Post

Desirable Properties of Optimizers

A discussion of the key desirable properties for optimization algorithms in machine learning, covering effectiveness, efficiency, robustness, invariance, practicality, and impact on model generalization.

Desirable Properties of Optimizers

In our previous post on iterative methods, we explored how algorithms iteratively search for solutions to optimization problems, distinguishing between gradient-free and gradient-based approaches. Now, we face a crucial question: what makes one optimizer “better” than another? This isn’t just an academic query; choosing the right optimizer can drastically impact training time, model performance, and resource consumption in machine learning.

1. Introduction: Why “Good” Matters in Optimization

Imagine you’re an engineer designing a race car. You could prioritize raw speed, fuel efficiency, handling on tight corners, or reliability over a long race. No single design excels at everything; trade-offs are inevitable. Similarly, in the world of optimization, there’s no universally “best” optimizer. The ideal choice depends heavily on the specific problem we’re trying to solve (e.g., training a deep neural network, fitting a logistic regression), the characteristics of the objective function \(f(x)\) (e.g., its shape, smoothness, dimensionality), available computational resources, and our ultimate goals (e.g., fast training vs. best possible generalization).

This post aims to establish a “wish list”—a set of desirable properties that we look for in an optimization algorithm. Understanding these properties will provide a framework for:

  • Evaluating and comparing existing optimizers.
  • Appreciating the design choices behind different algorithms.
  • Guiding the development of new, improved optimization methods.

We’ll group these properties into categories: core effectiveness, efficiency, robustness (including key mathematical invariances), and practical usability.

2. Core Effectiveness: Reaching the Goal and Solution Quality

The fundamental task of an optimizer is to find a solution. But “finding a solution” has several layers of meaning.

2.1. Convergence: The Journey’s End

The most basic requirement is that the optimizer should converge to some meaningful solution.

  • Guarantee of Convergence: We desire theoretical assurance that the sequence of iterates \(x_k\) generated by the optimizer actually approaches a solution point \(x^\ast\).
  • Type of Solution Point:
    • Stationary Point: The optimizer converges to a point \(x^\ast\) where \(\nabla f(x^\ast) = 0\). This is a common target.
    • Local Minimum: A point \(x^\ast\) such that \(f(x^\ast) \le f(x)\) for all \(x\) in a neighborhood around \(x^\ast\). This implies \(\nabla f(x^\ast) = 0\) and \(\nabla^2 f(x^\ast)\) is positive semi-definite.
    • Global Minimum: The ultimate goal: \(f(x^\ast) \le f(x)\) for all \(x\). This is very hard for non-convex functions.

Example. Getting Stuck

Consider minimizing the function \(f(x) = x^4 - 4x^2 + x\). This function has multiple local minima (two, in fact) and one local maximum. An optimizer might converge to one of these local minima, but not necessarily the global minimum. This illustrates how non-convex functions can pose challenges.

  • Underlying Principle: The Descent Property A core mechanism for ensuring progress in many iterative minimization algorithms is the descent property: each step should decrease the objective function value, i.e., \(f(x_{k+1}) < f(x_k)\), unless \(x_k\) is already a minimizer. This is typically ensured if the search direction \(p_k\) is a descent direction (satisfying \(\nabla f(x_k)^T p_k < 0\)) and an appropriate step size \(\alpha_k > 0\) is chosen (e.g., via a line search satisfying the Wolfe conditions).

  • Theoretical Benchmark: Convergence to Stationary Points Under suitable assumptions (e.g., \(f\) is continuously differentiable, bounded below, and its gradient is Lipschitz continuous), many algorithms can be proven to converge to a point \(x^\ast\) where \(\nabla f(x^\ast) = 0\). For line search methods, this is often analyzed using Zoutendijk’s condition, which states that if \(\sum_{k=0}^\infty \cos^2 \theta_k \Vert\nabla f(x_k)\Vert^2 < \infty\), where \(\theta_k\) is the angle between the search direction \(p_k\) and the negative gradient \(-\nabla f(x_k)\). If \(\cos \theta_k\) is bounded away from zero (i.e., the search direction isn’t becoming orthogonal to the steepest descent direction), this implies \(\Vert\nabla f(x_k)\Vert \to 0\).

2.2. Quality of the Solution

Beyond just converging, the quality of the solution matters.

  • Accuracy and Precision: The optimizer should be able to find a solution \(x_k\) that is very close to a true optimum \(x^\ast\), meaning the error \(\Vert x_k - x^\ast \Vert\) or the suboptimality \(f(x_k) - f(x^\ast)\) is small.
  • ML Context: Generalization and “Good” Minima In machine learning, finding the absolute minimum of the training loss isn’t the sole aim. We desire models that generalize well to unseen data.
    • “Flat” vs. “Sharp” Minima: Optimizers that find “flatter” regions of the loss landscape (where the function value doesn’t change much with small perturbations to the parameters) are often hypothesized to lead to better generalization than those converging to “sharp” minima.

      Example. Flat vs. Sharp Minima and Generalization

      Imagine training a model. Optimizer A finds parameters \(w_A^\ast\) corresponding to a minimum where the training loss \(L(w_A^\ast)=0.01\). This minimum is very “sharp”: small changes to \(w_A^\ast\) cause the loss to increase rapidly. Optimizer B finds parameters \(w_B^\ast\) where the training loss \(L(w_B^\ast)=0.05\). This minimum is “flat”: the loss surface is almost level around \(w_B^\ast\).

      When the model is evaluated on test data, which typically differs slightly from training data, this can be seen as evaluating the loss on a slightly perturbed version of the parameters (or, equivalently, the loss function itself is slightly different for the test set).

      • For model \(A\), if the parameters are perturbed even slightly (due to the train-test shift), the sharp nature of the minimum means the loss could increase significantly (e.g., test loss becomes 0.5).
      • For model \(B\), the flatness around \(w_B^\ast\) implies that similar perturbations to the parameters result in only a small change in loss (e.g., test loss becomes 0.06).

      In this scenario, optimizer B, despite achieving a slightly higher training loss, might lead to a model that generalizes better to unseen data. Stochastic Gradient Descent (SGD), due to the noise in its gradient estimates, is often observed to find flatter minima compared to some adaptive methods that might converge more aggressively to sharper minima. The precise mechanisms and guarantees linking flatness to generalization are complex and subjects of ongoing research.

    • Implicit Regularization: Some optimizers (e.g., SGD) are thought to possess an “implicit regularization” effect due to noise or specific update rules, guiding them towards solutions that generalize better, even if they don’t achieve the absolute lowest training loss.

3. Efficiency: Speed and Resource Management

An effective optimizer must also be efficient in terms of time and computational resources.

3.1. Rate of Convergence: The Pace of Progress

How quickly does the algorithm approach the solution?

  • Theoretical Convergence Rates: These describe how fast the error \(\Vert x_k - x^\ast \Vert\) (or \(f(x_k) - f(x^\ast)\)) goes to zero as \(k \to \infty\).

    Definition. Orders of Convergence

    An iterative method \(x_{k+1} = \mathcal{U}(x_k)\) is said to converge to \(x^\ast\) with order \(p\) and rate constant \(C > 0\) if:

    \[\lim_{k \to \infty} \frac{\Vert x_{k+1} - x^\ast \Vert}{\Vert x_k - x^\ast \Vert^p} = C\]

    Key orders include:

    • Sublinear: Slower than linear (e.g., \(\Vert x_k - x^\ast \Vert \sim 1/k\) or \(1/\sqrt{k}\)). Common for some stochastic or non-smooth methods.
    • Linear (\(p=1\), \(0 < C < 1\)): The error decreases by a constant factor at each step. Example: Gradient Descent on strongly convex functions.
    • Superlinear (\(p > 1\), or \(p=1\) and \(C=0\)): Faster than linear. Example: Quasi-Newton methods.
    • Quadratic (\(p=2\)): The number of correct digits roughly doubles each iteration (once close to \(x^\ast\)). Example: Newton’s method.

    A higher theoretical rate is generally better but often comes with trade-offs.

    Example. Convergence Rates on \(f(x) = \frac{1}{2}x^2\)

    Let’s minimize the simple quadratic function \(f(x) = \frac{1}{2}x^2\), starting from an initial point \(x_0 = 4\). The true minimum is at \(x^\ast=0\).

    • Gradient Descent (GD): The update rule is \(x_{k+1} = x_k - \alpha f'(x_k) = x_k - \alpha x_k = (1-\alpha)x_k\). For this function, GD converges if \(0 < \alpha < 2\). If we choose a learning rate \(\alpha=0.5\), the iterates are: \(x_0=4, \quad x_1=(1-0.5) \cdot 4 = 2, \quad x_2=(0.5) \cdot 2 = 1, \quad x_3=(0.5) \cdot 1 = 0.5, \quad \ldots\) The error \(\vert x_k - x^\ast \vert = \vert x_k \vert\) decreases by a constant factor of \(0.5\) at each step (\(\vert x_{k+1} \vert / \vert x_k \vert = 0.5\)), demonstrating linear convergence.
    • Newton’s Method: The update rule is \(x_{k+1} = x_k - \frac{f'(x_k)}{f''(x_k)}\). Here, \(f'(x) = x\) and \(f''(x) = 1\). So, \(x_{k+1} = x_k - \frac{x_k}{1} = x_k - x_k = 0\) Starting from \(x_0 = 4\), Newton’s method converges to the minimum \(x^\ast=0\) in a single step: \(x_1 = 0\). This is characteristic of Newton’s method on quadratic functions.

    This simple example highlights the potentially faster convergence of higher-order methods, though Newton’s method requires computing and inverting the Hessian, which can be costly for general functions.

  • Practical Speed Considerations:
    • Number of iterations (\(k\)): How many steps to reach desired accuracy.
    • Number of function/gradient/Hessian evaluations: Often the bottleneck.
    • Wall-clock time: Actual runtime, influenced by per-iteration cost and hardware.
  • Analytical Tool: Behavior on Quadratic Functions Analyzing an optimizer’s performance on a simple quadratic function \(f(x) = \frac{1}{2} x^T Q x - b^T x + c\) (where \(Q\) is positive definite) provides significant insight into its behavior on more general smooth functions locally.
    • Gradient Descent exhibits linear convergence with a rate dependent on the condition number \(\kappa(Q)\).
    • Newton’s Method converges in one step.
    • Conjugate Gradient methods converge in at most \(n\) steps (for exact arithmetic).

3.2. Computational Resources and Scalability

  • Low Per-Iteration Cost (Time): Each step \(x_k \to x_{k+1}\) should be computationally inexpensive. This includes the cost of evaluating \(f(x)\), \(\nabla f(x)\), or \(\nabla^2 f(x)\), plus algorithmic overhead.
  • Low Memory Footprint (RAM): The optimizer shouldn’t require excessive memory for parameters, gradients, or internal states (e.g., Hessian approximations).
  • Scalability to High Dimensions (\(n\)): Crucial for ML, where \(n\) can be millions or billions. Per-iteration costs like \(O(n^2)\) or \(O(n^3)\) (e.g., standard Newton’s method) become prohibitive. We prefer \(O(n)\) or \(O(n \log n)\).
  • Scalability to Large Datasets (ML context): When \(f(x)\) is a sum over many data points, methods like Stochastic Gradient Descent (SGD) that use mini-batches become essential for efficiency.

4. Robustness: Navigating and Adapting to Challenges

The optimization landscape can be treacherous. A robust optimizer performs reliably across diverse and difficult conditions.

4.1. Handling Difficult Optimization Landscapes

  • Non-Convexity:
    • Escaping “Bad” Local Minima: Ability to avoid getting stuck in poor local minima.
    • Navigating Saddle Points: Efficiently escaping saddle points, which are prevalent in high-dimensional non-convex problems.

      Example. Navigating a Saddle Point

      Consider the function \(f(x,y) = \frac{1}{2}x^2 - \frac{1}{2}y^2\). The gradient is \(\nabla f(x,y) = (x, -y)^T\), and the Hessian is \(\nabla^2 f(x,y) = \begin{pmatrix} 1 & 0 \\ 0 & -1 \end{pmatrix}\). The point \((0,0)\) is a critical point since \(\nabla f(0,0) = (0,0)^T\). The Hessian at \((0,0)\) has one positive eigenvalue (1) and one negative eigenvalue (-1), confirming that \((0,0)\) is a saddle point. Along the x-axis (where \(y=0\)), the function \(f(x,0)=\frac{1}{2}x^2\) looks like a minimum. Along the y-axis (where \(x=0\)), the function \(f(0,y)=-\frac{1}{2}y^2\) looks like a maximum.

      • Gradient Descent (GD): The updates are \(x_{k+1} = x_k - \alpha x_k = (1-\alpha)x_k\) and \(y_{k+1} = y_k - \alpha (-y_k) = (1+\alpha)y_k\). If started at \((x_0, y_0)\):
        • The \(x\) component, \(x_k = (1-\alpha)^k x_0\), converges to 0 (assuming \(0 < \alpha < 2\)).
        • The \(y\) component, \(y_k = (1+\alpha)^k y_0\), diverges from 0 if \(y_0 \neq 0\) and \(\alpha > 0\). GD thus “slides off” the saddle along the direction of negative curvature (the y-axis for this function). However, if the initial component \(y_0\) along the negative curvature direction is very small, convergence towards the saddle region can be slow before the diverging component eventually pushes the iterate away. In high-dimensional problems, saddle points can have many positive and negative curvature directions, making their geometry complex.
      • Newton’s Method: Newton’s method seeks points where \(\nabla f = 0\). The update step is \(p_k = -[\nabla^2 f(x_k)]^{-1} \nabla f(x_k)\). For \(f(x,y) = \frac{1}{2}x^2 - \frac{1}{2}y^2\), this gives \(p_k = -\begin{pmatrix} 1 & 0 \\ 0 & -1 \end{pmatrix}^{-1} \begin{pmatrix} x_k \\ -y_k \end{pmatrix} = -\begin{pmatrix} 1 & 0 \\ 0 & -1 \end{pmatrix} \begin{pmatrix} x_k \\ -y_k \end{pmatrix} = \begin{pmatrix} -x_k \\ -y_k \end{pmatrix}\). So, the next iterate is \(x_{k+1} = x_k + p_{k,x} = x_k - x_k = 0\) and \(y_{k+1} = y_k + p_{k,y} = y_k - y_k = 0\). From any starting point \((x_k, y_k)\), a pure Newton step jumps directly to the saddle point \((0,0)\). This occurs because Newton’s method targets any stationary point where \(\nabla f=0\). To specifically seek minima and effectively escape saddles (i.e., move towards points with lower function values), Newton-type methods often incorporate trust regions or modify the Hessian (e.g., by adding a multiple of the identity matrix if it’s not positive definite) to ensure descent.

      In practice, for escaping saddle points and finding local minima:

      • Stochastic Gradient Descent (SGD): The noise introduced by mini-batch gradients can provide the necessary “perturbation” to move the iterates away from saddle points, especially along directions where the gradient is small or corresponds to negative curvature.
      • Second-order methods with modifications: Algorithms like trust-region Newton methods or Cubic Regularized Newton’s method are designed to detect and exploit negative curvature directions to efficiently escape saddles and proceed towards minima.
  • Ill-Conditioning:
    • The problem is ill-conditioned if the Hessian \(\nabla^2 f(x)\) has a high condition number (level sets are like elongated ellipses).
    • Optimizers should handle this gracefully, avoiding excessive zig-zagging or slow convergence. Preconditioning or adaptive scaling helps.

      Example. Impact of Ill-Conditioning

      Consider minimizing the ill-conditioned quadratic function \(f(x,y) = \frac{1}{2}(x^2 + 100y^2)\). The minimum is at \((0,0)\). The Hessian matrix is \(\nabla^2 f(x,y) = \begin{pmatrix} 1 & 0 \\ 0 & 100 \end{pmatrix}\), and its condition number (ratio of largest to smallest eigenvalue) is \(\kappa = 100/1 = 100\). The level sets of this function are highly elongated ellipses. The gradient is \(\nabla f(x,y) = (x, 100y)^T\). The Gradient Descent update is: \(\begin{pmatrix} x_{k+1} \\ y_{k+1} \end{pmatrix} = \begin{pmatrix} x_k \\ y_k \end{pmatrix} - \alpha \begin{pmatrix} x_k \\ 100y_k \end{pmatrix} = \begin{pmatrix} (1-\alpha)x_k \\ (1-100\alpha)y_k \end{pmatrix}\) For stability, particularly for the \(y\) component, the learning rate \(\alpha\) must satisfy \(\vert 1-100\alpha \vert < 1\), which means \(0 < 100\alpha < 2\), or \(0 < \alpha < 0.02\).

      • If we choose \(\alpha\) close to the upper stability limit, say \(\alpha = 0.019\), then the convergence factor for \(x\) is \((1-0.019) = 0.981\) (slow), while for \(y\) it’s \(\vert 1-100 \cdot 0.019 \vert = \vert 1-1.9 \vert = 0.9\) (faster, but oscillating as \(1-1.9 = -0.9\)).
      • If we pick a smaller \(\alpha\) to ensure smooth convergence for \(y\), e.g., \(\alpha=0.01\), then \(x_{k+1} = 0.99 x_k\) (very slow convergence in the \(x\) direction) and \(y_{k+1} = (1-1)y_k = 0 \cdot y_k\) (immediate convergence in the \(y\) direction if \(y_0 \ne 0\)). This disparity in required learning rates across different directions forces Gradient Descent to take many small steps, effectively “zig-zagging” down the narrow valley of the objective function, leading to slow overall convergence. Newton’s method, being affine invariant (as discussed in Section 4.2), would rescale the problem variables appropriately and converge in one step for this quadratic. Adaptive methods (like AdaGrad, Adam) also aim to mitigate this issue by effectively using different step sizes for different parameters, thereby handling the disparate scales much better.

      </blockquote>

4.2. Fundamental Mathematical Robustness: Invariance Properties

These describe how an optimizer’s behavior changes (or, ideally, doesn’t) under transformations of the problem’s coordinate system or parameterization.

  • Affine Invariance:
    • Definition: An optimizer is affine invariant if its iterates transform covariantly under an affine transformation \(x = Ay + b\) (with \(A\) invertible). If \(x_0, x_1, \ldots\) are iterates for \(f(x)\), and \(y_0\) corresponds to \(x_0\), then the iterates \(y_k\) for \(g(y) = f(Ay+b)\) satisfy \(x_k = Ay_k + b\). The optimization path is geometrically equivalent in the transformed space.
    • Importance: The optimizer’s performance is independent of linear correlations or scaling of variables. This reduces the need for manual preconditioning.
    • Example: Newton’s Method. The pure Newton step \(x_{k+1} = x_k - [\nabla^2 f(x_k)]^{-1} \nabla f(x_k)\) is affine invariant.

      Proof Sketch. Newton’s Method Affine Invariance

      Let \(x = Ay+b\). The objective function in terms of \(y\) is \(g(y) = f(Ay+b)\). The gradients and Hessians are related by:

      \[\nabla_y g(y) = A^T \nabla_x f(Ay+b)\] \[\nabla_y^2 g(y) = A^T \nabla_x^2 f(Ay+b) A\]

      The Newton step for \(y\) is \(y_{k+1} = y_k - [\nabla_y^2 g(y_k)]^{-1} \nabla_y g(y_k)\). Substituting the transformations:

      \[y_{k+1} = y_k - [A^T \nabla_x^2 f(x_k) A]^{-1} A^T \nabla_x f(x_k)\] \[y_{k+1} = y_k - A^{-1} [\nabla_x^2 f(x_k)]^{-1} (A^T)^{-1} A^T \nabla_x f(x_k)\] \[y_{k+1} = y_k - A^{-1} ([\nabla_x^2 f(x_k)]^{-1} \nabla_x f(x_k))\]

      Multiplying by \(A\) and adding \(b\):

      \[Ay_{k+1} + b = Ay_k + b - ([\nabla_x^2 f(x_k)]^{-1} \nabla_x f(x_k))\]

      If \(x_k = Ay_k+b\), then \(x_{k+1} = Ay_{k+1}+b\). The sequence of iterates \(x_k\) is generated by the same Newton update rule, regardless of the affine transformation.

    • Gradient Descent is generally not affine invariant.
  • Scaling Invariance (Diagonal Scaling):
    • Definition: A special case of affine invariance where \(A\) is a diagonal matrix \(D\) (and \(b=0\)), so \(x = Dy\). The optimizer behaves consistently if individual variables are rescaled (e.g., changing units).
    • Importance: Crucial when variables have vastly different scales. Avoids the need for manual normalization.
    • Examples:
      • Newton’s method (being affine invariant).
      • Adaptive methods like AdaGrad, RMSProp, Adam achieve a form of diagonal scaling invariance by adapting per-parameter learning rates. For AdaGrad, the update for \(x_i\) is \(x_{i, t+1} = x_{i, t} - \frac{\eta}{\sqrt{G_{ii,t} + \epsilon}} g_{i,t}\). If \(x_i\) is scaled by \(s_i\), its gradient \(g_i\) scales by \(1/s_i\), \(G_{ii,t}\) by \(1/s_i^2\), so \(\frac{g_{i,t}}{\sqrt{G_{ii,t}}}\) is scale-invariant. Thus, the update \(\Delta x_i\) is scale-invariant w.r.t. \(x_i\).
  • Isometry Invariance (Rotations and Translations):
    • Definition: Behavior is unchanged if coordinates are rotated (\(x' = Rx + t\), \(R^TR=I\)) or translated.
    • Importance: Robustness to arbitrary choice of coordinate orientation or origin.
    • Examples: Newton’s method (implied by affine invariance). Gradient Descent with an invariant line search rule.

Invariance properties significantly enhance an optimizer’s robustness and reduce its sensitivity to problem formulation.

4.3. Resilience to Imperfections

  • Handling Noise and Stochasticity: If function or gradient evaluations are noisy (e.g., SGD using mini-batch gradients, simulation-based optimization), the optimizer should still converge robustly.
  • Insensitivity to Initialization: Performance should be consistent across different starting points \(x_0\). High sensitivity may lead to divergence or convergence to poor local minima if not initialized carefully.

4.4. Hyperparameter Management and Numerical Stability

  • Sensitivity, Number, and Tuning of Hyperparameters:
    • Few Hyperparameters: Simpler to use.
    • Robustness to Settings: Good performance across a reasonable range of hyperparameter values.
    • Clear Tuning Guidance: Heuristics or principles for setting them.
    • Adaptive/Parameter-Free Nature: The ideal is minimal or no tuning (e.g., adaptive learning rates).
  • Numerical Stability: Updates should not lead to numerical overflow or underflow (NaNs).

Tip. The Learning Rate Dilemma

The learning rate is often the most critical hyperparameter.

  • Too small: Slow convergence.
  • Too large: Overshooting, oscillation, or divergence. Optimizers less sensitive to this, or that adapt it, are highly valued.

An optimizer that performs well across a wide range of problems (convex, non-convex, smooth, mildly non-smooth) without requiring bespoke tuning for each is a hallmark of robustness.

5. Practicality and Broader Applicability

Theoretical excellence must be complemented by practical usability.

5.1. Ease of Implementation and Use

  • Simplicity of Algorithm: Conceptually simple algorithms are easier to implement, debug, and understand.
  • Availability in Libraries: Ready availability in standard ML frameworks (PyTorch, TensorFlow, JAX, scikit-learn) promotes adoption.

5.2. Derivative Requirements

  • Zeroth-Order (Derivative-Free): Use only \(f(x)\) values. Broadest applicability (non-smooth, black-box) but often slower.
  • First-Order (Gradient-Based): Use \(\nabla f(x)\). Common in ML (e.g., GD, SGD, Adam).
  • Second-Order: Use \(\nabla^2 f(x)\) (Hessian). Can be very fast (e.g., Newton) but expensive for large \(n\). The fewer derivatives needed, the wider the range of applicable problems, often at a cost to convergence speed or robustness on ill-conditioned problems.

5.3. Parallelism and Distributed Computation

  • Parallelizability: Optimizer computations (especially gradient calculation) should be parallelizable across cores/devices.
  • Communication Efficiency: In distributed settings, minimizing communication overhead is key for scaling.

5.4. (Desirable) Handling Constraints

Many real-world optimization problems involve constraints on \(x\) (e.g., \(g_i(x) \le 0\), \(h_j(x) = 0\)). While many ML optimizers focus on unconstrained problems (constraints often handled via regularization), general-purpose optimizers benefit from mechanisms to handle constraints (e.g., interior-point methods, augmented Lagrangians). This is often a more advanced topic.

6. The Optimizer’s Balancing Act: Trade-offs and No Silver Bullet

It’s crucial to understand that no single optimizer excels in all these desirable properties simultaneously. There are inherent trade-offs:

  • Rate vs. Cost: Methods with faster theoretical convergence rates (e.g., Newton’s quadratic rate) often have higher per-iteration computational costs and greater derivative requirements than simpler methods (e.g., Gradient Descent’s linear rate).
  • Global vs. Local: Algorithms that aim for global optima are typically much more computationally intensive and slower than those designed to find local optima.
  • Robustness vs. Peak Speed: Highly robust algorithms might be slightly slower on very well-behaved, simple problems compared to a specialized, finely-tuned algorithm.
  • Adaptivity vs. Optimality: Adaptive methods (like Adam) offer ease of use and robustness across many problems but might sometimes be outperformed in terms of final solution quality or generalization by simpler methods like SGD with meticulously tuned schedules on specific tasks.

The choice of an optimizer is an art informed by science, involving:

  • Understanding the characteristics of your problem (convexity, smoothness, size, noise).
  • Knowing the strengths, weaknesses, and theoretical underpinnings of different optimizers.
  • Considering available computational resources.
  • Empirical experimentation and validation.

The gap between theoretical guarantees (often based on simplifying assumptions) and practical performance on complex, high-dimensional, non-convex ML problems is an active research area.

7. Summary

We’ve detailed a “wish list” for optimizers, categorized as:

  • Core Effectiveness: Converging to high-quality solutions that (in ML) generalize well. This relies on properties like guaranteed convergence and the descent property.
  • Efficiency: Achieving solutions quickly and with minimal computational resources, characterized by convergence rates and scalability.
  • Robustness: Performing reliably across diverse and challenging scenarios, aided by mathematical invariances (affine, scaling), resilience to noise, good landscape navigation, and stable hyperparameter behavior.
  • Practicality & Broader Applicability: Being easy to implement and use, having manageable derivative requirements, and scaling in distributed environments.

Understanding these properties and their inherent trade-offs is key to selecting and developing effective optimization strategies.

8. Cheat Sheet: Desirable Optimizer Properties at a Glance

CategoryKey Desirable CharacteristicsImportance & Examples
Core EffectivenessGuaranteed convergence (to good minima), Accuracy/Precision, Good generalization (ML).Reliable training, high-quality solutions. Principles: Descent property.
EfficiencyFast theoretical rate (Linear, Superlinear, Quadratic), Low per-iteration cost (time/memory), Scalability (high-dim, large datasets).Reduced training time, feasible for large problems. Analysis: Behavior on quadratics.
Robustness: LandscapeHandles non-convexity (saddles, local minima), ill-conditioning.Works on complex, real-world problems. Techniques: Preconditioning, adaptive steps.
Robustness: InvarianceAffine invariance, Scaling invariance, Isometry invariance.Performance independent of problem parameterization/scaling. Examples: Newton (affine), Adam (scaling).
Robustness: OtherHandles noise/stochasticity, Insensitive to initialization, Stable numerics, Manageable hyperparameters.Reliable performance without excessive tuning or failure. Key for SGD, robust to starting points.
Practicality & ApplicabilityEasy to implement/use, Appropriate derivative requirements (0th, 1st, 2nd order), Parallelizable, Handles constraints (general).Accessible, applicable to diverse problems, efficient use of modern hardware.

9. Reflection

This exploration of desirable optimizer properties provides a critical lens for evaluating existing algorithms and a roadmap for future innovations. As we delve into specific optimizers in subsequent posts—from Gradient Descent to Adam and beyond—we will continuously refer back to this framework. We’ll analyze how each method strives to embody these properties, understand its strengths and weaknesses, and appreciate the clever design choices that have propelled progress in machine learning and optimization.

The quest for the “perfect” optimizer continues, driven by the ever-increasing complexity of models and datasets. By understanding what constitutes “good”, we are better equipped to navigate this evolving landscape. Our next post will offer a “Speedrun of common gradient-based ML optimizers”, providing a first look at how different algorithms put these principles into practice.


Further Reading and References

  • Nocedal, J., & Wright, S. J. (2006). Numerical Optimization. Springer.
  • Goodfellow, I., Bengio, Y., & Courville, A. (2016). Deep Learning. MIT Press.
  • Bottou, L., Curtis, F. E., & Nocedal, J. (2018). Optimization methods for large-scale machine learning. SIAM Review, 60(2), 223-311.
  • Ruder, S. (2016). An overview of gradient descent optimization algorithms. arXiv preprint arXiv:1609.04747.
  • Boyd, S., & Vandenberghe, L. (2004). Convex Optimization. Cambridge University Press.
This post is licensed under CC BY 4.0 by the author.