Principal component analysis, a minimal fluff tutorial

PCA explained with less words (and hopefully with more depth)

Goal of this blog post

The problem with most PCA explanations is that they are too long and use too many words (I typed this section before, the rest of the blog post and I did end up using a lot of words, hehe). They also omit a derivation of the principal components, which I do provide. In this blog post I also try to tie principal component analysis in with the the essence of linear algebra, an excellent video series by 3Blue1Brown, which I think is really insightful.

The introduction to PCA is meant to be understood without any knowledge about linear algebra. But concepts such as continues variables and scatterplots should be familiar.

In this blogpost you will find:

  • An Introduction to PCA
  • How data analysis with PCA is done (at least in psychology)
  • PCA from the “change of basis” perspective
  • A derivation of the principal components using Lagrange multipliers
  • How to make beautiful plots with base R

The derivation section is mainly based upon Joliffe’s (1955) PCA book, but with a slightly different notation, and some links to wikipedia. This section is optional.

Introduction to PCA

Lets assume you have a dataset with \(p\) continues variables and you want to learn something about the structure in your dataset, principal component analysis (PCA) might be the tool for you. Note: there is no dependent variable here.

What is PCA

PCA is an analysis technique that reweights your orginal \(p\) variables to \(p\) new variables, called principal components, in such a way that the first principal component is the most important, and then the next, then the next and so on. Important meaning, capturing the most variance in the data. Furthermore, the principal components are independent from each other, in other words: they are uncorrelated, when you know something about principal component 1, you know nothing about principal component 2.

Example of PCA

Lets look at an example of PCA. Below you can see plot 1A. In plot 1A you can see two heavily correlated variables: \(x_1\) and \(x_2\). In plot 1B you can see red arrows (called eigenvectors). These red arrows are the weights (in the context of PCA also called loadings) to create the principal components, but they are also pointing in the direction where the data has the most variance. You can see that in 1C where I plotted a normal distribution on top of the first eigenvector to make the visual argument that in that direction the data has most variance.

If we perform PCA on this example, we will get 2 new variables (principal components) the first having the most variation, then another one having the most variation and uncorrelad to the first one. In 2A you can see a plot of both principal components. Note that the points are still at exactly the same distance from each other compared the the original data, but \(z_1\) and \(z_2\) are not correlated anymore! PCA leaves the structure of your data intact, its just a way to look at exactly the same data but from a different perspective. A perspective that is more useful to us than the original perspective, because we know that \(z_1\) is more important than \(z_2\)!

How is PCA useful?

So how is this useful? Lets assume in this example \(x_1\) one is the length of a person and \(x_2\) is the height of a person. Not surprisingly these two are very correlated with each other. Now lets look at \(z_1\) and \(z_2\) (they are the same data but presented differently), what could \(z_1\) now be? \(z_1\) is a summary of \(x_1\) and \(x_2\) that captures most variation. It could, for example, be a “size of a person” component. Then what could \(z_2\) be? Maybe \(z_2\) captures individual differences in body morphology? Maybe its something else. We know its uncorrelated with our size component. If we think \(z_2\) is not interesting enough to us, we could choose to ignore it and call it junk.

If we ignore one dimension, and instead we continue with only \(z_1\) we have a dimension reduction from 2 to 1. That is why PCA is called a dimension reduction technique. In 2B you can see our new size component.

So, with PCA you can summarize your data and learn about its structure, which can be really valuable, if your case is not so straightforward as the case I presented here.

How is PCA typically used?

So how do you determine determine the number of principal components you want to look at, or want to retain? As I stated before, you have as many principal components as you have variables.

Do determine the number of components, you look at the variance of the individual principal components compared to the others. In this example \(z_1\) had a variance of 1.53 and \(z_1\) had a variance of 0.09. So \(z_1\) explains (\(1.53 / (1.53 + 0.09) \approx 0.94\)) 94% of the variance and \(z_2\) (\(0.09 / (1.53 + 0.09) \approx 0.06\)) 6% of the variance. One might find 6% not important enough and therefore choose to ignore \(z_2\) and not investigate it further.

But how do you determine what the components mean? When deriving meaning to the principal components, researchers typically inspect the weights (or loadings). In this example \(z_1\) is \(0.707 x_1 + 0.707 x_2 = z_1\). So we know \(z_1\) is a 0.7 height and 0.7 weight from that you can deduce that the combination of the 2 must be some size variable.

Lets do a larger fake example. Lets say we have 10 variables in a dataset (you can check them out in the table below) and we want to learn about their structure. Lets perform PCA on them.

First we inspect the variances of the 10 components. After inspecting the variances of the principal components we decided to keep 3 principal components, lets say, the first one explains 70% of the variance the second 10% and the third one 5%, the rest we found to be not important enough.

Below you can see a table with their weights/loadings.

Variable name pc1 pc2 pc3
Emotions 0.512 0.031 0.031
Struggels 0.894 0.011 0.011
Tough life 0.872 0.021 0.021
Cannot get out of bed 0.873 0.031 0.031
Have no pets 0.983 0.012 0.012
Sugar level in apple 0.031 0.782 0.031
Sugar content in peach 0.011 0.893 0.011
Sugar content in pear 0.021 0.890 0.021
Sugar content in pineapple 0.031 0.390 0.031
Sugar content in the perfume you wear 0.012 0.000 0.612

We found these weights/loadings for the variables, can you derive meaning to these components? The first principal component could be depression, the second one could be sugar content in fruits starting with the letter p, and the third component is sugar content in your perfume. The first one could be depression, because it more or less, only weights variables that might have something to do with depression. Same logic holds for component 2 or 3.

From PCA we learned the grouping, and if we want to continue with only 3 variables, we could use the principal components! (In this example, you could continue with the first and the second principal component, but for the third it might make more sense to just use the original variable)

Typically, the weights you find from PCA are all non-zero and do not show a clear structure like this (with a couple of large loadings, and then a couple of very small loadings). In order to force such a nice, structure researchers typically use different variations of PCA. You could force a structure using regularization (search for sparse PCA), or you can use the good ol’ rotation techniques (search for varimax rotation for example).

How to perform PCA?

In order to do PCA you can use any statistical software package. In R you have a couple of function that do PCA, one of them being princomp(). Most functions that perform PCA, allow you to inspect the variances of the principal components and then let you stare at the loadings. You can also directly look for the eigenvectors by applying the singular value decomposition to your data matrix, in R you can do that with svd(), if you don’t know what that meant, it does not matter.

PCA: a change of basis

In this sections I assume some very elementray linear algebra knowledge. I try to explain a very neat idea, which I believe is the core of principal component analysis, so you might want to try to read this section.

Lets think about our previous example, but now the points in the \(x\), \(y\) plane are arrows in the \(x\), \(y\) plane. In the figure below you can see the vector (arrow) representation of our data, note that nothing fundamentally changed here. Points are now arrows, nothing more. 3A is fundamentally the same as 1A.

I like to think about PCA as choosing a new basis for your data that makes more sense than the standard basis. Hopefully this sentence will make sense after reading this paragraph.

In our case the standard basis is one unit in the direction of the \(x\) axis and one unit in the direction of the \(y\) axis. We can express our basis using 2 vectors: \([1, 0]\) and \([0, 1]\). By combining these 2 vectors, we can reach any point in the \(x\),\(y\) plane.

Lets do an example, lets say we have vector \([2, 3]\), in order to know where the arrow head ends up, we walk 2 units to the right along the x-axis, and then we walk 3 units up along the y-axis.

In vector notation this would look like this:

$$ 2 \begin{bmatrix}1 \\ 0 \end{bmatrix} + 3 \begin{bmatrix} 0 \\ 1\end{bmatrix} = \begin{bmatrix} 2 \\ 3\end{bmatrix}. $$

In matrix notation (which is more convenient) this would be:

$$ \begin{bmatrix}1 & 0 \\ 0 & 1\end{bmatrix} \begin{bmatrix} 2 \\ 3\end{bmatrix} = \begin{bmatrix} 2 \\ 3\end{bmatrix}. $$

This seems almost trivial, but we do not have to use this basis. We can use a basis that might be more insightful.

We can also reach any point on the \(x\),\(y\) combining the red vectors (the eigenvectors) you can see in 3A. To reach point \([2, 3]\) we can walk x units along the first eigenvector and then y units along the second eigenvector. Lets try to reach \([2, 3]\) by combining the eigenvectors:

$$ 3.5353 \begin{bmatrix}0.7072 \\ 0.7069 \end{bmatrix} -0.7080 \begin{bmatrix} 0.7069 \\ -0.7072 \end{bmatrix} = \begin{bmatrix} 2 \\ 3\end{bmatrix}. $$

In matrix notation this would be:

$$ \begin{bmatrix} 0.7072 & 0.7069 \\ 0.7069 & -0.7072 \end{bmatrix} \begin{bmatrix}3.5353 \\ -0.7080 \end{bmatrix} = \begin{bmatrix} 2 \\ 3\end{bmatrix}. $$

The nice thing about this basis with eigenvectors is, is that this basis is way more informative then the standard basis. Walking along the first eigenvector is most important, because that is the direction where data has the most variation.

With the standard basis, in order to reach \([2, 3]\) we need to matrix multiply the standard basis times \([2, 3]\), and in order to reach \([2, 3]\) using the eigenvector basis, we need matrix multiply the eigenvector basis with \([3.5353, -0.7080]\), and these are precisely the values of the first and the second principal component!

Thus, we can talk about \(x_1, x_2\) using our old boring standard basis. Or we can talk about \(x_1, x_2\) using our more informative new eigenvector basis. So,

$$ \boldsymbol{A} \boldsymbol{z} = \boldsymbol{x}, $$

where \(\boldsymbol{A}\) is the matrix of eigenvectors, \(\boldsymbol{z}\) is a data point vector containing principal component values and \(\boldsymbol{x}\) is a vector containing our original data point. Eigenvectors have the nice property that \(\boldsymbol{A}^T \boldsymbol{A} = \boldsymbol{I}\). To go the other way around (from data to principal components) you can do,

$$ \boldsymbol{z} = \boldsymbol{A}^T \boldsymbol{x}. $$

where \(\boldsymbol{A}^T\) is a matrix containing the loadings. If we put the identity matrix back which I omitted from the previous equation, we have:

$$ \boldsymbol{I} \boldsymbol{z} = \boldsymbol{A}^T \boldsymbol{x}. $$

Now it very explicitely reads: the principal components expressed in the standard coordinate system can be obtained by matrix multiplying of the loadings times the data. That is exactly what I graphed in 3B (note that the principal components are in the standard coordinate system!).

I really think this drives the point home that PCA is a way of looking at exactly the same data from a different perspective.

You are not looking at the data from the standard coordinate system, but from the eigenvector coordinate system, which we know has nice properties!

If you want more information please check out this amazing series: the essence of linear algebra which most of you will probably know. And checkout chapter 3 and 13. And as the youtuber 3Blue1Brown says feel free to “pause and ponder” here!

Finding the weights of the principal components

In this section we will dive into the mathematical derivation of the the principal components. Lots of small properties from calculus, statistics and linear algebra come together here. I put links to the relevant pages if I use a property. This derivation follows the PCA book from Joliffe. I like this derivation because it uses few results that appear to be comming from nowhere. In contrast to the derivation of PCA on wikipedia that uses the Rayleigh quotient, short derivation, but it seems more "magical" (because I do not have enough knowledge) and therefore less satisfactory to me.

Lets start! Our goal is pick \(\boldsymbol{a}_1\) in such a way that the variance of the first principal component \(\boldsymbol{a}_1^{T} \boldsymbol{x} = \boldsymbol{z}_1\) is maximal. In this case \(\boldsymbol{x}\) is a vector containing \(p\) random variables. There are always \(p\) principal components to be found. But in practice we do not use them all.

First lets find the first principal component, we do this by solving:

$$ \begin{aligned} \underset{\boldsymbol{a}_1}{\text{maximize }} f(\boldsymbol{a}_1) &= \text{var}(\boldsymbol{a}_1^{T} \boldsymbol{x}) = \boldsymbol{a}_1^{T} \boldsymbol{\Sigma} \boldsymbol{a}_1 \\ \text{subject to } g(\boldsymbol{a}_1) &= \boldsymbol{a}_1^{T} \boldsymbol{a}_1 = 1. \end{aligned} $$

Note that \(\text{var}(\boldsymbol{a}_1^{T} \boldsymbol{x}) = \boldsymbol{a}_1^{T} \boldsymbol{\Sigma} \boldsymbol{a}_1\). You can verify this by checking the basic properties of the variance operator here.

The approach will be: we take the derivative of \(f(\boldsymbol{a}_1)\) and set it to \(0\) and solve for \(\boldsymbol{a}_1\). We will impose the contraint that \(\boldsymbol{a}_1^{T} \boldsymbol{a}_1 = 1\) else we could arbitrarily keep increasing \(\boldsymbol{a}_1\) to increase the variance. This constraint could be something else, but this particular constraint makes the derivation nice.

To maximize, \( \boldsymbol{a}_1^{T} \boldsymbol{\Sigma} \boldsymbol{a}_1 \) subject to \(\boldsymbol{a}_1^{T} \boldsymbol{a}_1 = 1 \) We are going to use the technique of Lagrange multipliers, please check the wikipedia page on Lagrange multipliers because it has an excelent intuitive explanation on how it works. So lets write the previous functions \(f\) and \(g\) into a new function, also called the Lagrangian:

$$ \underset{\boldsymbol{a}_1}{\text{maximize }} f(\boldsymbol{a}_1) = \boldsymbol{a}_1^{T} \boldsymbol{\Sigma} \boldsymbol{a}_1 - \lambda ( \boldsymbol{a}_1^{T} \boldsymbol{a}_1 - 1). $$

To maximze we take derivatives and set them to \(0\). I use this wiki page to check how to take derivatives involving matrices and vectors. $$ \begin{aligned} \frac{\partial f(\boldsymbol{a}_1)}{\partial\boldsymbol{a}_1} &= \boldsymbol{0} \\ 2 \boldsymbol{\Sigma} \boldsymbol{a}_1 - 2 \lambda \boldsymbol{a}_1 &= \boldsymbol{0} \\ \boldsymbol{\Sigma} \boldsymbol{a}_1 &= \lambda \boldsymbol{a}_1 \end{aligned} $$ This result shows us that \(\boldsymbol{a}_1\) is an eigenvector of \(\boldsymbol{\Sigma}\) (this is the definition of an eigenvector: \(\boldsymbol{\Sigma} \boldsymbol{a}_1 = \lambda \boldsymbol{a}_1 \), see here and this wonderful youtube video by 3Blue1Brown) and \(\lambda\) is the eigenvalue belonging to that eigenvector. Left multiplying the previous result by \(\boldsymbol{a}_1^{T}\) gives,

$$ \begin{aligned} \boldsymbol{\Sigma} \boldsymbol{a}_1 &= \lambda \boldsymbol{a}_1 \\ \boldsymbol{a}_1^{T} \boldsymbol{\Sigma} \boldsymbol{a}_1 &= \lambda \boldsymbol{a}_1^{T} \boldsymbol{a}_1 \\ \boldsymbol{a}_1^{T} \boldsymbol{\Sigma} \boldsymbol{a}_1 &= \lambda, \end{aligned} $$

which shows that the variance of the first principal component to is equal to the eigenvalue of the first principal component. We know that the covariance matrix \(\boldsymbol{\Sigma}\) is positive semi definite (check here) and that a positive semi definite matrix has non-negative eigenvalues (check here). Thus, to pick the weights for the first principal components we can pick \(\boldsymbol{a}_1\) as the eigenvector of \(\boldsymbol{\Sigma}\) with the largest eigenvalue.

Now for the second principal component we are going to do the same optimization but with the extra constraint that the second principal component is orthogonal/independent to the first principal component.

$$ \begin{aligned} &\underset{\boldsymbol{a}_2}{\text{maximize }} f(\boldsymbol{a}_2) = \text{var}(\boldsymbol{a}_2^{T} \boldsymbol{x}) = \boldsymbol{a}_2^{T} \boldsymbol{\Sigma} \boldsymbol{a}_2 \\ &\text{subject to } \boldsymbol{a}_2^{T} \boldsymbol{a}_2 = 1 \text{ and cov}(\boldsymbol{a}_1^{T} \boldsymbol{x}, \boldsymbol{a}_2^{T} \boldsymbol{x}) = 0. \end{aligned} $$

With,

$$ \text{cov}(\boldsymbol{a}_1^{T} \boldsymbol{x}, \boldsymbol{a}_2^{T} \boldsymbol{x}) = \boldsymbol{a}_2^{T} \Sigma \boldsymbol{a}_1 = \boldsymbol{a}_2^{T} (\lambda \boldsymbol{a}_1) = \lambda \boldsymbol{a}_2^{T} \boldsymbol{a}_1 = 0. $$

Here we used the variance rules again and the definition of eigenvectors. Next we put the equation with constraints in the Lagrangian form again,

$$ \underset{\boldsymbol{a}_2}{\text{maximize }} f(\boldsymbol{a}_2) = \boldsymbol{a}_2^{T} \boldsymbol{\Sigma} \boldsymbol{a}_2 - \lambda_2 ( \boldsymbol{a}_2^{T} \boldsymbol{a}_2 - 1) - \phi (\lambda
\boldsymbol{a}_2^{T} \boldsymbol{a}_1) $$

Next we take derivatives again and set them to zero (we first solve for \(\phi\), the lagrange multiplier for the second constraint),

$$ \begin{aligned} \frac{\partial f(\boldsymbol{a}_2)}{\partial\boldsymbol{a}_2} &= \boldsymbol{0} \\ 2 \boldsymbol{\Sigma} \boldsymbol{a}_2 - 2 \lambda_2 \boldsymbol{a}_2 - \phi \lambda \boldsymbol{a}_1 &= \boldsymbol{0} \\ 2 \boldsymbol{a}^{T}_1 \boldsymbol{\Sigma} \boldsymbol{a}_2 - 2 \lambda_2 \boldsymbol{a}^{T}_1 \boldsymbol{a}_2 - \phi \lambda \boldsymbol{a}^{T}_1 \boldsymbol{a}_1 &= \boldsymbol{a}^{T}_1 \boldsymbol{0} \\ \phi \lambda &= 0 \\ \phi &= 0 \end{aligned} $$

Now we know \(\phi = 0\), lets solve for \(\boldsymbol{a}_2\),

$$ \begin{aligned} 2 \boldsymbol{\Sigma} \boldsymbol{a}_2 - 2 \lambda_2 \boldsymbol{a}_2 &= \boldsymbol{0} \\ \boldsymbol{\Sigma} \boldsymbol{a}_2 &= \lambda_2 \boldsymbol{a}_2. \end{aligned} $$

The weights for to make the second principal component is also an eigenvector of \(\Sigma\), and \(\lambda_2\) its eigenvalue! We know that \(\boldsymbol{a}_2 \neq \boldsymbol{a}_2\) else the constraint \(\lambda \boldsymbol{a}_2^{T} \boldsymbol{a}_1 = 0\) would be violated. So, we can pick the second vector of weights to be the second eigenvector with the second largest eigenvalue!

You can repeat this process for the other \(3...p\) components and you find \(\boldsymbol{a}_3...\boldsymbol{a}_p\) to be the other eigenvectors with variances \(\lambda_3 ... \lambda_p\).

This was the theory. In practice you do not have the population covariance matrix, and you would use the sample covariance matrix to find its eigenvalues and eigenvectors.

R code that created the plots

Below you can find the R code to create the plots, I presented here. I made the plots in base R. These days base R is not sexy anymore and the cool kids use ggplot, at least that is what I see in presentations on conferences and in academic papers. But with ggplot2 you have to write a lot of code, and out of the box I don’t like the look of ggplot2. But base R on the other hand looks sexy and you can make whatever you want with very minimal code.

library(MASS)

#########################################################
# Data generation

# Generate some data and center it afterwards
set.seed(1)
S  <- matrix(c(1, 0.9, 0.9, 1), 2 , 2)
X <- mvrnorm(100, c(0.5, 1), S)
X <- scale(X, scale=FALSE)
x1 <- X[, 1]
x2 <- X[, 2]

#########################################################
# Estimation

# Perform the PCA analysis
# Extract the eigenvectors of the covariance matrix
V <- svd(X)$v
d <- svd(X)$d # extract the singular values (d = sqrt(eigenvalues))
d <- d / sum(d) # calculate percentage of variance per component

f <- 5 # some scaling factor to make the plots nice

# determine the principal components
Z <- X %*% V
z1 <- Z[, 1] # 1st component
z2 <- Z[, 2] # 2nd component

#########################################################
# Plots

# lower and upper limits of the plots
ll <- -3
up <- 3
pr <- c(ll, up) # plot range

#########################################################
# plots of x1, x2 and the eigenvectors

scale <- 1.5
# Open a "device" to draw the plot in ratio 3:1 
pdf("pca_plot1.pdf", width = 9 * scale, height = 3 * scale) 
# Set outerplot margins so title won't fall off
par(mfrow=c(1, 3), oma=c(0,0,3,0)) 

plot(x1, x2, xlim = pr, ylim = pr, asp = 1)
title("1A")
abline(h=0, col="grey")
abline(v=0, col="grey")

plot(x1, x2, xlim = pr, ylim = pr, asp = 1)
title("2A")
abline(h=0, col="grey")
abline(v=0, col="grey")
arrows(0, 0, V[1, 1] * d[1] * f, V[2, 1] * d[1] * f, col = "red")
arrows(0, 0, V[2, 1] * d[2] * f, V[2, 2] * d[2] * f, col = "red")

# Project the points in X, onto the first PC
# cbind those projected points to X
# I use the original points and the new projections in segments()
v1 <- V[, 1]
A <- cbind(X, t(v1 %*% t(v1) %*% t(X)))

# create a rotated normal-dist with variance equal to the 1st PC
norm_x <- seq(-4, 4, length=100)
norm_y <- dnorm(seq(-4, 4, length=100), mean = 0, sd = sqrt(var(z1)))
angle <- acos(V[, 1] %*% c(0, 1)) # the angle between the x-axis and the fist eigenvector
rot_mat <- matrix(c(cos(angle), sin(angle), -sin(angle), cos(angle)), 2, 2)
norm_x_y_rotated <- rot_mat %*% t(cbind(norm_x, norm_y * 5)) # rotate normal distribution
# shift the points up a fraction in the opposite direction of the 2nd eigenvector
norm_x_y_rotated <- t(norm_x_y_rotated + -1 * V[, 2] * 0.1) 


plot(x1, x2, xlim = pr, ylim = pr, asp = 1)
title("3A")
abline(h=0, col="grey")
abline(v=0, col="grey")
arrows(0, 0, V[1, 1] * d[1] * f, V[2, 1] * d[1] * f, col = "red")
arrows(0, 0, V[2, 1] * d[2] * f, V[2, 2] * d[2] * f, col = "red")
segments(x0 = A[, 1], y0 = A[, 2], x1 = A[, 3], y1 = A[, 4], col = "darkgreen")
lines(norm_x_y_rotated[, 1], norm_x_y_rotated[, 2])

title("Relationship between x1 and x2 with eigenvectors", outer = TRUE)
dev.off()



#########################################################
# plots of the principal components 

# Plots that show the principal components
pdf("pca_plot2.pdf", width = 6 * scale, height = 3 * scale) 
par(mfrow=c(1, 2), oma=c(0,0,3,0)) 

plot(z1, z2, xlim = pr, ylim = pr, asp  =1)
title("2A")
abline(h=0, col="grey")
abline(v=0, col="grey")
arrows(0, 0, f * d[1], 0, col = "red")
arrows(0, 0, 0, f * d[2], col = "red")

plot(z1, rep(1, length(z1)), type="b", ylab="", yaxt='n')
title("2B")
abline(v=0, col="grey")

title("Relationship between the principal components z1 and z2", outer = TRUE)
dev.off()


#########################################################
# Plots that show the linear transformation

pdf("pca_plot3.pdf", width = 6 * scale, height = 3 * scale) 
par(mfrow=c(1, 2), oma=c(0,0,3,0)) 

plot(x1, x2, xlim = pr, ylim = pr, asp = 1, type="n")
title("3A")
abline(h=0, col="grey")
abline(v=0, col="grey")
arrows(0, 0, V[1, 1] , V[2, 1] , col = "red", lwd="5")
arrows(0, 0, V[2, 1] , V[2, 2] , col = "red", lwd="5")
arrows(x0 = 0, y0 = 0, x1 = X[, 1], y1 = X[, 2], col = "black", length = 0.05, angle = 20)   


plot(z1, z2, xlim = pr, ylim = pr, asp = 1, type="n")
title("3B")
abline(h=0, col="grey")
abline(v=0, col="grey")
arrows(0, 0, 0 , 1, col = "red", lwd = "5")
arrows(0, 0, 1, 0, col = "red", lwd = "5")
arrows(x0 = 0, y0 = 0, x1 = z1, y1 = z2, col = "black",length = 0.05, angle = 20)  

title("Linear transformation from X to Z", outer = TRUE)
dev.off()


# Note: Afterwards I converted all pdfs to svg with the tool pdf2svg
# I think I could have created svg's from the beginning... :)