractable solutions via methods like the , and they can represent real-world scenarios, such as minimizing transportation costs in . In , particularly microeconomic theory, the describes the minimum required to produce a specified level of output given input prices, formally defined as C(y, w) = \min \{ w \cdot x : x \in V(y) \}, where y is the output , w > 0 is the of input prices, x is the input , and V(y) is the input requirement set. Key properties include non-negativity (C(y, w) \geq 0), positive linear homogeneity in prices (C(y, \lambda w) = \lambda C(y, w) for \lambda > 0), concavity in w, and monotonicity in output y, making it a to the and useful for analyzing firm behavior under varying market conditions. In , especially , a cost function (often interchangeably called a ) measures the discrepancy between model predictions and actual data, aggregated over examples to guide parameter optimization; for , it is commonly the : J(\theta) = \frac{1}{2m} \sum_{i=1}^m (h_\theta(x^{(i)}) - y^{(i)})^2, where \theta are model parameters, h_\theta is the function, m is the number of examples, x^{(i)} are inputs, and y^{(i)} are targets. This function is typically convex, enabling efficient minimization via , and its choice (e.g., squared error for or for ) directly impacts model performance and generalization.

General Concepts

Definition and Notation

A cost function is a mathematical construct that maps an input vector to a non-negative scalar value, quantifying the associated cost, error, or penalty, and is typically minimized in optimization problems across various disciplines. Formally, it is defined as C: \mathbb{R}^n \to \mathbb{R}_{\geq 0}, where the domain \mathbb{R}^n represents the space of inputs such as production quantities or model parameters, and the range ensures the output is a penalty measure starting from zero. This structure assumes familiarity with basic calculus concepts, including functions from multivariable inputs to scalar outputs and the notion of minimization as finding the input that yields the smallest output value. Common examples illustrate the form's simplicity and utility; for instance, a cost function C(x) = x^2 assigns costs proportional to the square of deviations from an optimal point, often arising in least-squares problems or penalty terms. More generally, the function can incorporate linear or nonlinear terms, but retains the core property of non-negativity to reflect cumulative penalties without offsets. Notation for cost functions varies by but follows consistent conventions for clarity. In general optimization and , C(\mathbf{x}) or C(\theta) is standard, with \mathbf{x} or \theta denoting the variable , such as input factors or parameters. In optimization literature, it may appear as the objective function f_0(\mathbf{x}), emphasizing its role in problem formulations like \min_{\mathbf{x}} f_0(\mathbf{x}). and statistical contexts often use J(\theta), where \theta specifically highlights tunable parameters, tracing back to dynamic programming traditions for "cost-to-go." This notation distinguishes cost functions from profit functions, which are maximized to capture net benefits (e.g., minus costs), or functions, which ordinalize preferences rather than penalize deviations. In , cost functions model production expenses as minimized expenditures for given outputs, while in , they evaluate predictive errors as loss functions.

Historical Development

The concept of the cost function originated in the mathematical field of during the 18th century, where it served as a tool for minimizing functionals representing physical or geometric quantities. Leonhard Euler pioneered this approach in his early work on variational problems around 1736, developing methods to find curves or paths that extremize integrals, such as those arising in . Euler's systematic treatment appeared in his 1744 Methodus inveniendi lineas curvas maximi minimive proprietate gaudentes, which established the foundations for solving optimization problems through differential equations derived from variational principles. Joseph-Louis Lagrange built upon Euler's ideas in the 1760s, providing a more general analytic framework. In his 1760–1761 memoir "Essai d'une nouvelle méthode pour déterminer les maxima et les minima des formules intégrales définies," Lagrange derived the Euler-Lagrange equations, which express the necessary conditions for a functional to achieve an extremum, thus formalizing the minimization of cost-like integrals without geometric intuitions. This advancement shifted the calculus of variations toward algebraic methods, influencing subsequent optimization theories. The economic interpretation of cost functions gained prominence in the late . formalized cost curves in his 1890 treatise Principles of Economics, where he linked production costs to output quantities, deriving supply schedules from considerations and integrating them with demand analysis. This represented a key step in , treating costs as functions of input factors and scale. In the mid-20th century, cost functions became integral to and optimization. Following , formulated in 1947, employing cost coefficients in objective functions to minimize linear combinations of variables subject to constraints, as applied to and . Concurrently, John von Neumann's contributions to in the 1940s, detailed in Theory of Games and Economic Behavior (1944) co-authored with , incorporated cost elements into payoff matrices for strategic interactions. The adoption of cost functions in began in the 1950s with early neural models. Frank Rosenblatt's , introduced in his 1958 paper "The Perceptron: A Probabilistic Model for Storage and Organization in the ," used error measures to iteratively adjust weights, effectively minimizing misclassification costs in binary tasks. This approach evolved significantly in the 1980s, as David E. Rumelhart, Geoffrey E. Hinton, and Ronald J. Williams demonstrated in their 1986 paper "Learning Representations by Back-Propagating Errors" how on differentiable loss functions—analogous to cost functions—could train multilayer networks efficiently.

Applications in Economics

Production Cost Models

In economics, the cost function represents the minimum total cost of producing a given of output, denoted as TC(q), where q is the output level. This function incorporates both fixed costs, which do not vary with output, and variable costs, which do. Fixed costs cover expenses like plant and equipment that remain constant regardless of volume, while variable costs include inputs such as labor and materials that with q. The total cost function thus captures the relationship between production scale and the resources required to achieve it under efficient operations. Common functional forms for the cost function reflect different assumptions about returns to scale and input efficiencies. A linear form, TC(q) = F + v q, assumes constant marginal cost v, implying no diminishing returns to variable inputs and constant average variable costs; this is suitable for scenarios with fixed capacity and linear input-output relationships. A quadratic form, such as TC(q) = F + v q + c q^2 with c > 0, models increasing marginal costs due to diminishing returns, leading to rising average costs at higher output levels. For more complex dynamics, a cubic form like TC(q) = F + v q + c q^2 + d q^3 (with appropriate coefficients) can produce a U-shaped average cost curve, where costs initially fall due to spreading fixed expenses before rising from inefficiencies. These forms are chosen based on empirical fit to data, with linear for constant returns, quadratic for early-stage increasing returns, and cubic for full U-shaped patterns observed in many industries. Deriving the cost function typically assumes in input markets, where producers are price-takers facing given input prices, and rational behavior in minimizing costs subject to a . Producers are assumed to optimize by selecting input combinations that achieve at least the target output q at lowest cost. For a two-input like the Cobb-Douglas form Q = A L^\alpha K^\beta, where L is labor, K is , A > 0 is , and $0 < \alpha, \beta < 1 reflect elasticities, the cost function emerges from solving the minimization problem. With input prices w for labor and r for , the total cost is TC(q) = \min \{ w L + r K \mid A L^\alpha K^\beta \geq q \}. Under constant returns to scale (\alpha + \beta = 1), this yields an explicit form TC(q) = q \left( \frac{w^\alpha r^\beta}{A \alpha^\alpha \beta^\beta} \right)^{1/(\alpha + \beta)}, showing costs scaling linearly with output after fixed adjustments. This derivation relies on the marginal rate of technical substitution equaling the input price ratio at the optimum, ensuring efficient input use. Empirically, cost functions often exhibit , where average cost AC(q) = TC(q)/q decreases initially as output rises—due to fixed costs being spread over more units and early efficiencies—before increasing due to diseconomies like managerial complexities or resource constraints. This results in a U-shaped average cost curve, commonly modeled by cubic forms to capture the transition from in real-world production processes. Such patterns are verified through regression on firm-level data, informing decisions on optimal plant size and output levels.

Short-Run and Long-Run Costs

In the short run, firms face constraints where at least one input, such as capital stock K, is fixed, while others, like labor L, are variable. The short-run total cost function is expressed as TC_{SR}(q) = w L(q, K) + r K, where w is the wage rate, r is the rental rate of capital, and q denotes output. Due to the law of diminishing marginal returns on the variable input, the marginal product of labor decreases as more labor is added to the fixed capital, leading to an upward-sloping segment in the short-run marginal cost curve MC_{SR}(q). This results in a characteristic U-shaped MC_{SR}(q) curve, initially declining due to increasing returns and then rising as diminishing returns dominate. In contrast, the long run allows all inputs to vary, enabling firms to adjust the fixed factors like capital to their optimal levels for any output q. The long-run total cost function is thus defined as the minimum over possible fixed input levels: TC_{LR}(q) = \min_K TC_{SR}(q \mid K). This makes the long-run average cost curve LAC(q) = TC_{LR}(q)/q the lower envelope of the family of short-run average cost curves SAC(q \mid K), where each SAC corresponds to a different fixed K, and tangency occurs at the output level where the chosen K is optimal for that q. The envelope theorem provides a key insight into the relationship between short-run and long-run marginal costs: at the optimal fixed input level for a given q, the derivative of the long-run total cost with respect to output equals the short-run marginal cost, so \frac{d TC_{LR}}{dq} = MC_{SR}(q \mid K^*), where K^* minimizes TC_{SR}. This equality holds because the indirect effect of output changes on the optimal fixed input is zero at the tangency point. In competitive industries, free entry and exit in the long run drive the market price down to the minimum of the LAC curve, aligning firm output with the minimum efficient scale where average costs are lowest.

Applications in Optimization

Role as Objective Function

In mathematical optimization, the cost function, denoted as f(\mathbf{x}), serves as the objective function that quantifies the performance or penalty associated with a decision variable \mathbf{x}, with the primary goal of minimizing it to identify optimal solutions. The canonical formulation of such problems is \min_{\mathbf{x} \in \mathbb{R}^n} f(\mathbf{x}) subject to inequality constraints g_i(\mathbf{x}) \leq 0 for i = 1, \dots, m and equality constraints h_j(\mathbf{x}) = 0 for j = 1, \dots, p, where the constraints define the feasible region over which the minimization occurs. This structure underpins optimization across disciplines, from engineering design to resource allocation, by encoding the trade-offs in achieving desirable outcomes. For unconstrained optimization problems, where no explicit constraints restrict \mathbf{x}, the minimization directly targets critical points of f(\mathbf{x}). Necessary first-order conditions require the gradient to vanish at a local minimum \mathbf{x}^*, i.e., \nabla f(\mathbf{x}^*) = 0, while second-order conditions stipulate that the Hessian matrix \nabla^2 f(\mathbf{x}^*) must be positive semi-definite to confirm a local minimum. In contrast, constrained optimization incorporates the restrictions through the Lagrangian function for equality constraints: \mathcal{L}(\mathbf{x}, \boldsymbol{\lambda}) = f(\mathbf{x}) + \sum_{j=1}^p \lambda_j h_j(\mathbf{x}), where \boldsymbol{\lambda} are the . Optimality then demands stationarity conditions \nabla_{\mathbf{x}} \mathcal{L}(\mathbf{x}^*, \boldsymbol{\lambda}^*) = 0 and h_j(\mathbf{x}^*) = 0, ensuring the objective aligns with the constraints at the solution. These conditions extend to inequalities via , but the Lagrangian framework remains foundational for handling equality-bound problems. When the cost function f(\mathbf{x}) is non-convex, optimization faces significant challenges due to the potential for multiple local minima and saddle points in the function's landscape, complicating the identification of the global minimum. Gradient descent and similar local search methods may converge to suboptimal local minima, necessitating advanced techniques like stochastic perturbations or global search heuristics to navigate these irregularities effectively. Such non-convexity arises frequently in real-world applications with intricate objective surfaces, highlighting the distinction between local and global optimality in practical problem-solving. In decision theory, the cost function embodies the negative of utility, framing minimization of expected cost as equivalent to maximization of expected utility under uncertainty, a principle central to rational choice models. This perspective, rooted in axiomatic foundations, allows cost minimization to capture risk-averse behaviors by weighting outcomes probabilistically.

Properties and Analysis

Cost functions in optimization exhibit several key mathematical properties that significantly influence the behavior of optimization algorithms and the solvability of associated problems. Among these, convexity is a fundamental property that ensures reliable global optimization. A function f: \mathbb{R}^n \to \mathbb{R} is convex if its domain is convex and satisfies f(\lambda x + (1-\lambda) y) \leq \lambda f(x) + (1-\lambda) f(y) for all x, y in the domain and \lambda \in [0,1]. This property implies that any local minimum is also a global minimum, and if the function is strictly convex, the global minimum is unique. Convex cost functions are particularly amenable to subgradient methods, which generalize gradient descent to handle non-differentiable points while guaranteeing convergence to the optimum under appropriate conditions. Differentiability further refines the analysis of cost functions by enabling the use of first-order information for optimization. A differentiable cost function allows the application of , where updates proceed in the direction of the negative gradient to reduce the function value iteratively. However, many practical cost functions are non-smooth, such as the L1 norm \|x\|_1 = \sum_{i=1}^n |x_i|, which is convex but not differentiable at zero. For such cases, optimization relies on or , which extend the notion of gradients to non-differentiable convex functions and facilitate convergence in . Continuity and Lipschitz continuity provide guarantees on the boundedness and stability of cost functions, which are crucial for algorithmic convergence. A function f is continuous with constant L > 0 if |f(x) - f(y)| \leq L \|x - y\| for all x, y in the domain, ensuring that small changes in the input lead to bounded changes in the output. When the is continuous—meaning \|\nabla f(x) - \nabla f(y)\| \leq L \|x - y\|—this bounds the variation in gradients, which is essential for establishing convergence rates in methods like . These properties collectively prevent erratic behavior in high-dimensional spaces and support theoretical analyses of optimization trajectories. In large-scale optimization, the scalability of cost functions is challenged by high dimensionality, often manifesting as the curse of dimensionality. This phenomenon, first articulated by Richard Bellman, describes how the volume of the search space grows exponentially with the number of variables, leading to increased computational demands and potential sparsity in feasible regions. For cost functions defined over high-dimensional domains, such as those in or engineering design, this effect complicates uniform sampling and gradient estimation, often requiring techniques or structured assumptions to maintain tractability. Sensitivity analysis examines how perturbations in the cost function affect the location and value of the optimum, providing insights into the robustness of solutions. Under suitable conditions, such as differentiability of the cost function and constraint qualifications, the guarantees that small changes in parameters defining f result in continuously varying optimal points. For instance, if the optimum satisfies \nabla f(x^*) = 0, perturbations \delta f induce shifts \delta x^* \approx -[\nabla^2 f(x^*)]^{-1} \nabla (\delta f)(x^*), allowing quantification of . This analysis is vital for understanding the reliability of optimized solutions in uncertain or parameterized environments.

Applications in Machine Learning

As Loss Functions

In machine learning, a cost function is adapted as a loss function, denoted L(\theta; D), to measure the discrepancy between a model's predictions \hat{y}(\theta, x) and the corresponding true labels y across a dataset D, typically by averaging this discrepancy over all samples in the dataset. This formulation enables the quantification of how well the parameterized model, governed by \theta, fits the observed data, serving as the core mechanism for guiding iterative improvements during training. The central training objective in supervised is , which seeks to identify \theta that minimize the average over the training : \min_{\theta} \frac{1}{|D|} \sum_{i=1}^{|D|} L(\hat{y}(\theta, x_i), y_i) This empirical risk approximates the true risk, defined as the expected \mathbb{E}[L] over the underlying data distribution, providing a practical surrogate for optimizing performance. computation and updates can occur in batch learning, where the full is used to evaluate the loss and perform a single gradient step, or in , where updates are made incrementally using individual samples or small mini-batches, as facilitated by to handle large-scale or more efficiently. To prevent , where models excessively fit data at the expense of , regularization terms are incorporated into the , such as L_{\text{reg}}(\theta) = L(\theta) + \lambda \|\theta\|^2, which imposes a penalty on magnitude and promotes simpler models. While functions drive the optimization process during by providing differentiable signals for gradient-based updates, they differ from evaluation metrics like accuracy, which assess overall model performance on held-out test data and may prioritize interpretability over direct optimizability.

Common Examples and Selection

In , the (MSE) serves as a fundamental cost function for tasks, defined as L(y, \hat{y}) = \frac{(y - \hat{y})^2}{2}, where y is the true value and \hat{y} is the predicted value. This formulation, rooted in the method, is differentiable, enabling efficient gradient-based optimization, and quadratically penalizes larger errors to emphasize accurate predictions for continuous outputs. For problems, the loss, also known as binary cross-entropy or log loss, is widely used, given by L(y, \hat{y}) = - y \log(\hat{y}) - (1-y) \log(1-\hat{y}), where y \in \{0,1\} and \hat{y} is the predicted probability. This function aligns closely with probabilistic model outputs from , maximizing likelihood by heavily penalizing confident incorrect predictions while being less sensitive to correct ones near certainty. In support vector machines (SVMs) for classification, the provides a margin-based approach, formulated as L(y, \hat{y}) = \max(0, 1 - y \hat{y}), where y \in \{-1, 1\} and \hat{y} is the model's raw output. Unlike differentiable losses, it is non-differentiable at the margin boundary, promoting sparsity by focusing on misclassified or near-margin points to maximize the separation between classes. The Huber loss addresses robustness in by hybridizing MSE and (), defined piecewise as L_\delta(y, \hat{y}) = \begin{cases} 0.5 (y - \hat{y})^2 & \text{if } |y - \hat{y}| < \delta \\ \delta |y - \hat{y}| - 0.5 \delta^2 & \text{otherwise} \end{cases}, with \delta typically set around 1.35 for assumptions. This design behaves quadratically for small errors like MSE but linearly for large ones like , reducing influence while remaining differentiable except at \pm \delta. Selecting an appropriate cost function depends on the task type, data characteristics, and practical constraints. For , MSE suits assumptions, but or is preferred when outliers are prevalent to enhance robustness, as MSE's penalty amplifies extreme deviations. tasks typically require probabilistic losses like for models outputting probabilities, whereas margin-based fits SVMs emphasizing decision boundaries over probabilities. Computational efficiency also guides choices; differentiable functions like MSE and integrate seamlessly with in neural networks, while non-differentiable ones like hinge may need subgradient methods, potentially increasing training time.

References

  1. [1]
    [PDF] Lecture 1: linear optimization
    Definition of cost / objective function. • Example of cost functions, affine functions, linear functions. • Definition of constraints.
  2. [2]
    [PDF] Linear Programming
    For example, the objective function may measure the profit or cost that occurs as a function of the amounts of various products produced. The objective ...
  3. [3]
    [PDF] COST FUNCTIONS 1.1. Understanding and representing ...
    The cost function is defined for output and input prices, and is the minimum cost to produce an output, defined as C(y, w) = min {wx : xǫV (y)}, y ∈ DomV,w > 0.
  4. [4]
    [PDF] CS229 Lecture notes
    We define the cost function: J(θ) = 1. 2 m. X i=1. (hθ(x(i)) − y(i))2. If you've seen linear regression before, you may recognize this as the familiar least ...<|control11|><|separator|>
  5. [5]
    None
    Below is a merged summary of the cost function and objective function in convex optimization, consolidating all information from the provided segments into a comprehensive response. To retain maximum detail and ensure clarity, I will use a structured format with tables where appropriate, followed by a narrative explanation. The response avoids redundancy while preserving all key details, including definitions, general forms, domains, ranges, contexts, section references, and useful URLs.
  6. [6]
    Optimization - Calculus I - Pauls Online Math Notes
    Nov 16, 2022 · ... cost function will always be concave up and so w=1.8821 w = 1.8821 must give the absolute minimum cost. All we need to do now is to find the ...
  7. [7]
    [PDF] Part I - Duality of Production, Cost, and Profit Functions
    The definition of the cost function as the result of an optimization yields strong mathematical properties, and establishes the cost function as a.
  8. [8]
    Cost function vs loss function vs error? - DeepLearning.AI Community
    Dec 24, 2024 · Definition: The cost function is the average or total loss over the entire training dataset. It aggregates the individual losses across all ...Missing: notation | Show results with:notation
  9. [9]
    Cost Function is No Rocket Science! - Analytics Vidhya
    Mar 20, 2024 · A cost function, also referred to as a loss function or objective function, is a key concept in machine learning. It quantifies the difference between ...Why to use a Cost function? · What is Cost Function in... · Types of Cost function in...
  10. [10]
    [PDF] LEONHARD EULER, BOOK ON THE CALCULUS OF VARIATIONS ...
    In this book Euler extended known methods of the calculus of variations to form and solve differential equations for the general problem of optimizing ...
  11. [11]
    [PDF] The original Euler's calculus-of-variations method - Edwin F. Taylor
    Leonhard Euler's original version of the calculus of variations (1744) used elementary mathematics and was intuitive, geometric, and easily visualized. In.
  12. [12]
    [PDF] J. L. Lagrange's changing approach to the foundations of the ...
    A central topic of this study concerns LAGRANGE'S changing derivation of the so-called EULER-LAGRANGE equations. Since the calculus of variations in its.
  13. [13]
    [PDF] LINEAR PROGRAMMING
    In the years from the time when it was first proposed in 1947 by the author (in connection with the planning activities of the military), linear programming and ...
  14. [14]
    [PDF] THEORY OF GAMES AND ECONOMIC BEHAVIOR
    The purpose of this book is to present a discussion of some funda,.- mental questions of economic theory which require a treatment different.
  15. [15]
    [PDF] The perceptron: a probabilistic model for information storage ...
    The perceptron: a probabilistic model for information storage and organization in the brain. · Frank Rosenblatt · Published in Psychology Review 1 November 1958 ...
  16. [16]
    Learning representations by back-propagating errors - Nature
    Oct 9, 1986 · Cite this article. Rumelhart, D., Hinton, G. & Williams, R. Learning representations by back-propagating errors. Nature 323, 533–536 (1986).
  17. [17]
    [PDF] Production and Cost Functions - NYU Stern
    Jan 2, 2012 · 11. Page 12. Under this assumption of a Cobb-Douglas production function, the Cost. function has the following form: C(Q, w, r)=¯ω +
  18. [18]
    Cost Functions | Types | Example and Graphs - XPLAIND.com
    Feb 11, 2019 · Typical cost functions are either linear, quadratic and cubic. A linear cost function is such that exponent of quantity is 1. It is ...
  19. [19]
    3 Main Types of Cost Functions - Economics Discussion
    The following points highlight the three main types of cost functions. The types are: 1. Linear Cost Function 2. Quadratic Cost Function 3. Cubic Cost ...
  20. [20]
    [PDF] Perfect Competition - Producer Theory - Columbia University
    The second assumption tells us that the first derivative of the production function has to be positive. We call the first derivative of the production function ...
  21. [21]
    Economies of Scale | Microeconomics - Lumen Learning
    The normal shape for a short-run average cost curve is U-shaped with decreasing average costs at low levels of output and increasing average costs at high ...
  22. [22]
    [PDF] Short Run Cost Functions
    In the short run, one or more inputs are fixed, so the firm chooses the variable inputs to minimize the cost of producing a given amount of output. With several ...
  23. [23]
    [PDF] Cost Functions - UCLA Economics
    In very short run, all inputs are fixed. •. In short run, some inputs fixed with others are flexible. •. In medium run, all inputs ...
  24. [24]
    [PDF] Envelopes for Economists: Housing Hedonics and Other Applications
    Because a long-run cost function is the envelope of a family of short-run cost functions with different plant sizes or scales, a long-run cost function can ...<|control11|><|separator|>
  25. [25]
    8.3 Entry and Exit Decisions in the Long Run - UH Pressbooks
    Entry and exit to and from the market are the driving forces behind a process that, in the long run, pushes the price down to minimum average total costs so ...
  26. [26]
    Numerical Optimization | SpringerLink
    Numerical Optimization presents a comprehensive and up-to-date description of the most effective methods in continuous optimization.Derivative-Free Optimization · Sequential Quadratic... · Quasi-Newton Methods
  27. [27]
    [PDF] 1 Theory of convex functions - Princeton University
    Let's first recall the definition of a convex function. Definition 1. A function f : Rn → R is convex if its domain is a convex set and for all x, y.
  28. [28]
    [PDF] Notes for Optimization Algorithms Spring 2023 - Purdue Math
    2.2.2 Convergence for Lipschitz continuous ∇f . . . . . . . 21. 2.2.3 Convergence for convex functions . . . . . . . . . . . . 24. 2.2.4 Convergence for ...
  29. [29]
    Overcoming the Curse of Dimensionality for Control Theory - IPAM
    Oct 30, 2015 · This is the “curse of dimensionality” a term coined by Richard Bellman in 1957. Jerome Darbon and Stanley Osher were motivated to think about ...
  30. [30]
    [PDF] High Dimensional Geometry, Curse of Dimensionality, Dimension ...
    We encounter the so-called curse of dimensionality which refers to the fact that algorithms are simply harder to design in high dimensions and often have a ...Missing: optimization | Show results with:optimization
  31. [31]
    [PDF] LECTURE 7: CONSTRAINED OPTIMIZATION - NC State ISE
    Sensitivity analysis. • Consider NLP with equality constraints: Page 3. Basic Idea of Implicit Functions. Page 4. Example. Page 5. Implicit function theorem. • ...
  32. [32]
    Sensitivity Analysis of Nonlinear Programs and Differentiability ...
    This paper is concerned with a study of differentiability properties of the optimal value function and an associated optimal solution of a parametrized ...
  33. [33]
    [PDF] a survey and taxonomy of loss functions in machine learning - arXiv
    Nov 18, 2024 · Loss functions are objective functions that measure an algorithm's performance, critical for machine learning, and are defined as a mapping of ...
  34. [34]
    [PDF] Task-based Loss Functions in Computer Vision - arXiv
    Loss functions quantify the difference between predicted outputs and ground truth labels, guiding the optimization process to minimize errors in deep learning.
  35. [35]
    [PDF] An overview of gradient descent optimization algorithms - arXiv
    Jun 15, 2017 · Batch gradient descent performs redundant computations for large datasets, as it recomputes gradients for similar examples before each parameter ...
  36. [36]
    [PDF] On the training dynamics of deep networks with L2 regularization
    We study the role of L2 regularization in deep learning, and uncover simple relations between the performance of the model, the L2 coefficient, the learning ...<|separator|>
  37. [37]
    3.1. Linear Regression - Dive into Deep Learning
    It follows that minimizing the mean squared error is equivalent to the maximum likelihood estimation of a linear model under the assumption of additive Gaussian ...
  38. [38]
    Robust Estimation of a Location Parameter - Project Euclid
    This paper contains a new approach toward a theory of robust estimation; it treats in detail the asymptotic theory of estimating a location parameter for ...