The Derivative (and why it matters)


“Whatcha doing George?” “Oh nothing Lenny, just working out some gradients.” “On paper? I’m not sure if you’ll be able to call loss.backward() there.


Machine learning, especially deep learning, is built almost entirely on differentiation. In this post, we will give a short introduction to differentiation, from derivatives, to gradients and Jacobians. From there, we will discuss their connection to the underlying idea behind many popular deep learning frameworks: automatic differentiation.

Open In Colab

Derivatives

While derivatives can be explained and generalized to quite a few domains, we will start with the most elementary formulation: in the case of a function of a single, real-valued variable.

Concretely, a derivative in this setting measures the sensitivity of the output with respect to a changes in input: graphically in one dimension, this is the slope of the tangent line at . In physics, we refer to changes in position (output) with respect to time (input) as velocity, and changes in velocity with respect to time, acceleration.

Derivatives can be generalized to both higher orders and higher dimensions. Higher orders of derivatives are just differentiating with respect to the input once again; for example, the second order derivative of position with respect to time brings us back to acceleration.

To understand higher dimensional derivatives, we can introduce partial derivatives. When given a vector-valued input , a partial derivative is just a derivative of the output with respect to a particular component of . When dealing with scalar-valued functions (i.e whose output is scalar, or notationally, ), the vector of first order partial derivatives is called the gradient. If the output itself is vector-valued, the matrix of first-order partial derivatives is called the Jacobian; in fact, the Jacobian is just a generalization of the gradient.

Just as the derivative of a real-valued function is the slope of the tangent line, the gradient’s direction is the direction of greatest increase, where the rate of that increase is given by the magnitude. Often, we will use the notation to denote the gradient.

While higher-order derivatives of scalar-valued functions are sometimes used in machine learning (the matrix of second-order derivatives of a vector-valued function is called the Hessian, and when the derivatives are taken with respect to parameters instead of the inputs, the matrix is called the Fisher Information Matrix), they deserve their own treatment, and are not central to discussion here.

Machine Learning and Gradients

In machine learning, we care about objective functions and how to optimize them, whether we are working with supervised, deep, reinforcement, … learning. To optimize a function, we search for optima, or where and the second-derivative test has certain criteria.

While generally, we cannot make the assumption our objective function is convex (which means that we cannot guarantee any local optima are globally optimal), we can still use gradients to find local minima (or maxima, depending on the problem formulation).

This gives rise to algorithms like gradient descent (or more popularly, its stochastic cousin, SGD); since we know the gradient gives us the direction of fastest increase, we can take the opposite direction to minimize an objective.

While blindly following first order gradients can get also us into trouble (saddles also have just like optima, and unfortunately are ubiquitous in nonconvex optimization; while SGD is generally considered effective at escaping these regions, the dynamics of gradient descent algorithms is still an active area of research), we defer these discussions to a later post.

In deep learning we are often concerned with continuous, differentiable functions. It is important to remember that many optimization problems are non-differentiable, but we can still perform some form of hill climbing, or iterated local search procedure. Below is a naïve implementation of a hill climbing algorithm which tries perturbing each input by a positive, negative or zero value, and returns the input corresponding to the greatest response.

fun search(P̂: ℝᵐ→ℝ, c: ℝ, g₀: ℝ, g₁: ℝ, …, gₘ₋₁: ℝ): ℝᵐ
    if m = 1:
        return argmax { P̂(g₀+c), P̂(g₀-c), P̂(g₀) }
    
    return argmax { P̂(g₀+c)◦search(P̂(g₀+c), c, g₁, …, gₘ₋₁),
                    P̂(g₀-c)◦search(P̂(g₀-c), c, g₁, …, gₘ₋₁),
                    P̂(g₀)◦search(P̂(g₀), c, g₁, …, gₘ₋₁) }

fun climb(P̂: ℝᵐ→ℝ, c: ℝ, g: ℝᵐ): ℝᵐ
    Δ ← check(P̂, c, g)

    if Δ = g:
        return g

    return climb(P̂, c, Δ)

The search algorithm is where is the input dimensionality. If is differentiable, this algorithm is not particularly efficient, but offers a framework upon which we can make various improvements by assuming further structure on .

Calculating Gradients, Numerically

To understand the numerical calculation of gradients, we must refer back to the definition of derivatives, in terms of limits:

However, since limits are a theoretical tool, they offer little help in the practical calculation of gradients.

Numerically, we can replace the infinitesimal with an arbitrarily small one, leading us to approximate the derivative as:

This is known as a first-order divided difference.

Numerical differentiation (known as differential quadrature) are still the leading methods to solve partial differential equations, however they incur errors on an order of magnitude of . First-order divided differences incur errors on the order of , and while improved methods such as symmetric divided difference can lower this, differential quadrature methods can be prone to floating point approximation errors and implementation efficiency issues.

Calculating Gradients, Analytically

``In a recent article R. Wengert suggested a technique for machine evaluation of the partial derivatives of a function given in analytical form [that] appears very attractive from the programming viewpoint and permits the treatment of large systems of differential equations which might not otherwise be undertaken.’’

–Richard E. Bellman (1964) 1

Numerical approaches tell us the approximate sensitivity of a function with respect to its inputs at a single point. Symbolic differentiation gives us a closed-form expression for the derivative at any point in its input.

In calculus, we are taught many rules for symbolic differentiation. These rules are convenient identities to remember, but almost all differential calculus can be recovered from three simple rules for scalar differentiation:

The same notion which appears in the differential calculi can be applied in many different contexts from regular expressions to λ-calculus to linear logic. In machine learning, we are chiefly interested in calculating derivatives for vector fields or some mechanical representation thereof. As long as a number system admits the standard arithmetic notation (addition, multiplication and their inverses), it is possible to symbolically derive an expression which, when evaluated at a point, will equate to the finite difference approximation.

We make the following two claims:

  1. Symbolic differentiation can be performed mechanically by replacing the expressions on the left hand side with their right hand side equivalents. This process often requires less computation than the numerical method described above.
  2. The same rules can be applied to functions whose inputs and outputs are vectors, with exactly the same algebraic semantics. Using matrix notation requires far less computation than elementwise scalar differentiation.

Firstly, let us examine the first claim. A naïve implementation of the finite difference method requires at least two evaluations each time we wish to compute the derivative at a certain input. While algebraic rewriting can help to reduce the computation, but we are still left with the . It is often more efficient to derive a closed form analytical solution for the derivative at every input.

Secondly, partial differentiation of vector functions is a specific case of higher dimensional derivatives that are often more convenient to represent as a matrix, or Jacobian, which is defined as follows:

The matrix representation often requires far less memory and computation than a naïve implementation which iteratively computes derivatives for each element of a vector function.

TODO: give an example

The Chain Rule, Reexamined

TODO: Leibniz notation shows us something very interesting. What happens if we take the derivative with respect to a function?

Sum of the Parts are Greater than the Whole

TODO: A section on how simple arithmetic operations and simple functions can be composed to efficiently calculate derivatives. Composition of linear maps is a linear map(!)

From Analytical to Automatic Differentiation

TODO: Examine how to implement these rules using operator overloading, computation graphs and recursion.

Forward Mode and Backward Mode

TODO: Introduce linear function chains, optimal Jacobian accumulation problem and some heuristic solutions. Demonstrate algorithmic complexity of forward and reverse accumulation where and vis versa.

Suppose we have a function , where and . Using the chain rule for vector functions, the Jacobian can be defined as:

Recall that matrix multiplication is associative . In which order should we evaluate this product in order to minimize the computational complexity? We consider two cases, where , and .

Linear Regression from an AD Perspective

Recall the matrix equation for linear regression, where and :

Imagine we are given the following dataset:

Our goal in ordinary least squares (OLS) linear regression is to minimize the loss, or error between the data and the model’s prediction:

Finite Difference Method

First, we consider the scalar case, where . Since are considered to be fixed, we can rewrite as simply:

To find the minimizer of , we need . There are various ways to compute this, of which we consider two: (1) the finite difference method (FDM), and (2) symbolic differentiation. First, let’s see FDM with centered differences:

Using computer algebra, these equations can be simplified considerably:

Partial Differentiation

Alternatively, we can calculate the partials analytically, by applying the chain rule:

Notice how analytical differentiation gives us the same answer as the finite difference method (this is not by accident), with much less algebra. We can rewrite these two solutions in gradient form, i.e. as a column vector of partial derivatives:

Matrix solutions

Having reviewed the scalar procedure for linear regression, let us now return to the general form of . Matrix notation allows us to simplify the loss considerably:

Matrix notation allows us to derive the gradient and requires far less algebra:

For completeness, and to convince ourselves the matrix solution is indeed the same:

Notice how we recover the same solution obtained from partial differentiation and finite difference approximation, albeit in a more compact form. For a good introduction to matrix calculus, the textbook by Magnus & Neudecker (2007) 2 is an excellent guide, of which Petersen & Pedersen (2012) 3 and Parr & Howard (2018) 4 offer a review of important identities.

Unlike most optimization problems in machine learning, OLS linear regression is a convex optimization problem. If is invertible, i.e. full-rank, this implies a unique solution , which we can solve for directly by setting :

Solving this requires computing which is at least 5 to the best of our knowledge, i.e. quadratic with respect to the number of input dimensions. Another way to find is by initializing and repeating the following procedure until convergence:

Typically, . Although hyperparameter tuning is required to find a suitable (various improvements like Nesterov momentum and quasi-Newton methods also help to accelerate convergence), this procedure is guaranteed to be computationally more efficient than matrix inversion for sufficiently large and . In practice, the normal equation is seldom used unless is very small.