Book Image

Mastering Machine Learning Algorithms

Book Image

Mastering Machine Learning Algorithms

Overview of this book

Machine learning is a subset of AI that aims to make modern-day computer systems smarter and more intelligent. The real power of machine learning resides in its algorithms, which make even the most difficult things capable of being handled by machines. However, with the advancement in the technology and requirements of data, machines will have to be smarter than they are today to meet the overwhelming data needs; mastering these algorithms and using them optimally is the need of the hour. Mastering Machine Learning Algorithms is your complete guide to quickly getting to grips with popular machine learning algorithms. You will be introduced to the most widely used algorithms in supervised, unsupervised, and semi-supervised machine learning, and will learn how to use them in the best possible manner. Ranging from Bayesian models to the MCMC algorithm to Hidden Markov models, this book will teach you how to extract features from your dataset and perform dimensionality reduction by making use of Python-based libraries such as scikit-learn v0.19.1. You will also learn how to use Keras and TensorFlow 1.x to train effective neural networks. If you are looking for a single resource to study, implement, and solve end-to-end machine learning problems and use-cases, this is the book you need.
Table of Contents (22 chapters)
Title Page
Dedication
Packt Upsell
Contributors
Preface
13
Deep Belief Networks
Index

Features of a machine learning model


In this section, we're going to consider supervised models, and try to determine how it's possible to measure their theoretical potential accuracy and their ability to generalize correctly over all possible samples drawn from pdata. The majority of these concepts were developed before the deep learning age, but continue to have an enormous influence on research projects. The idea of capacity, for example, is an open-ended question that neuroscientists keep on asking themselves about the human brain. Modern deep learning models with dozens of layers and millions of parameters reopened the theoretical question from a mathematical viewpoint. Together with this, other elements, like the limits for the variance of an estimator, again attracted the limelight because the algorithms are becoming more and more powerful, and performances that once were considered far from any feasible solution are now a reality. Being able to train a model, so as to exploit its full capacity, maximize its generalization ability, and increase the accuracy, overcoming even human performances, is what a deep learning engineer nowadays has to expect from his work.

Capacity of a model

If we consider a supervised model as a set of parameterized functions, we can define representational capacity as the intrinsic ability of a certain generic function to map a relatively large number of data distributions. To understand this concept, let's consider a function f(x) that admits infinite derivatives, and rewrite it as a Taylor expansion:

We can decide to take only the first n terms, so to have an n-degree polynomial function. Consider a simple bi-dimensional scenario with six functions (starting from a linear one); we can observe the different behavior with a small set of data points:

Different behavior produced by six polynomial separating curves

The ability to rapidly change the curvature is proportional to the degree. If we choose a linear classifier, we can only modify its slope (the example is always in a bi-dimensional space) and the intercept. Instead, if we pick a higher-degree function, we have more possibilities to bend the curvature when it's necessary. If we consider n=1 and n=2 in the plot (on the top-right, they are the first and the second functions), with n=1, we can include the dot corresponding to x=11, but this choice has a negative impact on the dot at x=5.

Only a parameterized non-linear function can solve this problem efficiently, because this simple problem requires a representational capacity higher than the one provided by linear classifiers. Another classical example is the XOR function. For a long time, several researchers opposed perceptrons (linear neural networks), because they weren't able to classify a dataset generated by the XOR function. Fortunately, the introduction of multilayer perceptrons, with non-linear functions, allowed us to overcome this problem, and many whose complexity is beyond the possibilities of any classic machine learning model.

Vapnik-Chervonenkis capacity

 A mathematical formalization of the capacity of a classifier is provided by the Vapnik-Chervonenkis theory. To introduce the definition, it's first necessary to define the concept of shattering. If we have a class of sets C and a set M, we say that C shatters M if:

In other words, given any subset of M, it can be obtained as the intersection of a particular instance of C (cj) and M itself. If we now consider a model as a parameterized function:

We want to determine its capacity in relation to a finite dataset X:

According to the Vapnik-Chervonenkis theory, we can say that the model f shatters X if there are no classification errors for every possible label assignment. Therefore, we can define the Vapnik-Chervonenkis-capacity or VC-capacity (sometimes called VC-dimension) as the maximum cardinality of a subset of X so that f can shatter it.

For example, if we consider a linear classifier in a bi-dimensional space, the VC-capacity is equal to 3, because it's always possible to label three samples so that f shatters them; however, it's impossible to do it in all situations where N > 3. The XOR problem is an example that needs a VC-capacity higher than 3. Let's explore the following plot:

XOR problem with different separating curves

This particular label choice makes the set non-linearly separable. The only way to overcome this problem is to use higher-order functions (or non-linear ones). The curve lines (belonging to a classifier whose VC-capacity is greater than 3) can separate both the upper-left and the lower-right regions from the remaining space, but no straight line can do the same (while it can always separate one point from the other three).

Bias of an estimator

Let's now consider a parameterized model with a single vectorial parameter (this isn't a limitation, but only a didactic choice):

The goal of a learning process is to estimate the parameter θ so as, for example, to maximize the accuracy of a classification. We define the bias of an estimator (in relation to a parameter θ):

In other words, the bias is the difference between the expected value of the estimation and the real parameter value. Remember that the estimation is a function of X, and cannot be considered a constant in the sum.

An estimator is said to be unbiased if:

Moreover, the estimator is defined as consistent if the sequence of estimations converges (at least with probability 1) to the real value when k → ∞:

Given a dataset X whose samples are drawn from pdata, the accuracy of an estimator is inversely proportional to its bias. Low-bias (or unbiased) estimators are able to map the dataset X with high-precision levels, while high-bias estimators are very likely to have a capacity that isn't high enough for the problem to solve, and therefore their ability to detect the whole dynamic is poor. 

Let's now compute the derivative of the bias with respect to the vector θ (it will be useful later):

Consider that the last equation, thanks to the linearity of E[•], holds also if we add a term that doesn't depend on x to the estimation of θ. In fact, in line with the laws of probability, it's easy to verify that:

Underfitting

A model with a high bias is likely to underfit the training set. Let's consider the scenario shown in the following graph:

Underfitted classifier: The curve cannot separate correctly the two classes

Even if the problem is very hard, we could try to adopt a linear model and, at the end of the training process, the slope and the intercept of the separating line are about -1 and 0 (as shown in the plot); however, if we measure the accuracy, we discover that it's close to 0! Independently from the number of iterations, this model will never be able to learn the association between X and Y. This condition is called underfitting, and its major indicator is a very low training accuracy. Even if some data preprocessing steps can improve the accuracy, when a model is underfitted, the only valid solution is to adopt a higher-capacity model.

In a machine learning task, our goal is to achieve the maximum accuracy, starting from the training set and then moving on to the validation set. More formally, we can say that we want to improve our models so to get as close as possible to Bayes accuracy. This is not a well-defined value, but a theoretical upper limit that is possible to achieve using an estimator. In the following diagram, we see a representation of this process:

Accuracy level diagram

Bayes accuracy is often a purely theoretical limit and, for many tasks, it's almost impossible to achieve using even biological systems; however, advancements in the field of deep learning allow to create models that have a target accuracy slightly below the Bayes one. In general, there's no closed form for determining the Bayes accuracy, therefore human abilities are considered as a benchmark. In the previous classification example, a human being is immediately able to distinguish among different dot classes, but the problem can be very hard for a limited-capacity classifier. Some of the models we're going to discuss can solve this problem with a very high target accuracy, but at this point, we run another risk that can be understood after defining the concept of variance of an estimator.

Variance of an estimator

At the beginning of this chapter, we have defined the data generating process pdata, and we have assumed that our dataset X has been drawn from this distribution; however, we don't want to learn existing relationships limited to X, but we expect our model to be able to generalize correctly to any other subset drawn from pdata. A good measure of this ability is provided by the variance of the estimator:

The variance can be also defined as the square of the standard error (analogously to the standard deviation). A high variance implies dramatic changes in the accuracy when new subsets are selected, because the model has probably reached a very high training accuracy through an over-learning of a limited set of relationships, and it has almost completely lost its ability to generalize. 

Overfitting

If underfitting was the consequence of a low capacity and a high bias, overfitting is a phenomenon that a high variance can detect. In general, we can observe a very high training accuracy (even close to the Bayes level), but not a poor validation accuracy. This means that the capacity of the model is high enough or even excessive for the task (the higher the capacity, the higher the probability of large variances), and that the training set isn't a good representation of pdata. To understand the problem, consider the following classification scenarios:

Acceptable fitting (left), overfitted classifier (right)

The left plot has been obtained using logistic regression, while, for the right one, the algorithm is SVM with a sixth-degree polynomial kernel. If we consider the second model, the decision boundaries seem much more precise, with some samples just over them. Considering the shapes of the two subsets, it would be possible to say that a non-linear SVM can better capture the dynamics; however, if we sample another dataset from pdata and the diagonal tail becomes wider, logistic regression continues to classify the points correctly, while the SVM accuracy decreases dramatically. The second model is very likely to be overfitted, and some corrections are necessary. When the validation accuracy is much lower than the training one, a good strategy is to increase the number of training samples to consider the real pdata. In fact, it can happen that a training set is built starting from a hypothetical distribution that doesn't reflect the real one; or the number of samples used for the validation is too high, reducing the amount of information carried by the remaining samples. Cross-validation is a good way to assess the quality of datasets, but it can always happen that we find completely new subsets (for example, generated when the application is deployed in a production environment) that are misclassified, even if they were supposed to belong to pdata. If it's not possible to enlarge the training set, data augmentation could be a valid solution, because it allows creating artificial samples (for images, it's possible to mirror, rotate, or blur them) starting from the information stored in the known ones. Other strategies to prevent overfitting are based on a technique called regularization, which we're going to discuss in the last part of this chapter. For now, we can say that the effect of regularization is similar to a partial linearization, which implies a capacity reduction with a consequent variance decrease.

The Cramér-Rao bound

If it's theoretically possible to create an unbiased model (even asymptotically), this is not true for variance. To understand this concept, it's necessary to introduce an important definition: the Fisher information. If we have a parameterized model and a data-generating process pdata, we can define the likelihood function by considering the following parameters:

This function allows measuring how well the model describes the original data generating process. The shape of the likelihood can vary substantially, from well-defined, peaked curves, to almost flat surfaces. Let's consider the following graph, showing two examples based on a single parameter:

Very peaked likelihood (left), flatter likelihood (right)

We can immediately understand that, in the first case, the maximum likelihood can be easily reached by gradient ascent, because the surface is very peaked. In the second case, instead, the gradient magnitude is smaller, and it's rather easy to stop before reaching the actual maximum because of numerical imprecisions or tolerances. In worst cases, the surface can be almost flat in very large regions, with a corresponding gradient close to zero. Of course, we'd like to always work with very sharp and peaked likelihood functions, because they carry more information about their maximum. More formally, the Fisher information quantifies this value. For a single parameter, it is defined as follows:

The Fisher information is an unbounded non-negative number that is proportional to the amount of information carried by the log-likelihood; the use of logarithm has no impact on the gradient ascent, but it simplifies complex expressions by turning products into sums. This value can be interpreted as the speed of the gradient when the function is reaching the maximum; therefore, higher values imply better approximations, while a hypothetical value of zero means that the probability to determine the right parameter estimation is also null.

When working with a set of K parameters, the Fisher information becomes a positive semidefinite matrix:

This matrix is symmetric, and also has another important property: when a value is zero, it means that the corresponding couple of parameters are orthogonal for the purpose of the maximum likelihood estimation, and they can be considered separately. In many real cases, if a value is close to zero, it determines a very low correlation between parameters and, even if it's not mathematically rigorous, it's possible to decouple them anyway.

At this point, it's possible to introduce the Cramér-Rao bound, which states that for every unbiased estimator that adopts x (with probability distribution p(x; θ)) as a measure set, the variance of any estimator of θ is always lower-bounded according to the following inequality:

In fact, considering initially a generic estimator and exploiting Cauchy-Schwarz inequality with the variance and the Fisher information (which are both expressed as expected values), we obtain:

Now, if we use the expression for derivatives of the bias with respect to θ, considering that the expected value of the estimation of θ doesn't depend on x, we can rewrite the right side of the inequality as:

If the estimator is unbiased, the derivative on the right side is equal to zero, therefore, we get:

In other words, we can try to reduce the variance, but it will be always lower-bounded by the inverse Fisher information. Therefore, given a dataset and a model, there's always a limit to the ability to generalize. In some cases, this measure is easy to determine; however, its real value is theoretical, because it provides the likelihood function with another fundamental property: it carries all the information needed to estimate the worst case for variance. This is not surprising: when we discussed the capacity of a model, we saw how different functions could drive to higher or lower accuracies. If the training accuracy is high enough, this means that the capacity is appropriate or even excessive for the problem; however, we haven't considered the role of the likelihood p(X| θ).

High-capacity models, in particular, with small or low-informative datasets, can drive to flat likelihood surfaces with a higher probability than lower-capacity models. Therefore, the Fisher information tends to become smaller, because there are more and more parameter sets that yield similar probabilities, and this, at the end of the day, drives to higher variances and an increased risk of overfitting. To conclude this section, it's useful to consider a general empirical rule derived from the Occam's razor principle: whenever a simpler model can explain a phenomenon with enough accuracy, it doesn't make sense to increase its capacity. A simpler model is always preferable (when the performance is good and it represents accurately the specific problem), because it's normally faster both in the training and in the inference phases, and more efficient. When talking about deep neural networks, this principle can be applied in a more precise way, because it's easier to increase or decrease the number of layers and neurons until the desired accuracy has been achieved.