R Shiny in the Classroom

Three years ago, while I was teaching elementary statistics at a community college in California, finally fed up with the cost and deficiencies of the technology foisted on students by proprietary textbook sites and math homework websites, I decided to create a suite of R Shiny applications which would provide the students with a free computing resource to accomplish three things. It would:

(1) give students the means to perform all basic statistical computations associated with an elementary statistics course. In particular, it would allow students to perform tests and compute confidence interval constructions beginning either raw data or, as is often the case in textbook problems, working from summary statistics,

(2) provide data simulations and interactive demonstrations to illustrate fundamental statistical concepts, and

(3) assist students in deciding what sort of statistical tool/computation is appropriate to use in a given situation through the use of an ‘expert system’ front end.

As matters eventuated, I ended up quitting my teaching and going on a very long bike ride across Western Europe before I could implement my idea (a different and perhaps more interesting tale). But I finally did do so when I returned to the US, deploying my apps using an R Shiny server deployed on an NGINX webserver on a cloud instance of Ubuntu 16.04 which I initiated specifically for the purpose. The result is here:

http://104.236.178.75:3838/sort/sorta/ Many of the apps could certainly be refined and stabilized. I am completely self-educated in all things computing, in particular as regards R programming and R Shiny app development. These apps were written as learning projects for me as much as they were intended to address the three goals enumerated above. But as an outgrowth of writing the statistics apps, it became clear to me that Shiny apps are widely adaptable to a host of pedagogical purposes, really across disciplines, although, because of the nature of R, applications for use in the teaching of math, statistics, and the sciences come first to mind.

I went on to create a few dozen Shiny apps for non-statistical but mostly pedagogical applications. Those of vaguely mathematical theme are here


https://www.apclam.com/shinymath.html


whereas an over-arching point of access to all such apps is here

https://www.apclam.com/shinydir.html .

My point here is not so much to draw attention to the apps which I have developed and deployed which, as I freely admit are rudimentary in many respects, as rather to call the attention of deans, department chairwomen and chairmen, and faculty members at high schools, colleges and univesities to the opportunity to use R Shiny to liberate curricula, and budgets, from the strictures imposed by propriety websites and software served up by textbook publishers and various other academic hangers-on. Here some key points militating in favor of doing so:

(1) The ‘pro’ version of the R Shiny server is available to academic institutions at no to very low cost.

(2) The learning curve involved in writing and deploying R Shiny apps is not steep. For people who already have experience writing R scripts it can be fairly characterized as easy. Even for people who have limited programming background, it is a manageable task, particularly if given institutional guidance in the form of workshops or training materials.

(3) Web activities can be created which are tuned to the very specific needs of one’s own curriculum, institution and even classroom. In particular, R’s extensive graphics capabilities in its base graphics package can be leveraged to create learning tools which integrate numerical, symbolically and graphical approaches to the demonstration of and interaction with key concepts.

(4) Developing and deploying R Shiny apps gives faculty and staff a constructive outlet for creativity which provides opportunities for collaboration within and across disciplines and results in a tangible (OK, internet ‘tangible’) product which can be used by students and faculty throughout the world.

I would be happy to share my experiences in this realm to help anyone or any institution pursue this path. I think the possibilies in this direction are almost without limit. Feel free to email me at [email protected].

COVID-19 Posts: A Public Dataset Containing 400+ COVID-19 Blog Posts

Over the last few months, we’ve been collecting hundreds of COVID-19 blog posts from the R community. Today, we are excited to share this dataset publicly, to help bloggers who want to analyze COVID-19 data by unleashing R and the resources of its community by being able to research such posts.

So far, we have found and recorded 423 COVID posts in English. In an effort to encourage others to explore such posts, we’ve published a Shiny web app which allow users to find the names of the 231 bloggers who wrote those posts, their roles, and their country of focus. The app also lets users interactively search the collection of posts by primary topic, post title, date, and whether the post uses a particular mathematical technique or data source. To learn more about the evolution of this dataset, one of the authors (Rees) has published nine articles on Medium, which you can find here.

We encourage users to submit their own posts-or others’ posts-for inclusion, which can be done on this Google Form. Our dataset, as well as the code for the Shiny app, is available on GitHub. If anyone has corrections to the dataset, please write Rees (at) ReesMorrison (dot) com.

The remainder of this post highlights some of the findings from the dataset of COVID-19 posts. As will be made evident by the plots that follow, this is by no means a comprehensive review of every COVID-19 R blog post, but rather an overview of the data that we have found.

Posts Over Time

As the pandemic has progressed, fewer bloggers have engaged with COVID-related data, as we notice that blog posts peaked in March of 2020.

Some bloggers have been prolific; many more have been one and done. The plot below shows the names and posts of the 23 bloggers who have so far published at least four posts. For an example of how to read the plot, Tim Churches, at the bottom of the y-axis, has published a total of nine posts, but none after early April.

The color of the points corresponds to the work role of the blogger as explained in the legend at the bottom. It is immediately apparent that professors and academic researchers predominate in this group of bloggers. If you include the postgraduate students, universities writ large account for nearly all of the prolific bloggers.

Roles of Authors

The bloggers in our dataset describe their work-day roles in a variety of ways. One of the authors (Rees) standardized these job roles by categorizing the multitude of terms and descriptions, but it is quite possible that this effort misrepresented what some of these bloggers do for a living. We welcome corrections.

We’ve further categorized roles into a broad typology where professions fall into one of five categories: university, corporate, professional, government, and nonprofit. Those broader categories are represented as columns in the following chart.

Data Sources

A greater number of data sources related to COVID-19 will yield richer insights. Combining different datasets can shed new light on an issue, yield improvements, and allow authors to contruct better indices and measures. For that reason, one of the authors (Rees) extracted dataset information from our collection of blog posts.

For the most part, bloggers identified the data source they drew on for their analysis. On occasion, we had to apply some effort to standardize the 140 data sources.

By far the most prevalent data source is Johns Hopkins University, who early, comprehensively and consistently has set the standard for COVID-19 data collection and dissemination to the public.

Blog Post Topics

It may also be the case that readers want a summary of blogs, or to only look at posts that pertain to a certain topic. Assigning each blog post a primary topic introduces a fair amount of subjectivity, to be sure, but the hope is that these broad topics will help researchers find content and colleagues who share similar interests.

Here, a balloon plot shows various categories that the 423 posts address as their primary topic. Topics fall on the y axis and the blogger’s category of employment is on the x axis. The size (and opacity) of each bubble represents the count of posts that match that combination. Epidemiology leads the way, as might be expected, but quite a few posts seem to use COVID data to showcase something else, or apply R in novel ways.

Concluding Thoughts

We encourage you to use our Shiny application to explore the data for yourself. If you’d like to submit your post to be included, fill out this Google Form.

As we note in the footer of the application, the R community is intelligent and produces interesting content, but not all of us are experts when it comes to COVID-19. Engaging with these posts will allow you to better understand the application of R to our current moment, and perhaps provide feedback to post authors. We do not endorse the findings of any particular author and encourage you to find accurate, relevant, and recent information from reputable sources such as the CDC and the WHO.


Originally published on my blog. Follow the authors Rees and Connor on Twitter.

Most popular on Netflix, Disney+, Hulu and HBOmax. Weekly Tops for last 60 days

Couple months ago I published Most popular on Netflix. Daily Tops for last 60 days – small research based on daily scraping data answering following questions: How many movies (titles) made the Netflix Daily Tops? What movie was the longest #1 on Netflix? For how many days movies / TV shows stay in Tops and as #1? etc.
This time I am sharing analysis of the most popular movies / TV shows across Netflix, Disney+, Hulu and HBOmax on weekly basis, instead of daily, with anticipation of better trends catching.

So, let`s count how many movies made the top5, I assume it is less than 5 *60…
library(tidyverse)
library (gt)

platforms <- c('Disney+','HBOmax', 'Hulu', 'Netflix') # additionally, load CSV data using readr 
Wrangle raw data – reverse (fresh date first), take top 5, take last 60 days
fjune_dt % rev () %>% slice (1:5) %>% select (1:60)
fdjune_dt % rev () %>% slice (1:5) %>% select (1:60)
hdjune_dt % rev () %>% slice (1:5) %>% select (1:60)
hulu_dt % rev () %>% slice (1:5) %>% select (1:60)
Gather it together and count the number of unique titles in Top5 for 60 days
fjune_dt_gathered <- gather (fjune_dt)
fdjune_dt_gathered <- gather (fdjune_dt)
hdjune_dt_gathered <- gather (hdjune_dt)
hulu_dt_gathered <- gather (hulu_dt)
unique_fjune_gathered % length ()
unique_fdjune_gathered % length ()
unique_hdjune_gathered % length ()
unique_hulu_gathered % length ()
unique_gathered <- c(unique_fdjune_gathered, unique_hdjune_gathered, unique_hulu_gathered, unique_fjune_gathered)
unique_gathered <- as.data.frame (t(unique_gathered), stringsAsFactors = F)
colnames (unique_gathered) <- platforms
Let`s make a nice table for the results
unique_gathered_gt %
tab_header(
  title = "Number of unique movies (titles) in Top5")%>%
  tab_style(
    style = list(
      cell_text(color = "purple")),
    locations = cells_column_labels(
      columns = vars(HBOmax)))%>%
  tab_style(
    style = list(
      cell_text(color = "green")),
    locations = cells_column_labels(
      columns = vars(Hulu))) %>%
  tab_style(
    style = list(
      cell_text(color = "red")),
    locations = cells_column_labels(
      columns = vars(Netflix)))
unique_gathered_gt


Using similar code we can count the number of unique titles which were #1 one or more days


What movie was the longest in Tops / #1?
table_fjune_top5 <- sort (table (fjune_dt_gathered$value), decreasing = T) # Top5
table_fdjune_top5 <- sort (table (fdjune_dt_gathered$value), decreasing = T)
table_hdjune_top5 <- sort (table (hdjune_dt_gathered$value), decreasing = T)
table_hulu_top5 <- sort (table (hulu_dt_gathered$value), decreasing = T)
Plotting the results
bb5fdjune <- barplot (table_fdjune_top5 [1:5], ylim=c(0,62), main = "Days in Top5, Disney+", las = 1, col = 'blue')
text(bb5fdjune,table_fdjune_top5 [1:5] +2,labels=as.character(table_fdjune_top5 [1:5]))
bb5hdjune <- barplot (table_hdjune_top5 [1:5], ylim=c(0,60), main = "Days in Top5, HBO Max", las = 1, col = 'grey', cex.names=0.7)
text(bb5hdjune,table_hdjune_top5 [1:5] +2,labels=as.character(table_hdjune_top5 [1:5]))
bb5hulu <- barplot (table_hulu_top5 [1:5], ylim=c(0,60), main = "Days in Top5, Hulu", las = 1, col = 'green')
text(bb5hulu,table_hulu_top5 [1:5] +2,labels=as.character(table_hulu_top5 [1:5]))
bb5fjune <- barplot (table_fjune_top5 [1:5], ylim=c(0,60), main = "Days in Top5, Netflix", las = 1, col = 'red')
text(bb5fjune,table_fjune_top5 [1:5] +2,labels=as.character(table_fjune_top5 [1:5]))

The same for the movies / TV shows reached the first place in weekly count


Average days in top distribution
#top 5
ad5_fjune <- as.data.frame (table_fjune_top5, stringsAsFActrors=FALSE)
ad5_fdjune <- as.data.frame (table_fdjune_top5, stringsAsFActrors=FALSE)
ad5_hdjune <- as.data.frame (table_hdjune_top5, stringsAsFActrors=FALSE)
ad5_hulu <- as.data.frame (table_hulu_top5, stringsAsFActrors=FALSE)
par (mfcol = c(1,4))
boxplot (ad5_fdjune$Freq, ylim=c(0,20), main = "Days in Top5, Disney+")
boxplot (ad5_hdjune$Freq, ylim=c(0,20), main = "Days in Top5, HBO Max")
boxplot (ad5_hulu$Freq, ylim=c(0,20), main = "Days in Top5, Hulu")
boxplot (ad5_fjune$Freq, ylim=c(0,20), main = "Days in Top5, Netflix")

The same for the movies / TV shows reached the first place in weekly count (#1)

DN Unlimited 2020: Europe’s largest data science gathering | Nov 18 – 20 online

  • The DN Unlimited Conference will take place online for the first time this year
  • More than 100 speakers from the fields of AI, machine learning, data science, and technology for social impact, including from The New York Times, IBM, Bayer, and Alibaba Cloud
  • Fully remote networking opportunities via a virtual hub

Europe’s largest data science community launches new digital platform for this year’s conference

The Data Natives Conference, Europe’s biggest data science gathering, will take place virtually and invite data scientists, entrepreneurs, corporates, academia, and business innovation leaders to connect on November 18-20, 2020. The conference’s mission is to connect data experts, inspire them, and let people become part of the equation again. With its digital networking platform, DN Unlimited expects to reach a new record high with 5000+ participants. Visitors can expect keynotes and panels from the industry experts and a unique opportunity to start on new collaborations during networking and matchmaking sessions. In 2019, the sold-out Data Natives conference gathered over 3000 data, technology professionals and decision-makers from over 30 countries, including 29 sponsors, 45 community and media partners, and 176 speakers.The narrative of DN Unlimited Conference 2020 focuses on assisting the digital transformation of businesses, governments, and communities by offering a fresh perspective on data technologies – from empowering organizations to revamp their business models to shedding light on social inequalities and challenges like Climate Change and Healthcare accessibility.

Data science, new business models and the future of our society

In spring 2020, the Data Natives community of 80.000 data scientists mobilised to tackle the challenges brought by the pandemic – from the shortage of medical equipment to remote care – in a series of Hackcorona and EUvsVirus hackathons. Through the collaboration of governments such as the Greek Ministry for Digital Governance, institutions such as the Charité and experts from all over Europe, over 80 data-driven solutions have been developed. DN Unlimited conference will continue to facilitate similar cooperation.

The current crisis demonstrates that only through collaboration, businesses can thrive.

While social isolation may be limiting traditional networking opportunities, we are more equipped than ever before to make connections online. “…The ability to connect to people and information instantly is so common now. It’s just the beginning of an era of even more profound transformation. We’re living in a time of monumental change. And as the cloud becomes ambiguous, it’s literally rewriting entire industries” – says Gretchen O’Hara, Microsoft VP; DN Unlimited & Humanaize Open Forum speaker.

The crisis has called for a digital realignment from both companies and institutions. Elena Poughia, the Founder of Data Natives, perceives the transformation as follows: “It’s not about deploying new spaces via data or technology – it’s about amplifying human strengths. That’s why we need to continue to connect with each other to pivot and co-create the solutions to the challenges we’re facing. These connections will help us move forward.” 

The DN Unlimited Conference will bring together data & technology leaders from across the globe – Christopher Wiggins (Chief Data Scientist, The New York Times), Lubomila Jordanova (CEO & Founder, Plan A), Angeli Moeller (Bayer AG, Head Global Data Assets), Jessica Graves (Founder & Chief Data Officer, Sefleuria) and many more will take on the virtual stages to talk about the growing urge for global data literacy, resources for improving social inequality and building a data culture for agile business development. 

On stage among others:
  • Erika Cheung, Executive Director, Ethics in Entrepreneurship
  • Cory Doctorow, Science Fiction Author, Activist, and Journalist.
  • Whurley (William Hurley), Eisenhower Fellow, a Senior Member of the Institute of Electrical and Electronics Engineers (IEEE)
  • Alistair Croll, Founder, Solve for Interesting
  • Clare Jones, Chief Commercial Officer, what3words
  • Mark Turrell, Orcasci, CEO
  • And many more

Spatial Data Analysis Using Artificial Neural Networks, Part 2

Spatial Data Analysis Using Artificial Neural Networks, Part 2

Richard E. Plant

Additional topic to accompany Spatial Data Analysis in Ecology and Agriculture using R, Second Edition http://psfaculty.plantsciences.ucdavis.edu/plant/sda2.htm Chapter and section references are contained in that text, which is referred to as SDA2. Data sets and R code are available in the Additional Topics through the link above.

Forward

This post originally appeared as an Additional Topic to my book as described above. Because it may be of interest to a wider community than just agricultural scientists and ecologists, I am posting it here as well. However, it is too long to be a single post. I have therefore divided it into two parts, posted successively. I have also removed some of the mathematical derivations. This is the second of these two posts. In the first post we discussed ANNs with a single hidden layer and constructed one to see how it functions. This second post discusses multilayer and radial basis function ANNs,  ANNs for multiclass problems, and the determination of predictor variable importance.  Some of the graphics and formulas became somewhat blurred when I transfered them from the original document to this post. I apologize for that, and again, if you want to see the clearer version you should go to the website listed above and download the pdf file.

3. The Multiple Hidden Layer ANN

Although artificial neural networks with a single hidden layer can provide quite general solution capabilities (Venables and Ripley, 2002, p. 245), preference is sometimes expressed for multiple hidden layers. The most widely used R ANN package providing multiple hidden layer capability is neuralnet (Fritsch et al., 2019). Let’s try it out on our training data set using two hidden layers with four cells each (Fig. 19).

> set.seed(1)
> mod.neuralnet <- neuralnet(QUDO ~ MAT +
+     Precip,
data = Set2.Train,
+     hidden = c(4, 4), act.fct = “logistic”,

+     linear.output = FALSE)
> Predict.neuralnet <-
+     predict(mod.neuralnet, Set2.Test)

> plotnet(mod.neuralnet, cex = 0.75) # Fig. 19a
> p <- plot.ANN(Set2.Test,
+  Predict.neuralnet, 0.5,
“neuralnet Predictions, Two Hidden Layers”) 

The argument hidden = c(4, 4) specifies two hidden layers of four cells each, and the argument linear.output = FALSE specifies that this is a classification and not a regression problem. As with the single layer ANNs of the previous section, the two layer ANNs of neuralnet() produce a variety of outputs, some tame and some not so tame (Exercise 6).



Figure 19. (a) Diagram of the neuralnet ANN with two hidden layers; (b) prediction map of the ANN.

 
Instead of using Newton’s method found in the function nlm() and approximated in nnet(),  it has become common in ANN work to use an alternative method called gradient descent. The neuralnet package permits the selection of various versions of the gradient descent algorithm. Suppose for definiteness that the objective function J is the sum of squared errors (everything works the same if it is, for example, the cross entropy, but the discussion is more complex). Denote the combined sets of biases and weights by wi, 1 = 1,…,nw,  (for our example of Section 2, with one hidden layer and M = 4, nw = 17). If the ANN returns an estimate of the vector Y of class labels then we can then write the sum of squared errors of the ANN as

.

(7)
The intent of this notation is to indicate explicitly that the vector of estimates  depends on the biases and weights. Minimizing  J(w1,…,wnw) is the unconstrained nonlinear minimization problem to which the gradient descent method is applied. The set of values of J(w1,…,wnw) describes a surface (technically a “hypersurface”) sitting “above” the nw dimensional space of the weights and biases. If you have trouble visualizing this, it might help to imagine the case in which nw = 2, in which case the surface of values of J(w1,w2)  lies above the plane of values of (w1,w2). The Wikipedia article discussing the gradient descent method is truly excellent, and I will not try to improve on it. From this article we can see that in order to find the minimum of the objective function J, starting from a particular point in the  space we move in the direction of the gradient

.

(8)
Here the ei are unit vectors in the space of the weights and biases, and the gradient points in the direction of steepest descent (and of steepest ascent – which would take us in the wrong direction) of J. As described in the Wikipedia article, the algorithm moves in a sequence of steps

.

(9)
Here w(k) is the estimate of the vector w of biases and weights at the kth gradient descent step, and  is the length of the step taken in the direction of the negative gradient (sometimes called the learning rate). All of the partial derivatives in Equation (8) must be estimated numerically, so it is evident that although the problem is much smaller than that involving the computation of the Hessian matrix of second partial derivatives, it still involves a lot of computation. For that reason, ANN algorithms often use an approximate form of gradient descent. The most common of these is stochastic gradient descent. In this method, instead of estimating all nw partial derivatives at each step, a random subset of a fixed size (possibly as small as one) is chosen and the direction is determined based on minimizing this subset. If you have read the Wikipedia article above then this short description provides a good explanation of stochastic gradient descent, and this discussion provides more detail. In particular, you can see from these articles that it is even possible that the stochastic gradient descent method can actually move past a local minimum and find the global minimum. In fact, the gradient in Equation (9) is not actually computed directly. Instead, an algorithm called back propagation is generally used to estimate the partial derivatives. The Additional Topic on which this post is based includes a mathematical description of back propagation, but I have decided to remove that from this post, both to bring the post to a reasonable length and to reduce the mathematical complexity. The basic idea of back propagation is to compute the partial derivatives in Equation (8) by repeated applications of the chain rule of differential calculus. You can see the full derivation by downloading the Additional Topic. The neuralnet() function has available several forms of back propagation. As a default it uses the resilient propagation algorithm (Riedmiller et al., 1993). This algorithm recognizes that the size of the step from w(k to w(k+1 depends not only on the size of  but also on the sizes of the components of , and that this latter dependence may be disadvantageous. For this reason resilient propagation develops a step size that is less dependent on the magnitudes of the components of , using these values primarily to determine the direction of the step. There are other methods available and you can see the neuralnet package documentation for details. The package also makes available the cross-entropy error function in addition to the default error sum of squares. Now that we have an idea of how the function neuralnet() works, let’s try it out on the full QUDO data set. We will use two hidden layers of four cells each, similar to Fig 19a. When I worked with on this problem it turned out that neuralnet() failed to converge in several attempts when applied to the full data set. The largest training set for which convergence was obtained had n = 200 data records evenly divided between QUDO = 0 and QUDO = 1. Fig. 20a shows the result of applying neuralnet() to this training data set using a cutoff of 0.73, which was determined from a ROC curve analysis (code not shown). The error rate is 0.225, which is not as good as that obtained in Section 2 using nnet().


                          
                                       (a)                                                      (b)
Figure 20. (a) Plot of prediction regions of neuralnet() with two hidden layers and a training data set of 200 randomly selected records; (b) application of the resulting model to the full set.

  Sections 2 and 3 have given us a general idea of how a multilayer perceptron works. Although the MLP is probably the most commonly used ANN for the type of classification problem discussed here, several alternatives exist. The next section discusses one of them, the radial basis function ANN.

4. The Radial Basis Function ANN

The seemingly obscure term radial basis function ANN is actually not hard to parse out. We will start with word basis. As you may remember from linear algebra, a basis is a set of vectors in the vector space that have the property that any vector in the vector space can be described as a linear combination of elements of this set. For discussion and a picture see here. If you had a course in Fourier series, you may remember that the Fourier series functions form a basis for a set of functions on an infinite dimensional vector space. Radial basis functions do the same sort of thing. A radial function is a function of the form
                  , (10)
where c is the center point of a vector space. In other words, a radial function is radially symmetric and depends only on the distance of the vector x from the center. Again we are using the terms “point” and “vector” synonymously. The radial basis function commonly used in ANNs is the Gaussian function (i.e., the normal distribution). When x is a two dimensional quantity, as it is in our case (MAT and Precip), the function  can be visualized, and a picture of it is shown here. Thus, as with a Fourier series, a radial basis function forms a basis with which to approximate other functions – in particular, in our case, functions associated with the process of classification via an ANN.        
Figure 21. Architecture of the radial basis function ANN.

  The way a radial basis function (RBF) works is as follows. Suppose again that the cost function J is based on the sum of squared errors. Referring to Fig. 21, which shows a schematic of a radial basis function ANN with a single hidden layer, the input activations for the ith data record are, as with the multilayer perceptron, simply the values Xi,j. We will describe the combined output of all of the hidden cells in terms of the basis function . Specifically, for a fixed i define a function f(X) by
. (11)
Here cm is the center of the mth radial basis function and jm measures the size of the contribution of this component to the overall function . If our objective function is the sum of squared errors then we plug Equation (11) into the definition of J to get
. (12)
Each of the M elements of the sum  forms one of the cells in the hidden layer of the ANN (Fig. 19) (note that the order of the summation has been reversed from Section 2 so that first the summation over i takes place first in each cell and then second over the m hidden cells). The links between the hidden cells and the output cell hold the weights jm. The weighted outputs of the m cells, given by , are summed to determine the activation a(3) of the output cell. The architecture described here and shown in Fig. 21 is the simplest form of a radial basis function ANN. More bells and whistles can be added; for example, biases can be incorporated as constant terms in the sums of Equation (11). In addition, a logistic or other function can be interposed between the output cell level internal value  and the activation a(3), similar to that of the MLP (Fig. 8).
The first step in the RBF computational process is to determine a set of centers cm in Equation (11). The determination is made by trying to place the centers so that their location matches as closely as possible the distribution in the data space of values . Some RBF systems do this by carrying out a k-means clustering (SDA2 Sec. 12.6.1) of the space of data records, where the number of means k equals the number M of hidden cells (e.g. Nunes da Silva et al., 2017, p.121). Another possibility is to carry out a Kohonen training process. This is basically an unsupervised clustering algorithm analogous to k-means clustering, although, as is often the case with methods related to ANNs, the terminology is a bit more pretentious: in this case the result is called a “self-organizing map.” We will be using the function rbf() of the RSNNS package (Bergmeir and Benitez, 2012). It does not use k-means clustering, but the use of the Kohonen training process is available as an option. The default method, which in my experience works the best, is called the “rbf weights” method. Fundamentally, the rbf weights method is, in the words of the creators of the package (Zell et al, 1998, p. 177), “rather simple.” Essentially, the weights are distributed in the data space to match a random sample of the values of the data records. There are a few adjustments that can be used, but this description captures the basics. Once the values of the centers cm are established the final step of the training process is to determine the values of the link weights jm along with the values of the biases if there are any. This is accomplished in rbf() by gradient descent. Similarly to the multilayer perceptron ANNs described in Section 3, the gradient descent algorithm carries out an iterative process in which the link weights are adjusted by a series of small steps of the form

,

(13)
where   is a small constant. This brief introduction is sufficient to allow us to try out an RBF ANN. The package RSNNS is a port to R of the Stuttgart Neural Network Simulator (Zell et al, 1998), or SNNS, which contains implementations of many ANNs. As a segue to the use of the function rbf(), which creates a radial basis function ANN, we will first test the function mlp() from the same package. This provides a multilayer perceptron solution that can be compared with those of Sections 2 and 3. The use of functions with the SNNS package is slightly different from what we have seen in Sections 2 and 3. Working initially with the 50 record Training data set, we first prepare its data records to serve as arguments for the ANN functions.

library(RSNNS)
Train.Values <- Set2.Train[,c(“MAT”, “Precip”)]
Train.Targets <-
+   decodeClassLabels(as.character(Set2.Train$QUDO))
 

We will start by testing the RSNNS multilayer perceptron using a single hidden layer with four cells. The package has a very nice function called plotIterativeError() that permits one to visualize how the error declines as the iterations proceed. We will first run the ANN and check out the output. These are slight differences in how the polymorphic R functions are implemented from those of the previous sections.


> set.seed(1)
> mod.SNNS <- mlp(Train.Values, Train.Targets,
+    size = 4)

> Predict.SNNS <- predict(mod.SNNS,
+     Set2.Test[,1:2])

> head(Predict.SNNS)
         [,1]      [,2]
[1,] 0.6385500 0.3543552
[2,] 0.6247426 0.3684126
[3,] 0.6107571 0.3826691
[4,] 0.5966275 0.3970889
[5,] 0.5823891 0.4116347
[6,] 0.5680778 0.4262682
 

For this binary classification problem the output of the function predict() is in the form of a matrix whose columns sum to 1, reflecting the alternative predictions. Using this information, we can plot the prediction region (Fig. 22a).

  

Figure 22. Plots of (a) the predicted regions and (b) the iterative error with the default iteration limit.

  Next we plot the iterative error (Fig. 22b).


> plotIterativeError(mod.SNNS) # Fig. 22b
 

The error clearly declines, and it may or may not be leveling off. To check, we will set the argument maxit to a very large value and see what happens.


> mod.SNNS <- mlp(Train.Values, Train.Targets,
+    size = 4,
maxit = 50000)
> stop_time <- Sys.time()
> stop_time – start_time
Time difference of 5.016404 secs
> Predict.SNNS <- predict(mod.SNNS,
+    Set2.Test[,1:2])

> p <- plot.ANN(Set2.Test, Predict.SNNS[,2],
+    0.5,

+    “RSNNS mlp Prediction, 50,000 Iterations”)
> plotIterativeError(mod.SNNS) # Fig. 23b

                    
                                  (a)                                            (b)
Figure 23. Plots of (a) the predicted regions and (b) the iterative error with an iteration limit of 50,000.

  The iteration error clearly declines some more. The prediction region does not look like most of those that we have seen earlier, and does not look very intuitive, but of course the machine doesn’t know that. The time to run 50,000 iterations for this problem is not bad at all.
Now we can move on to the radial basis function ANN, implemented via the function rbf(). First we will try it with all arguments set to their default values (Fig. 24).


> set.seed(1)
> mod.SNNS <- rbf(Train.Values, Train.Targets)
> Predict.SNNS <- predict(mod.SNNS, Set2.Test)
> p <- plot.ANN(Set2.Test, Predict.SNNS[,2],
+   0.5,
“RSNNS rbf Prediction,
+   Default Values”) 

Figure 24. Prediction regions for the model developed using rbf() with default values.

 
The “radial” part of the term radial basis function becomes very apparent! A glance at the documentation via ?rbf reveals that the default value for the number of hidden cells is 5. Many sources discussing radial basis function ANNs point out that a large number of hidden cells is often necessary to correctly model the data. We will try another run with the number of hidden cells increased to 40.


> set.seed(1)
> mod.SNNS <- rbf(Train.Values, Train.Targets,
+    size = 40)
 
The rest of the code is identical to what we have seen before and is not shown. The prediction region is similar to that of Fig. 25a below and is not shown. The error rate is about what we have seen from the other methods.

> length(which(Set2.Pred$TrueVal !=
+   Set2.Pred$PredVal)) /
nrow(Set2.Train)
[1] 0.22
                              
                                       (a)                                      (b)
Figure 25. (a) Prediction regions and (b) error estimate for the model developed using rbf() with the full data set.

  The performance of rbf() on the full data set is not particularly spectacular.

> length(which(Set2.Pred$TrueVal !=
+   Set2.Pred$PredVal)) /
nrow(data.Set2U)
[1] 0.2324658  

Now that we have had a chance to compare several ANNs on a binary selection problem, an inescapable conclusion is that, for the problem we are studying at least, the old workhorse nnet() acquits itself pretty well. In Section 5 we will try out various ANNs on a multi-category classification process. In this case each of the ANNs we are considering has an output cell for each class label.  

5. A Multicategory Classification Problem

The multicategory classification problem that we will study in this chapter is the same one that we studied in the Additional Topic on Comparison of Supervised Classification Methods. A detailed introduction is given there. In brief, the data set is the augmented Data Set 2 from SDA2. It is based on the Wieslander survey of oak species in California (Wieslander, 1935) and contains records of the presence or absence of blue oak (Quercus douglasii, QUDO), coast live oak (Quercus agrifolia, QUAG), canyon oak (Quercus chrysolepis, QUCH), black oak (Quercus kelloggii, QUKE), valley oak (Quercus lobata, QULO), and interior live oak (Quercus wislizeni, QUWI). As implemented it is the data frame data.Set2U; until now we have grouped all of the non-QUDO species together and assigned them the value QUDO = 0. In this section we will create the class label species and assign the appropriate character value to each record.

> species <- character(nrow(data.Set2U))
> species[which(data.Set2U$QUAG == 1)] <- “QUAG”
> species[which(data.Set2U$QUWI == 1)] <- “QUWI”
> species[which(data.Set2U$QULO == 1)] <- “QULO”
> species[which(data.Set2U$QUDO == 1)] <- “QUDO”
> species[which(data.Set2U$QUKE == 1)] <- “QUKE”
> species[which(data.Set2U$QUCH == 1)] <- “QUCH”
> data.Set2U$species <- as.factor(species)
> table(data.Set2U$species)
QUAG QUCH QUDO QUKE QULO QUWI
551   99  731  717   47  122
 

As can be seen, the data set is highly unbalanced. This property always poses a severe challenge for classification algorithms. The data are also highly autocorrelated spatially. In the Additional Topic on Method Comparison we found that 83% of the data records have the same species classification as their nearest neighbor. As in other Additional Topics, we want to be able to take advantage of locational information both through spatial relationships and by including Latitude and Longitude in the set of predictors. I won’t repeat the discussion about this since it is covered in the other Additional Topics. We will use the same spatialPointsDataFrame (Bivand et al, 2011) D of rescaled predictors and species class labels that was used in the Additional Topic on method comparison. Again, if you are not familiar with the spatialPointsDataFrame object, don’t worry about it: you can easily understand what is going anyway.


> library(spdep)
> coordinates(data.Set2U) <-
+   c(“Longitude”, “Latitude”)
> proj4string(data.Set2U) <-
+    CRS(“+proj=longlat +datum=WGS84”)
> D <- cbind(coordinates(data.Set2U),
+    data.Set2U@data[,c(2:19, 28:34, 43)])
> for (i in 1:2)) D[,i] <- rescale(D[,i]) 

  In this section we will use the predictors selected as optimal in the variable selection process for support vector machine analysis. This established as optimal the predictors Precip, CoastDist, the distance from the Pacific Coast, TempR, the average annual temparature range, Elevation, and Latitude. We will first try out nnet() using these predictors (Fig. 26). After playing around a bit with the number of hidden and skip layers, here is a result.

> set.seed(1)
> modf.nnet <- nnet(species ~ Precip +
+    CoastDist + TempR + Elevation + Latitude,

+    data = D, size = 10, skip = 4,
+    maxit = 10000)
# weights:  156
initial  value 3635.078808
iter2190 value 925.764437
     *     *     *     *
final  value 925.760573
converged
> plotnet(modf.nnet, cex = 0.6)

> Predict.nnet <- predict(modf.nnet, D)
> pred.nnet <- apply(Predict.nnet, 1,
+      which.max)

> 1 – (sum(diag(conf.mat)) / nrow(D))
[1] 0.1433613
> print(conf.mat <- table(pred.nnet, D$species))

pred.nnet QUAG QUCH QUDO QUKE QULO QUWI
        1  514    9   29    7   16   24
        2    1   27    2    6    2    1
        3   27    7  661   19   11   22
        4    1   54   27  673    4   20
        5    1    0    4    0   12    0
        6    7    2    8   12    2   55
> print(diag(conf.mat) /
+   table(data.Set2U$species), digits = 2)

QUAG QUCH QUDO QUKE QULO QUWI
0.93 0.27 0.90 0.94 0.26 0.45
 

The overall error rate is about 0.14, and, as usual, the rare species are not predicted particularly well.   Figure 26. Schematic of the multicategory ANN.

  In the Additional Topic on method comparison it was noted that, owing to the high level of spatial autocorrelation of these data, one can obtain an error rate of about 0.18 by simply predicting that a particular data record will be the same species as that of its nearest neighbor. We will try using that to our advantage by adding the species of the nearest and second nearest geographic neighbors to the set of predictors.


> set.seed(1)
> start_time <- Sys.time()
> modf.nnet <- nnet(species ~ Precip +
+    CoastDist + TempR +
Elevation + Latitude +
+    nearest1 + nearest2,
data = D, size = 12,
+    skip = 2, maxit = 10000)

> stop_time <- Sys.time()
> stop_time – start_time
Time difference of 11.64763 secs
> plotnet(modf.nnet, cex = 0.6)

> Predict.nnet <- predict(modf.nnet, D)
> pred.nnet <- apply(Predict.nnet, 1,
+    which.max)

> print(conf.mat <- table(pred.nnet,
+    D$species))

pred.nnet QUAG QUCH QUDO QUKE QULO QUWI
        1  527    3   13    4   11   12
        2    1   48    0    3    1    0
        3   12    2  695    7    7   18
        4    1   45   17  698    1   16
        5    6    0    2    0   27    0
        6    4    1    4    5    0   76
> 1 – (sum(diag(conf.mat)) / nrow(D))

[1] 0.08645787
> print(diag(conf.mat) /
+     table(data.Set2U$species), digits = 2)

QUAG QUCH QUDO QUKE QULO QUWI
0.96 0.48 0.95 0.97 0.57 0.62
 

The error rate is a little under 9%, which is really quite good. The improvement is particularly good for the rare species. Of course, this only works if one knows the species of the nearest neighbors! Nevertheless, it does show that incorporation of spatial information, if possible, can work to one’s advantage. The neuralnet ANN was not as successful in my limited testing. The best results were obtained when nearest neighbors were not used. The code is not shown, but here are the results.


> stop_time – start_time
Time difference of 2.726899 mins
> print(conf.mat <- table(pred.neuralnet,
+    D$species))

pred.neuralnet QUAG QUCH QUDO QUKE QULO QUWI
             1  499   22   31   16   21   35
             2    0    8    2    2    0    0
             3   51    5  654   27   22   47
             4    1   62   34  672    4   23
             6    0    2   10    0    0   17
> 1 – (sum(diag(conf.mat[1:4,1:4])) +

+         conf.mat[5,6]) / nrow(D)
[1] 0.1839435  

No records were predicted to be QULO. The performance of the SNNS MLP was about the same. The RBF ANN performed even worse.


> print(conf.mat <- table(pred.SNNS,
+   D$species))
pred.SNNS QUAG QUCH QUDO QUKE QULO QUWI
        1  471   24   42   18   13   33
        3   80   15  665   63   31   68
        4    0   60   24  636    3   21
> 1 – (conf.mat[1:1] +conf.mat[2,3] +
+      conf.mat[3,4]) / nrow(D)

[1] 0.2183502  

Once again, despite being the oldest and arguably the simplest package, nnet has performed the best in our analysis. There may, of course, be problems for which it does not do as well as the alternatives. Nevertheless, we will limit our remaining work to this package alone. The last topic we need to cover is variable selection for ANNs.

6. Variable Selection for ANNs

There have been numerous papers published on variable selection for ANNs (try googling the topic!). Olden and Jackson (2002), Gevery et al. (2003), and Olden et al. (2004) provide, in my opinion, particularly useful ones for the practitioner. Gevery et al. (2003) discuss some stepwise selection methods in which variables enter and leave the model based on the mean squared error (see SDA2 Sec. 8.2.1 for a discussion of stepwise methods in linear regression). These methods actually have much to recommend them, but they are not much different from stepwise selection methods we have discussed in SDA2 and in other Additional Topics and we won’t consider them further here. The unique distinguishing feature of ANNs in terms of variable selection is their connection weights, and several methods for estimating variable importance based on these weights have been developed. The most commonly used methods for ANNs involve a single output variable, implying only binary classification or regression. We will therefore discuss them by returning to the QUDO classification problem of Sections 2-4. We will carry over the data frame D from Section 5. Our model will include three predictors. Two of them are MAT and Precip, as used in Sections 2-4. The third, called Dummy, is just that: a dummy variable consisting of a uniformly distributed set of random numbers having no predictive power at all. A good method should identify Dummy as a prime candidate for elimination.

> set.seed(1)
> D$Dummy <- runif(nrow(D))
> size <- 2
> set.seed(1)  

For simplicity our demonstration ANN will have only two hidden cells (Fig. 27). We will first select the random number seed with the best performance, using the same procedure as that carried out in Section 2. Here is the result.


> seeds [1]  1  2  3  4  5  6  7  8  9 10 11 12 13 14 15 16 17 18 19 20> sort(error.rate)
[1] 0.2077636 0.2077636 0.2082047 0.2254080 0.2271725 0.2271725 0.2271725
[8] 0.2271725 0.2271725 0.2271725 0.2271725 0.2271725 0.2276136 0.2280547
[15] 0.2284958 0.2289369 0.3224526 0.3224526 0.3224526 0.3224526
> print(best.seed <- which.min(error.rate))

[1] 10
> mod.nnet <- nnet(X.t, Y.t, size = size,
+     maxit = 10000)

> plotnet(mod.nnet, cex_val = 0.75)

  Fig. 27 shows the plotnet() plot of this ANN. Here are the numerical values of the biases and weights


> mod.nnet$wts [1] -80.82405907   0.87215226  85.41369339 108.51778012  -4.8336825
[6]   0.04523967   2.83945431  -0.84126067  -2.81207163  -2.34322065
[11]  71.37131603
 

The order of biases and weights is {B1H1, I1H1, I2H1, I3H1, B1H2, I1H2 I2H2, I3H2, B2O1, H1O1, H2O1}.

Figure 27. plotnet() diagram of the simple ANN.

  Since the biases act on all predictors equally, they are not considered in the algorithm. One of the earliest ANN methods for determining the relative importance of each predictor was developed by Garson (1991) and Goh (1995). We will present this method following the discussion of Olden and Jackson (2002). The following replicates for our current case the set of tables displayed in Box 1 of that paper. First, we compute the matrix of weights.


> print(n.wts <- length(mod.nnet$wts))
[1] 11
> # Remove H1O1, H2O1
> print(M.seq <- seq(1,(n.wts-size)))

[1] 1 2 3 4 5 6 7 8 9
> n.X <- 3

> print(M.out <-  seq(1,(n.wts-size),
+    by = n.X + 1))

[1] 1 5 9
> print(M1 <- matrix(mod.nnet$wts[M.seq[-M.out]],

+   n.X, size), digits = 2)
      [,1]   [,2]
[1,]   0.87  0.045
[2,]  85.41  2.839
[3,] 108.52 -0.841
 

Next we compute the output weights.

> print(O <- matrix(mod.nnet$wts
+   [seq(n.wts-size+1,n.wts)],

+   1, size), digits = 2)
    [,1] [,2]
[1,] -2.3   71
 

You may want to compare the numerical values with the sign and shade of the links in Fig. 27. Table 4 shows the weights and output in in tabular form, analogous to the first table in Box 1 of Olden and Jackson (2002) (from this you can see the reason for the convention involving the arrangement of subscripts described in Section 2).  
Table 4. Input – hidden – output link weights.
  Hidden H1 Hidden H2
Input I1 w(1)11 = 0.87 w(1)12 = 0.045
Input I2 w(1)21 = 84.41 w(1)22 = 2.839
Input I3 w(1)31 = 108.52 w(1)32 = –0.841
Output w(2)1 = -2.3 w(2)2 = 71
    If your output is different from mine then of course your table will have different values, but you should be able to check them by comparing your output with mine. Next we compute the matrix C.m, which contains the contribution of each input cell to the output cell (C is a reserved word in R and can’t be used).


> C.m <- matrix(0, n.X, size)
> for (i in 1:n.X) C.m[i,] <- M1[i,] * O[1,]
> print(C.m, digits = 2)
    [,1]  [,2]
[1,]   -2   3.2
[2,] -200 202.7
[3,] -254 -60.0
 

These row sums represent the total weight that each input applies to the computation of the output. Next, we compute the matrix R, which is defined the ratio of the absolute value of each element of C.m to the column sum. This is considered to be the relative contribution of each input cell to the activation of each output cell. We won’t be generalizing this code, and it is easier to write it using the fixed values of n.X and size.


> R <- matrix(0, 3, 2)
> for (i in 1:3) R[i,] <-
+    abs(C.m[i,]) / (abs(C.m[1,]) +
+    abs(C.m[2,]) + abs(C.m[3,]))
> print(R, digits = 2)
      [,1]  [,2]
[1,] 0.0045 0.012
[2,] 0.4385 0.762
[3,] 0.5571 0.226
 

Next, we compute the matrix S, which is the sum of the relative contributions of each input cell to the activation of each output cell.

> S <- matrix(0, 3, 1)
> for (i in 1:3) S[i] <- R[i,1] + R[i,2]
> print(S, digits = 2)
     [,1]
[1,] 0.017
[2,] 1.201
[3,] 0.783
 

Finally, we compute the matrix RI, which is the relative importance of each input cell to the activation of each output cell.

> RI <- matrix(0, 3, 1)
> for (i in 1:3) RI[i,1] <- S[i,1] / sum(S)
> print(RI, 2)
      [,1]
[1,] 0.0083
[2,] 0.6003
[3,] 0.3914

The package NeuralNetTools contains the function garson(), which carries out these computations.


> print(garson(mod.nnet, bar_plot = FALSE),
+   digits = 2)

      rel_imp
Dummy   0.0083
MAT     0.6003
Precip  0.3914
 
The function garson() can also be used to create a barplot of these values (Fig. 29a).

   
                     (a)                                             (b)
Fig. 29. a) barplot of Garson importance values; b) barplot of Olden-Jackson importance values.


  Olden and Jackson (2002) criticized Garson’s algorithm on the grounds that by computing absolute values of link weights it ignores the possibility that a large positive input-hidden link weight followed by a large negative hidden-output link weight might cancel each other out. In a comparison of various methods for evaluating variable importance, Olden et al. (2004) found that Garson’s algorithm fared the worst among alternatives tested. Olden and Jackson (2002) suggest a permutation test (SDA2 Sec. 3.4) as an alternative to Garson’s algorithm. In reality, this method provides useful information even without employing a permutation test, and we will present it in that way. The test involves the following steps (we will use the same numbering as Olden and Jackson).

Step 1. Test several random seeds and select the one giving the smallest error. This step has already been carried out.
Step 2a
. Compute the matrix C.m above for the selected weight set. Again, this step has already been carried out.

> print(C.m, digits = 2)
[,1]  [,2]
[1,]   -2   3.2
[2,] -200 202.7
[3,] -254 -60.0
 

Step 2b
. Compute the row sums of C.m.

> print(C.sum <- apply(C.m, 1, sum),
+      digits = 2)

[1]    1.2    2.5 -314.3  

The package NeuralNetTools has a function olden() to carry out these computations and to plot a barplot (Fig. 29b).

> print(olden(mod.nnet, bar_plot = FALSE), digits = 2)
      importance
Dummy         1.2
MAT           2.5
Precip     -314.3
 

In this particular example, one might say that the Garson method acquits itself better than the Olden-Jackson method. This shows that one must keep an open mind when playing with these toys. Lek (1996) developed a method that is also discussed by Gevery et al. (2003). We will follow the discussion in this latter publication as well as the Details section of ?lekprofile. It is easiest to present the method via an example. The package NeuralNetTools contains the function lekprofile() that does the computations and sets up the graphical output. The default output of the function is a ggplot (Wickham, 2016) object.

> library(ggplot2)
> lekprofile(mod, steps = 5,
+    group_vals = seq(0, 1, by = 0.5))   

The figure is shown below after the explanation. The procedure divides the range of each predictor into (steps – 1) equally spaced segments. In our case, steps = 5, so there are 4 segments and 5 steps (the default steps value is 100). It is also possible to obtain the numerical values to compute the values represented in the plot. Here is a portion of it.

> lekprofile(mod.nnet, steps = 5,
+    group_vals = seq(0, 1, by = 0.5),
+    val_out = TRUE)

[[1]] Explanatory     resp_name    Response Groups exp_name
1  0.0006052661      QUDO 0.095468020      1    Dummy
2  0.2504365980      QUDO 0.096018057      1    Dummy
3  0.5002679299      QUDO 0.096577123      1    Dummy
4  0.7500992618      QUDO 0.097145396      1    Dummy
5  0.9999305937      QUDO 0.097723053      1    Dummy
6  0.0006052661      QUDO 0.189228401      2    Dummy
7  0.2504365980      QUDO 0.195418722      2    Dummy
8  0.5002679299      QUDO 0.201826448      2    Dummy
9  0.7500992618      QUDO 0.208457333      2    Dummy
10 0.9999305937      QUDO 0.215317020      2    Dummy
11 0.0006052661      QUDO 0.231719048      3    Dummy
         *         *            *           *
30 1.0000000000      QUDO 0.263814961      3      MAT
31 0.0000000000      QUDO 0.095468020      1   Precip
32 0.2500000000      QUDO 0.086684129      1   Precip
33 0.5000000000      QUDO 0.080092615      1   Precip
34 0.7500000000      QUDO 0.017896195      1   Precip
35 1.0000000000      QUDO 0.007309402      1   Precip
36 0.0000000000      QUDO 0.867860787      2   Precip
37 0.2500000000      QUDO 0.213604259      2   Precip
38 0.5000000000      QUDO 0.119027224      2   Precip
39 0.7500000000      QUDO 0.070552909      2   Precip
40 1.0000000000      QUDO 0.045104426      2   Precip
41 0.0000000000      QUDO 0.977067535      3   Precip
42 0.2500000000      QUDO 0.902834305      3   Precip
43 0.5000000000      QUDO 0.720770997      3   Precip
44 0.7500000000      QUDO 0.468084251      3   Precip
45 1.0000000000      QUDO 0.263814961      3   Precip
 
[[2]]
            Dummy       MAT    Precip
0%   0.0006052661 0.0000000 0.0000000
50%  0.4781180343 0.7601916 0.2747424
100% 0.9999305937 1.0000000 1.0000000
 

The numerical output is a list with two elements. The first element is a data frame and the second is a matrix. The explanation is clearest if we start with the second element. The components of the matrix are percentile values of each of the predictors. For example, the fiftieth percentile value of MAT is 0.7601916. Note that the 0 and 100 percentile values of MAT and Precip are 0 and 1 because these variables have been rescaled. The number of percentiles is specified via the argument group_vals, which has the default value seq(0, 1, by = 0.2). In the example we use a coarser sequence to make the interpretation easier.
Moving to the first element of the list, the data frame, the first data field is the column labeled explanatory. Its first five values represent the subdivision into five (as specified by the value of steps) equal segments of the range of values of Dummy from the 0% to the 100% percentile. For each explanatory variable, the data field exp_name, there are three Groups; the 0, 50 and 100 percentile groups, as specified by the group_vals argument. The values of the input to the mod.nnet ANN object are cycled as follows. Starting with MAT, this variable is first set to its 0 percentile value, which is 0. This forms Group 1 for MAT. Holding MAT at this value, the other two inputs are both set successively to the values specified by steps: 0, 0.25, 0.5, 0.75, and 1. The output of the ANN is calculated for each of these combinations of values and recorded in the Response data field. Thus, for example, the second data record has Response = 0.096018057, which is the value of QUDO computed by the ANN for the input vector {Dummy, MAT, Precip} = {0, 0.25, 0.25}. This value of Response is encircled by the small black circle.

 
Figure 30. a) Lek plot of the three-variable model. The response to {Dummy, MAT, Precip} = {0, 0.25, 0.25} is encircled by the small black circle; b) plot of the data.

  Now we can interpret Fig. 30a. For each of the three predictors, the colored lines represent the ANN response as that predictor is held fixed and the other two predictors are successively increased in value. For Dummy there is, unsurprisingly, very little response. Discounting the effect of Dummy, at low values of MAT (Group 1), the response QUDO is high until Precip reaches a high value. Fig 30b shows (again) the data, helping to interpret the plots in Fig. 30a. The Lek profile allows us to interpret how the predictors interact to generate the value of the class label. In terms of variable selection, we can see that the Lek plot would indeed suggest that we eliminate Dummy, since the values of QUDO are virtually unresponsive to it. Depending on how we implement it, the Lek method can be used for stepwise selection, by bringing explanatory variables in and out of the model depending on how they line up in plots such as Fig. 30, or one can simply carry out comparison of all of the explanatory variables at once. The Details section of ?lekprofile describes other uses of the results of an analysis based on the Lek method.

7. Further Reading

Bishop (1995) is probably the most widely cited introductory book on ANNs. The text by Nunes da Silva et al. (2017) is also a good resource. Venables and Ripley (2002) provide an excellent discussion at a more advanced and theoretical level. A blog by Hatwell (2016) provides a nice discussion of the variable selection methods in a regression context. The manual by Zell et al. (1998), although specifically developed for the SNNS library, gives a good practical discussion of many aspects of ANNs. Finally, this site, developed by Daniel Smilkov and Shan Carter, is a “neural net playground” where you can try out various ANN configurations and see the results (I am grateful to Kevin Ryan for pointing this site out to me).

8. Exercises

1) Repeat the analysis of Section 2 using nnet(), but apply it to a data set in which MAT and Precip are normalized rather than rescaled. Determine the difference in predicted value of the class labels.   2) Explore the use of the homemade ANN ANN.1() with differently sized hidden layers in terms of the stability of the prediction.   3) By computing the sum of squared errors for a set of predictions of nnet(), determine whether it is converging to different local minima for different random starting values and number of hidden cells.   4) Using a set of predictions of nnet(), compare the error rates when entropy is TRUE with the default of FALSE.   5) If you have access to Venables and Ripley (2002), read about the softmax option. Repeat Exercise (4) for this option.   6) Generate a “bestiary” of predictions of neuralnet() analogous to that of Fig. 9 for ANN1.().

7) Look up the word “beastiary” in the dictionary.

9. References

Beck, M. W. (2018). NeuralNetTools: Visualization and Analysis Tools for Neural Networks. _Journal of Statistical Software 85: 1-20. doi: 10.18637/jss.v085.i11 (URL: https://doi.org/10.18637/jss.v085.i11).   Bishop, C.M. (1995). Neural Networks for Pattern Recognition. Clarendon Press, Oxford, UK.   Bergmeir, C. and J. M. Benitez (2012). Neural Networks in R Using the Stuttgart Neural Network Simulator: RSNNS. Journal of Statistical Software, 46(7): 1-26. URL http://www.jstatsoft.org/v46/i07/   Bivand, R. (with contributions by M. Altman, L. Anselin, R. Assunçăo, O. Berke, A. Bernat, G. Blanchet, E. Blankmeyer, M. Carvalho, B. Christensen, Y. Chun, C. Dormann, S. Dray, R. Halbersma, E. Krainski, N. Lewin-Koh, H. Li, J. Ma, G. Millo, W. Mueller, H. Ono, P. Peres-Neto, G. Piras, M. Reder, M. Tiefelsdorf and and D. Yu) (2011). spdep: Spatial dependence: weighting schemes, statistics and models. http://CRAN.R-project.org/package=spdep.   Du, K.-L. and M.N.S. Swamy (2014). Neural Networks and Statistical Learning. Springer, London   Fritsch, S., F. Guenther and M. N. Wright (2019). neuralnet: Training of Neural Networks. R package version 1.44.2. https://CRAN.R-project.org/package=neuralnet   Gevery, M. I. Dimopoulos and S. Lek (2003). Review and comparison of methods to study the contribution of variables in artificial neural network models. Ecological Modelling 160: 249-264. Guenther, F. and S. Fritsch (2010). neuralnet: Training of Neural Networks. The R Journal 2:30-38.   Hatwell, J. (2016) Artificial Neural Networks in R. https://rpubs.com/julianhatwell/annr   Meyer, D, E. Dimitriadou, K. Hornik, A. Weingessel and F. Leisch (2019). e1071: Misc Functions  of the Department of Statistics, Probability Theory Group (Formerly: E1071), TU Wien. R package version 1.7-3. https://CRAN.R-project.org/package=e1071   Nunes DaSilva, I.N, D. Hernane Spatti, R. Andrade Flauzino, L.H. Bartocci Liboni and S.F. dos Reis Alves (2017). Artificial Neural Networks: A Practical Course. Springer International, Switzerland.   Riedmiller M. and H. Braun (1993) A direct adaptive method for faster backpropagation learning: The RPROP algorithm. Proceedings of the IEEE International Conference on Neural Networks: 586-591. San Francisco, CA.   Ripley, B.D. (1996). Pattern Recognition and Neural Networks. Cambridge University Press, Cambridge, UK   Rosenblatt, F. (1958). The perceptron: A probabilistic model for information storage and organization in the brain. Psychological Review, 65(6), 386–408. https://doi.org/10.1037/h0042519   Venables, W. N. and B.D. Ripley (2002) Modern Applied Statistics with S. Fourth Edition. Springer, New York.   Wickham, H. (2016). ggplot2: Elegant Graphics for Data Analysis. Springer-Verlag New York.   Wieslander, A. E. (1935). A vegetation type map of California. Madroño 3: 140-144.   Zell, A. and 27 others (1998), SNNS Stuttgart Neural Network Simulator User Manual, Version 4.2. IPVR, University of Stuttgart and WSI, University of Tubingen. URL http://www.ra.cs.uni-tuebingen.de/SNNS/        

Climate Change & AI for GOOD | Online Open Forum Oct 15th

Join Data Natives for a discussion on how to curb Climate Change and better protect our environment for the next generation. Get inspired by innovative solutions which use data, machine learning and AI technologies for GOOD. Lubomila Jordanova, Founder of Plan A, and featured speaker, explains that “the IT sector will use up to 51% of the global energy output in 2030. Let’s adjust the digital industry and use Data for Climate Action, because carbon reduction is key to making companies future-proof.” When used carefully, AI can help us solve some of the most serious challenges. However, key to that success is measuring impact with the right methods, mindsets, and metrics.

The founders of startups that developed innovative solutions to combat humanity’s biggest challenge, will share their experiences and thoughts: Brittany Salas (Co-Founder at Active Giving) Peter Sänger (Co-Founder/Executive Managing Director at Green City Solutions GmbH) Shaheer Hussam (CEO & Co-Founder at Aetlan) | Lubomila Jordanova (Founder at Plan A)  Oliver Arafat (Alibaba Cloud’s Senior Solution Architect)

Details
What? Climate Change & AI for GOOD | DN Unlimited Open Forum powered by Alibaba Cloud
When? October 15th at 6 PM CET
Where? Online, worldwide
Register for FREE here: https://datanatives.io/climate-change-ai-for-good-open-forum/

Zoom talk on “Organising exams in R” from the Grenoble R user group

Due to the current sanitary situation, the Grenoble (France) R user group has decided to switch 100% online. The advantage to this is that our talks will now open for anybody around the globe 🙂

The next talk will be on October 22nd, 2020 at 5PM (FR):

Organising exams in R
Package {exams} enables creating questionnaires that combine program output, graphs, etc in an automatised and dynamic fashion. They may be exported in many different formats: html, pdf, nops and most intresting xml. Xml is compatible with moodle which allows to reproducibly generate random questions in R and create a great number of different exams and make them an online moodle exam.

Link to the event: https://www.eventbrite.com/e/organising-exams-in-r-tickets-125308530187

Link to Zoom: https://us04web.zoom.us/j/76885441433?pwd=bUhvejdUb2sxa29saEk5M3NlMldBdz09

Link to the Grenoble R user group 2020/2021 calendar:  https://r-in-grenoble.github.io/sessions.html

Hope to see you there!

Spatial Data Analysis Using Artificial Neural Networks Part 1

Spatial Data Analysis Using Artificial Neural Networks, Part 1

Richard E. Plant

Additional topic to accompany Spatial Data Analysis in Ecology and Agriculture using R, Second Edition

http://psfaculty.plantsciences.ucdavis.edu/plant/sda2.htm
Chapter and section references are contained in that text, which is referred to as SDA2. Data sets and R code are available in the Additional Topics through the link above.

Forward

This post originally appeared as an Additional Topic to my book as described above. Because it may be of interest to a wider community than just agricultural scientists and ecologists, I am posting it here as well. However, it is too long to be a single post. I have therefore divided it into two parts, to be posted successively. I have also removed some of the mathematical derivations. This is the first of these two posts in which we discuss ANNs with a single hidden layer and construct one to see how it functions. The second post will discuss multilayer and radial basis function ANNs,  ANNs for multiclass problems, and the determination of predictor variable importance. All references are listed at the end of the second post (if you are in a hurry, you can go to the link above and get the full Additional Topic). Finally, some of the graphics and formulas became somewhat blurred when I transfered them from the original document to this post. I apologize for that, and again, if you want to see the clearer version you should go to the website listed above and download the pdf file.

1. Introduction

Artificial neural networks (ANNs) have become one of the most widely used analytical tools for both supervised and unsupervised classification. Nunes da Silva et al. (2017) give a detailed history of ANNs and the interested reader is referred to that source. One of the first functioning artificial neural networks was constructed by Frank Rosenblatt (1958), who called his creation a perceptron. Rosenblatt’s perceptron was an actual physical device, but it was also simulated on a computer, and after a hiatus of a decade or so work began in earnest on simulated ANNs. The term “perceptron” was retained to characterize a certain type of artificial neural network and this type has become one of the most commonly used today. Different authors use the term “perceptron” in different ways, but most commonly it refers to a type of ANN called a multilayer perceptron, or MLP. This is the first type of ANN that we will explore, in Section 2.
                                               (a)                                                                   (b)
Figure 1. Two plots of the same simple ANN. (a) Display of numerical values of connections, (b) schematic representation of connection values.

Fig. 1 shows two representations of a very simple ANN constructed using the R package neuralnet (Fritsch et al. 2019). The representation in Fig. 1a shows greater detail, displaying numerical values at each link, or connection, indicating the weight and sign. The representation in Fig. 1b of the same ANN but using the plotnet() function of the NeuralNetTools package (Beck, 2018) is more schematic, with the numerical value of the link indicated by its thickness and the sign by its shade, with black for positive and gray for negative. The circles in this context represent artificial nerve cells. If you have never seen a sketch of a typical nerve cell, you can see one here. In vast oversimplification, a nerve cell, when activated by some stimulus, transmits an action potential to other nerve cells via its axons. At the end of the axon are synapses, which release synaptic vesicles that are taken up by the dendrites of the next nerve cell in the chain. The signal transmitted in this way can be either excitatory, tending to make the receptor nerve cell develop an action potential, or inhibitory, tending to impede the production of an action potential. Multiple input cells combine their transmitted signal, and if the strength of the combined signal surpasses a threshold then the receptor cell initiates an action potential of its own and transmits the signal down its axon to other cells in the network. In this context the circles in Fig. 1 represent “nerve cells” and the links correspond to axons linking one “nerve cell” to another. The “artificial neural network” of Fig. 1 is a multilayer perceptron because it has at least one layer between the input and the output. The circles with an I represent the inputs, the circles with an H are the “hidden” cells, so called because they are “hidden” from the observer, and circle with an O is the output. The numerical values beside the arrows in Fig. 1a, which are called the weights, represent the strength of the connection, with positive values corresponding to excitatory connections and negative values to inhibitory connections. The response of each cell is typically characterized by a nonlinear function such as a logistic function (Fig. 2). This represents schematically the response of a real nerve cell, which is nonlinear. The circles at the top in Fig. 1, denoted with a 1 in Fig. 1a and with a B in Fig. 1b, represent the bias, which is supposed to correspond to the activation threshold in a nerve cell. Obviously, a system like that represented in Fig. 1 does not rise to the level of being a caricature of the neural network of even the simplest organism. As Venables and Ripley (2002) point out, it is better to drop the biological metaphor altogether and simply think of this as a very flexible nonlinear model of the data. Nevertheless, we will conform to universal practice and use the term “artificial neural network,” abbreviated ANN, to characterize these constructs.

Figure 2. A logistic function.

In Section 2 we introduce the topic by manually constructing a multilayer perceptron (MLP) and comparing it to an MLP constructed using the nnet package (Venables and Ripley, 2002), which comes with the base R software. In Section 3 (in Part 2) we will use the neuralnet package (Fritsch et al., 2019) to develop and test an MLP with more than one hidden layer, like that in Fig. 1. Although the MLP is probably the most common ANN used for the sort of classification and regression problems discussed here, it is not the only one. In Section 4 we use the RSNNS package (Bergmeir and Benitez, 2012) to develop a radial basis function ANN. This is, for these types of application, probably the second most commonly used ANN. These first three sections apply the ANN to a binary classification problem. In Section 5 (in Part 3) we examine them in a classification problem with more than two alternatives. In Section 6 we take a look at the issue of variable selection for ANNs. Section 7 contains the Exercises and Section 8 the references. We do not discuss regression problems, but the practical aspects of this sort of application are virtually identical to those of a binary classification problem.

2. The Single Hidden Layer ANN

We will start the discussion by creating a training data set like the one that we used in our Additional Topic on the K nearest neighbor method to introduce that method. The application is the classification of oak woodlands in California using an augmented version of Data Set 2 of SDA2. In SDA2 that data set was only concerned with the presence or absence in the California landscape of blue oaks (Quercus douglasii), denoted in Data Set 2 by the variable QUDO. The original Wieslander survey on which this data set is based (Wieslander, 1935), however, contains records for other oak species as well. Here we employ an augmented version of Data Set 2, denoted Data Set 2A, that also contains records for coast live oak (Quercus agrifolia, QUAG), canyon oak (Quercus chrysolepis, QUCH), black oak (Quercus kelloggii, QUKE), valley oak (Quercus lobata, QULO), and interior live oak (Quercus wislizeni, QUWI). The data set contains records for the basal area of each species as well as its presence/absence, but we will only use the presence/absence data here. Of the 4,101 records in the data set, 2,267 contain only one species and we will analyze this subset, denoted Data Set 2U. If you want to see a color coded map of the locations, there is one here.

In this and the following two sections we will restrict ourselves to a binary classification problem: whether or not the species at the given location is QUDO. In the general description we will denote the variable to be classified by Y. As with the other Additional Topics discussing supervised classification, we will refer to the Yi, i = 1,…,n,  as the class labels. The data fields on which the classification is to be based are called the predictors and will in the general case be denoted Xi, . When we wish to distinguish the individual data fields we will write the components of the predictor Xi as Xi,j, j = 1,…,p. As mentioned in the Introduction, the application of ANNs to problems in which Y takes on continuous numerical values (i.e., to regression as opposed to classification problems) is basically the same as that described here. Because ANNs incorporate weights like those in Fig. 1a, it is important that all predictors have approximately the same numerical range. Ordinarily this is assured by normalizing them, but for ease of graphing it will be more useful for us to rescale them to the interval 0 to 1. There is little or no difference between normalizing and rescaling as far as the results are concerned (Exercise 1). Let’s take a look at the setup of the data set.

data.Set2A <- read.csv(“set2\\set2Adata.csv”,
+ header = TRUE)

> data.Set2U <- data.Set2A[which(data.Set2A$QUAG +
+   data.Set2A$QUWI +
data.Set2A$QULO +
+   data.Set2A$QUDO + data.Set2A$QUKE +
+   data.Set2A$QUCH == 1),]
> names(data.Set2U)
[1] “ID”        “Longitude” “Latitude”  “CoastDist” “MAT”     
[6] “Precip”    “JuMin”     “JuMax”     “JuMean”    “JaMin”  
[11] “JaMax”     “JaMean”    “TempR”     “GS32”      “GS28”   
[16] “PE”        “ET”        “Elevation” “Texture”   “AWCAvg” 
[21] “Permeab”   “PM100”     “PM200”     “PM300”     “PM400”  
[26] “PM500”     “PM600”     “SolRad6”   “SolRad12”  “SolRad” 
[31] “QUAG”      “QUWI”      “QULO”      “QUDO”      “QUKE”   
[36] “QUCH”      “QUAG_BA”   “QUWI_BA”   “QULO_BA”   “QUDO_BA”
[41] “QUKE_BA”   “QUCH_BA” 


Later on we are going to convert the data frame to a SpatialPointsDataFrame (Bivand et al. 2011), so we will not rescale the spatial coordinates. If you are not familiar with the SpatialPointsDataFrame concept, don’t worry about it – it is not essential to understanding what follows.

> rescale <- function(x) (x – min(x)) / (max(x) – min(x))
> for (j in 4:30) data.Set2U[,j] <- rescale(data.Set2U[,j])
> nrow(data.Set2U)
[1] 2267
> length(which(data.Set2U$QUDO == 1)) / nrow(data.Set2U)
[1] 0.3224526


The fraction of data records having the true value QUDO = 1 is about 0.32. This would be the error rate obtained by not assigning QUDO = 1 to any records and represents an error rate against which others can be compared. To simplify the explanations further we will begin by using a subset of Data Set 2U called the Training data set that consists of 25 randomly selected records each of QUDO = 0 and QUDO = 1 records.

> set.seed(5)
> oaks <- sample(which(data.Set2U$QUDO == 1), 25)
> no.oaks <- sample(which(data.Set2U$QUDO == 0), 25)
> Set2.Train <- data.Set2U[c(oaks,no.oaks),]


As usual I played with the random seed until I got a training set that gave good results.

We will begin each of the first three sections by applying the ANN to a simple “practice problem” having only two predictors (i.e., p = 2), mean annual temperature MAT and mean annual precipitation Precip. Fig. 3 shows plots of the full and training data sets in the data space of these predictors. We begin our discussion with the oldest R ANN package, nnet (Venables and Ripley, 2002), which includes the function nnet(). Venables and Ripley provide an excellent description of the package, and the source code is available here. An ANN created using nnet() is restricted to one hidden layer, although a “skip layer,” that is, a set of one or more links directly from the input to the output, is also allowed. A number of options are available, and we will begin with the most basic. We will start with a very simple call to nnet(), accepting the default value wherever possible. We will set the random number seed before each run of any ANN. The algorithm begins with randomly selected initial values, and this way we achieve repeatable results. Again, your results will probably differ from mine even if you use the same seed.

                  
                                     (a)                                       (b)

Figure 3. (a) Plot of the full data set, (b) plot of the training data set.

> library(nnet)
> set.seed(3)
> mod.nnet <- nnet(QUDO ~ MAT + Precip,
+   data = Set2.Train, size = 4)

# weights:  17
initial  value 12.644382
iter  10 value 8.489889
      *      *      *
iter 100 value 6.158126
final  value 6.158126
stopped after 100 iterations
 
If you are not familiar with the R formula notation QUDO ~ MAT + Precip see here. As already stated, nnet() allows only one hidden layer, and the argument size = 4 specifies four “cells” in this layer. From now on to avoid being pedantic we will drop the quotation marks around the word cell. The output tells us that there are 17 weights to iteratively compute (we will see in a bit where this figure comes from), and nnet() starts with a default maximum value of 100 iterations, after which the algorithm has failed to converge. Let’s try increasing the value of the maximum number of iterations as specified by the argument maxit.

> set.seed(3)
> mod.nnet <- nnet(QUDO ~ MAT + Precip,
+   data = Set2.Train,
size = 4, maxit = 10000)
# weights:  17
initial  value 12.644382
iter  10 value 8.489889
     *      *       *
iter 170 value 6.117850
final  value 6.117849
converged
 
This time it converges. Sometimes it does and sometimes it doesn’t, and again your results may be different from mine.

                         

Fig. 4. Schematic diagram of an ANN with a single hidden layer having four cells.


Fig. 4 shows a plot of the ANN made using the plotnet() function. Again, the thickness of the lines indicates the strength of the corresponding link, and the color indicates the direction, with black indicating positive and gray indicating negative. Thus, for example, MAT has a strong positive effect on hidden cell 3 and Precip has a strong negative effect. The function nnet() returns a lot of information, some of which we will explore later. The nnet package includes a predict() function similar to the ones encountered in packages covered in other Additional Topics. We will use it to generate values of a test data set. Unlike other Additional Topics, the test data set here will not be drawn from the QUDO data. Instead, we will cover the data space uniformly so that we can see the boundaries of the predicted classes and compare them with the values of the training data set. As is usually true, there is an R function available that can be called to carry out this operation (in this case it is expand.grid()). However, I do not see any advantage in using functions like this to perform tasks that can be accomplished more transparently with a few lines of code.

> n <- 50
> MAT.test <- rep((1:n), n) / n
> Precip.test <- sort(rep((1:n), n) / n)     
> Set2.Test <- data.frame(MAT = MAT.test,
+   Precip = Precip.test)

> Predict.nnet <- predict(mod.nnet, Set2.Test)

Let’s first take a look at the structure of Predict.nnet.

> str(Predict.nnet)
 num [1:2500, 1] 0 0 0 0 0 0 0 0 0 0 …
 – attr(*, “dimnames”)=List of 2
  ..$ : chr [1:2500] “1” “2” “3” “4” …
  ..$ : NULL
  It is a 2500 x 1 array, with rows named 1 through 2500. We can examine the range of values by constructing a histogram (Fig. 5)




Figure 5. Histogram of values of Predict.nnet.

This shows a range of values between 0 and 1. Those values above a certain threshold will be interpreted as QUDO = 1. With more complex models we will choose the threshold by constructing a receiver operating characteristic curve (i.e., a ROC curve), but for this simple situation it is pretty clear that the value one half will suffice. Since we will be constructing many plots of a similar nature, we will develop a function to carry out this process.

> plot.ANN <- function(Data, Pred, Thresh,
+      header.str){

+   Data$Color <- “red”
+   Data$Color[which(Pred < Thresh)] <- “green”
+   with(Data, plot(x = MAT, y = Precip, pch = 16,
+     col = Color,
cex = 0.5, main = header.str))
+   }  

Next we apply the function.

> plot.ANN(Set2.Test, Predict.nnet, 0.5, “nnet
+     Predictions”) # Fig. 6

> with(Set2.Train, points(x = MAT, y = Precip,
+    pch = 16, col = Color))




Figure 6. A “bestiary” of predictions delivered by the function nnet(). Most of the time the function returns “boring” predictions like the first and fourth, but occasionally an interesting one is produced.

If we run the code sequence several times, each time using a different random number seed, each run generates a different set of starting weights and returns a potentially different prediction. Fig. 6 shows some of them. Why does this happen? There are a few possibilities. The plots in Fig. 6 represent the results of the solution of an unconstrained nonlinear minimization problem (we will delve into this problem more deeply in a bit). Solving a problem such as this is analogous to trying to find the lowest point in hilly terrain in a dense fog. Moreover, this search is taking place in the seventeen-dimensional space of the biases and weights. It is possible that some of the solutions represent local minima, and some may also arise if the area around the global minimum is relatively flat so that convergence is declared before reaching the true minimum. It is, of course, also possible that the true global minimum gives rise to an oddly shaped prediction. Exercise 3 explores this issue a bit more deeply.

Now that we have been introduced to some of the concepts, let’s take a deeper look at how a multilayer perceptron ANN works. Since each individual cell is a logistic response function, we can start by simply fitting a logistic curve to the data.  Logistic regression is discussed in SDA Sec. 8.4. The logistic regression model is generally written as (cf. SDA2 Equation 8.20)

                                   , (1)
where is the logistic regression estimator of . Based on the discussion in SDA2 we will use the function glm() of the base package to determine the logistic model for our training data set.

> mod.glm <- glm(QUDO ~ MAT + Precip,
+    data = Set2.Train, family = binomial)
> Predict.glm <- predict(mod.glm, Set2.Test)
> p <- with(Set2.Train, c(MAT, Precip, QUDO))
 
The results are plotted as a part of Fig. 7 below after first switching to a perspective more in line with ANN applications. Specifically, the function glm() computes a logistic regression model based not on minimizing the error sum of squares but on a maximum likelihood procedure. As a transition from logistic regression to ANNs we will compute a least squares solution to Equation (1) using the R base function nlm(). In this case, rather than being interpreted as a probability as in Equation (1), the logistic function is interpreted as the response of a cell such as those shown in Fig. 1. We will change the form of the logistic function to
                                               . (2)
Note that as Z increases, approaches 1, and as decreases toward minus infinity,  approaches 0. This form of the function is different from Equation (1) but it is an easy exercise to show that they are equivalent. The form of the function in Equation (2) follows that of Venables and Ripley (2002, p. 224) and is more common in the ANN literature.

We will write the R code for the logistic function (this was used to plot Fig. 2) as follows.

> # Logistic function
> lg.fn <- function(coefs, a){
+    b <- coefs[1]
+    w <- coefs[2:length(coefs)]
+    return(1 / (1 + exp(-(b + sum(w * a)))))
+    }

Here the b corresponds to the bias terms B and the w corresponds to the weights terms W in Fig. 1b. The a corresponds to the level of activation that the cell receives as input. Thus each cell receives an input activation a and produces an output activation . Next we create a function SSE() to compute the sum of squared errors. For simplicity of code we will restrict our models to the case of two predictors (i.e., p = 2).

> SSE <- function(w.io, X.Y, fn){
+    n <- length(X.Y) / 3
+    X1 <- X.Y[1:n]
+    X2 <- X.Y[(n+1):(2n)]
+    Y <- X.Y[(2
n+1):(3*n)]
+    Yhat <- numeric(n)
+    for (i in 1:n) Yhat[i] <- fn(w.io,
+       c(X1[i], X2[i]))

+    sse <- 0
+    for (i in 1:n) sse <- sse + (Y[i] – Yhat[i])^2
+    return(sse)
+    }

Here the argument w.io contains the biases and weights (the b and w terms), which are the nonlinear regression coefficients, X.Y contains the Xi and Yi terms, and fn is the function to compute (in this case lg.fn()). Here is the code to compute the least squares regression using nlm(), compare its coefficients and error sum of squares with that obtained from glm(), and draw Fig. 7. As with nnet() above, the iteration in nlm() is started with a set of random numbers drawn from a unit normal distribution.

> print(coefs.glm <- coef(mod.glm))
(Intercept)       MAT    Precip
 -0.7704258   3.1803868  -5.3868261
> SSE(coefs.glm, p, lg.fn)
[1] 8.931932
> print(coefs.nlm <- nlm(SSE, rnorm(3), p,
+      lg.fn)$estimate)

[1] -0.9089667  3.3558542 -4.9486441
> SSE(coefs.nlm, p, lg.fn)
[1] 8.90217
> plot.ANN(Set2.Test, Predict.glm, 0.5, “Logistic
+     Regression”) # Fig. 7

> with(Set2.Train, points(x = MAT, y = Precip,
+  pch = 16,
col = Color))  
> c <- coefs.nlm
> abline(-c[1]/c[3], -c[2]/c[3])




Figure 7. Prediction zones computed using glm(). The black line shows the boundary of the prediction zones computed by minimizing the error sum of squares using nlm().

Now that we have a function to compute the logistic function and one to minimize the sum of squares, we are ready to construct a simple ANN with one hidden layer.  

Table 1. Notation used in the single hidden cell ANN.
Subscript or superscript Significance Range symbol Max value
( ) Layer I, H, M (1), (2), (3) (3)
i Data record n 50
j Predictor p 2
m Hidden cell number M 4
  Table 1 shows the notation that we will use. There are p input cells at layer 1, numbered 1,…,p. In our case, p = 2. Each input cell receives an input activation pair Xi,j, j = 1,…,n. For the training data, n = 50. Fig. 8 shows a schematic of an ANN in which there is one hidden layer with four cells. As shown in the figure, the layers of cells are numbered in increasing order, so that the input layer is layer 1, the hidden layer is layer 2, and the output layer is layer 3. The layers are indicated by superscripts in parentheses. The connections between input cells and the hidden cells pass the weighted signal to the hidden layer cells. In Fig. 8 for each data record i each cell receives a signal consisting of the pair Xi,j, j = 1, 2, and passes it along to the hidden layer cells as an activation . Here we have to be a bit careful. The superscript (1) refers to the layer, i indexes the data record, which ranges from 1 to 50, j refers to the component of the predictor variable and has the value 1 or 2, and m indexes the hidden layer cell number and ranges from 1 to M = 4. Figure 8. Diagram of an ANN with one hidden layer having four cells.

  The subscripts in Fig. 8 are the reverse of what one might expect in that information flows from input cell j in layer 1, the input layer, to cell m in layer 2, the hidden layer (for example, we write   instead of ; here j =2 identifies the input cell and m = 4 identifies the hidden layer cell). This seemingly backwards notation is traditional in the ANN literature and is used because the set of values is often represented as a matrix in which the values in the same layer form the rows. We will encounter this usage in Section 6. The activation   is part of the information processed by hidden cell m in layer 2. The hidden cells respond to the sum of input activations. according to the equation
                            . (3)
This action takes place for each of the  data records. The one output cell (layer 3) receives the activation values  of each of the M hidden cells and processes these according to the equation
                 . (4)
The quantity Qi represents the estimate of Yi given the current set of weights. Assuming that the sum of squared errors is used as the objective function, the value of the objective function

(5)
is computed. This quantity is then minimized to yield the final estimate . Table 2 summarizes this sequence of steps.  
Table 2 Sequence of steps in the computation of the ANN weights and biases.

We are now ready to start constructing our homemade ANN. First let’s see how many weights we will need. Fig. 8 indicates that if there are M hidden cells and two predictors then there will be 2M weights to cover the links from the input, plus M for the bias , for a total of 3M. From the hidden cells to the output cell there will be M, plus a final weight for the bias  for a grand total of 4M + 1 weights. Recall that the function nnet() reported 17 weights; this is where that number came from. We must pass all of the weights through the code as an array. Table 3 shows how the code of our homemade ANN arranges this array for an ANN with four hidden cells. The arrangement is organized to make the programming as simple as possible. Note, by the way, that the bias  is applied to layer 2 and the bias  is applied to layer 3. This seems confusing, but again is be nearly universal practice, so I will go with it as well. Remember that  is the weight applied to the activation from the  cell in layer l applied to the  cell in layer l + 1.

Table 3. Arrangement of the 17 biases and weights for an ANN with four hidden cells.
We will first construct a prediction function pred.ann() that, given a set of weights  and a pair of input values  for a fixed i calculates an estimate according to Equation (3).

> pred.ANN1 <- function(w.io, X){
+     b.2 <- w.io[1] # First beta is output bias
+     b.w <- w.io[-1] # Biases and weights
+     M <- length(b.w) / 4 # Number of hidden cells
# vector to hold hidden cell output activation
+     a.2 <- numeric(M)
# Hidden cell biases and weights

+     w.2 <- b.w[(3 * M + 1):(4 * M)]
+     for (m in 1:M){
+        bw.1 <-
+           b.w[(1 + (m – 1) * 3):(3 + (m – 1) * 3)]

+        a.2[m] <- lg.fn(bw.1, X)
+        }
+     phi.out <- lg.fn(c(b.2, w.2), a.2)
+     return(phi.out)
+     }
}

The arguments w.io and X contain the biases and weights and the predictor values X, respectively. The first two lines extract the output cell bias b.2, which corresponds to in Table 3, and the third lines determines the number of hidden cells by dividing the number of weights by 4. The next three lines establish locations to hold the values of the quantities in Table 3. The remaining lines of code compute the quantity Qi in Equation (4).

The next function generates the ANN itself, which we call ANN.1().

> ANN.1 <- function(X, Y, M, fn, niter){
+    w.io <- rnorm(4*M + 1)
+    X.Y <- c(as.vector(X), Y)
+    return(nlm(SSE, w.io, X.Y, fn, iterlim = niter))
+    } 

The arguments are respectively X, a matrix or data frame whose columns are the Xi,j, Y, the vector of Y values, M, the number of hidden cells, fn, the function defining the cells’ response, and niter, the maximum allowable number of iterations. The first line in the function generates the random starting weights and the second line arranges the data as a single column. The function then calls nlm() to carry out the nonlinear minimization of the sum of squared errors. There are other error measures beside the sum of squared errors that we could use, and some will be discussed later. Now we must set the quantities to feed to the ANN and pass them along.

> set.seed(1)
> X <- with(Set2.Train, (c(MAT, Precip)))
> mod.ANN1 <- ANN.1(X, Set2.Train$QUDO, 4, pred.ANN,
+    1000)
 
The function ANN.1() returns the full output of nlm(), which includes the coefficients as well as a code indicating the conditions under which the function terminated. There are five possible values of this code; you can use ?nlm to see what they signify. One point, however, must be noted. The function nlm() makes use of Newton’s method to estimate the minimum of a function. The specific algorithm is described by Schnabel et al. (1985). The important point is that the method makes use of the Hessian matrix, which is the matrix of second derivatives. If there are thousands of weights, as there might be in an ANN, this is an enormous matrix. For this reason, newer ANN algorithms generally use the gradient descent method, for which only first derivatives are needed. We will return to this issue in Section 3. Here is the output from ANN.1(). Again, your output will probably be different from mine.

> print(beta.ann <- mod.ANN1$estimate)
[1]  191.48348  198.97750 -407.10859  269.84118 -824.73518
[6]  898.75456  826.91247  473.50764 -469.37435 -241.65738
[11]  443.55838 -201.72335 -149.85410 -188.09470 -168.33946
[16]  -96.12398   73.12316
> print(code.ann <- mod.ANN1$code)

[1] 3  
The code value 3 indicates that at least a local minimum was probably found. The output is plotted in Fig. 9.

> Set2.Test$QUDO <- 0
> for (i in 1:length(Set2.Test$QUDO))
+    Set2.Test$QUDO[i] <- pred.ANN1(beta.ann,
+        c(Set2.Test$MAT[i],
Set2.Test$Precip[i]))
> plot.ANN(Set2.Test, Set2.Test$QUDO, 0.5,
+    “ANN.1 Predictions”)

> with(Set2.Train, points(x = MAT, y = Precip,
+    pch = 16,
col = Color)) # Fig. 9   
Similar to Fig. 6, this figure shows a “bestiary” of predictions of this artificial neural network. This was created by sequencing the random number seed through integer values starting at 1 and picking out results that looked interesting.

  Figure 9. A “bestiary” of ANN.1() predictions. As with nnet(), most of the time the output is relatively tame, but sometimes it is not.  

In order to examine more closely the workings of the ANN we will focus on the lower middle prediction in Fig. 9, which was created using the random number seed set to 7, and examine this for a fixed value of the predictor Precip. Fig. 10 plots the ANN output with the set of values Precip = 0.8 highlighted.



Figure 10. Replot of the sixth output displayed in Figure 8 with values Precip = 0.8 highlighted.

 

Figure 11. Activation  of each individual hidden cell.

Fig. 11 shows the hidden cell activation  together with the value of the weight  multiplying this response in Equation (3). Note that the individual cells can be said to “organize” themselves to respond in different ways to the inputs. For this reason during the heady days of early work in ANNs systems like this were sometimes called “self-organizing systems.” Also, the first two cells appear to have similar response, although, as can be seen by inspecting the bias and weight values in the output of ANN.1() above (with reference to Table 1), these values are actually very different. Fig. 12 is a plot along the highlighted line Precip = 0.8 of the same output as displayed in Fig. 11.   Figure 12. Plot of the hidden cell activation   along the highlighted line in Fig. 10.   Figure 13 shows the weighted activation values  of the hidden cells along the row of highlighted values. Note the different abscissa ranges of the plots in Figs. 12 and 13. Figure 13. Plot of the weighted hidden cell activation  along the highlighted line in Fig. 10. Fig. 14a shows the sum passed to the output cell, given by  in Equation (3), and Fig. 14b shows the sum  when the bias  is added. Finally, Fig. 14c shows the output value .
               (a)                                       (b)                                       (c) Figure 14. Referring to Equation (3): (a) sum  of inputs to the output cell; (b) bias added to sum of weighted inputs, ; (c) final output. The values of the first cell are universally low, so it apparently plays almost no role in the final result, we might be able achieve some simplification by pruning it. Let’s check it out.

> Set2.Test$QUDO1 <- 0
> beta.ann1 <- beta.ann[c(1,5:16)]
> for (i in 1:length(Set2.Test$QUDO))
+    Set2.Test$QUDO1[i] <- pred.ANN1(beta.ann1, c(Set2.Test$MAT[i],
+         Set2.Test$Precip[i]))
> max(abs(Set2.Test$QUDO – Set2.Test$QUDO1))  
[1] 1


Obviously, cell 1 does play an important role for some values of the predictors. Sometimes a cell can be pruned in this way without changing the results and sometimes not. Algorithms do exist for determining whether individual cells in an ANN can be pruned (Bishop, 1995). We will see in Section 6 that examination of the contributions of cells can also be used in variable selection. We mentioned earlier that the function nlm(), used in our function SSE() to locate the minimum of the sum of squared errors, could be consumptive of time and memory for large problems. Ripley (1996, p. 158), writing at a time when computers were much slower than today, asserted that any ANN with fewer than around 1,000 weights could be solved effectively using Newton’s method. The function nnet() uses an approximation of Newton’s method called the BGFS method (see ?nnet and ?optim). We can to some extent test Newton’s method by expanding the size of our ANN to ten hidden cells and running it on the full data set of 2267 records.

> set.seed(3)
> X <- with(data.Set2U, (c(MAT, Precip)))
> start_time <- Sys.time()
> mod.ANN1 <- ANN.1(X, data.Set2U$QUDO, 10,
+     pred.ANN1, 1000)

> print(code.ann <- mod.ANN1$code)
[1] 1
> stop_time <- Sys.time()
> stop_time – start_time
Time difference of 9.562427 secs


The return code of 1 indicates iteration to convergence. When using other random seeds the program occasionally goes down a rabbit hole and has to be manually stopped by hitting the <esc> key. When it does converge, it usually doesn’t take too long. We now return to the nnet package and the nnet() function, which we will apply to the entire data set. First, however, we need to discuss the use of this function for classification as opposed to regression. Our use of the numerical values 0 and 1 for the class label QUDO makes nnet() treat this as a regression problem. We can change to a classification problem easily by defining a factor valued data field and using this in the nnet() model. This seemingly small change, however, gives us the opportunity to look more closely at how the ANN computes its values and how error is interpreted. As with logistic regression, when QUDO is an integer quantity, the estimate returns values on the interval (0,1) (i.e., all numbers greater than 0 and less than 1). The use of nnet() in classification is discussed by Venables and Ripley (2002, p. 342). We will define a new quantity QUDOF as a factor taking on the factor values “0” and “1”.


> data.Set2U$QUDOF <- as.factor(data.Set2U$QUDO)
> unique(data.Set2U$QUDOF)
[1] 1 0
Levels: 0 1
modf.nnet <- nnet(QUDOF ~ MAT + Precip,
+    data = data.Set2U,
size = 4, maxit = 1000)
Let’s take a closer look at the predicted values

> hist(Predict.nnet, breaks = seq(0, 1, 0.1),
+   main = “Histogram of nnet Predictions”) #Fig. 15

Figure 15.  Histogram of predicted values of the function nnet().   As shown in Fig. 15, although QUDOF is a nominal scale quantity (see SDA2, Sec. 4.2.1), the function nnet() still returns values on the interval (0,1). Unlike the case with logistic regression, these cannot be interpreted as probabilities, but we can interpret the output as an indication of the “strength” of the classification. Since the returned values are distributed all along the range from 0 to 1, it makes sense to establish the cutoff point using a ROC curve. We will use the functions of the package ROCR to do so.

> library(ROCR)
> pred.QUDO <- predict(modf.nnet, data.Set2U)
> pred <- prediction(pred.QUDO, data.Set2U$QUDO)
> perf <- performance(pred, “tpr”, “fpr”)
> plot(perf, colorize = TRUE,
+      main = “nnet ROC Curve”) # Fig. 16a
> perf2 <- performance(pred, “acc”)

> plot(perf2, main =
+   “nnet Prediction Accuracy”)  # Fig 16b
> print(best.cut <- which.max([email protected][[1]]))

[1] 565
> print(cutoff <- [email protected][[1]][best.cut])

[1] 0.4547653
                                   (a)                                                  (b)
Figure 16. (a) ROC curve for the nnet() prediction of QUDO presence/absence; (b) plot of prediction accuracy against cutoff value. The ROC curve analysis indicates that the cutoff value leading to the highest prediction accuracy is around 0.45. Fig. 17a shows a replot of the full data set plotted in Fig. 3a and Fig. 17b shows a plot of the prediction regions.                   
                                      (a)                                              (b) Figure 17.  (a) Plot of the QUDO data; (b) plot of the prediction regions. The ROC curve can be used to compute a widely used performance measure, the area under the ROC curve, commonly known as the AUC. Looking at Fig. 16a, one can see that if the predictor provides a perfect prediction (zero false positives or false negatives) then the AUC will be one, and if the predictor is no better than flipping a coin then the AUC will be one half. The Wikipedia article on the ROC curve provides a thorough discussion of the AUC. Here is the calculation for the current case.

> auc.perf <- performance(pred,”auc”)
> print(auc <- as.numeric([email protected]))
[1] 0.8872358

 Returning to the nnet() function, another feature of this function when the class labels are factors is its default error measure (this is discussed in the Details section of ?nnet). We can see this by applying the str() function to the nnet() output.

> str(modf.nnet)
List of 19
  *    *    Deleted    *    *
$ entropy      : logi TRUE
  *    *    Deleted    *    *

The argument entropy is by default set to TRUE. If you try this with the nnet object created earlier with QUDO taking on integer values, you will see that the default value of entropy in this case is FALSE (the value of entropy can be controlled in nnet() via the entropy argument). When entropy is TRUE, the prediction error is not measured by the sum of squared errors as defined in Equation (5). Instead, it is measured by the cross entropy. For a factor valued quantity we can define this as (Guenther and Frisch, 2010, Venables and Ripley, 2002, p. 245)
                                                 , (6)
where ti is 1 for those values of Qi that do not match the corresponding Yi, and 0 for those that do, and ai is a measure of the activity of the output cell. There is some evidence (e.g., Du and Swamy 2014, p. 37) that this measure avoids the creation of “flat spots” in the surface (or hypersurface) over which the optimization is being computed. Fig. 18 gives a sense of the difference. We will meet the cross entropy again in Section 3.

                           Figure 18. Plots of the sum of squared errors and the cross entropy for a range of error measures between 0 and 1. A potentially important issue in the development of an ANN is the number of cells. The tune() function of the e1071 package (Meyer et al., 2019), discussed in the Additional Topic on Support Vector Machines, has an implementation tune.nnet(). The primary intended use of the function is for cross-validation, which will be discussed below. The function can, however, also be used to rapidly assess the effect of increasing the number of hidden cells.

> library(e1071)
> tune.nnet(QUDOF ~ MAT + Precip, data = data.Set2U,
+    size = 4)
Error estimation of ‘nnet’ using 10-fold cross validation: 0.2056282
Time difference of 9.346614 secs
> tune.nnet(QUDOF ~ MAT + Precip, data = data.Set2U,

+    size = 8) Error estimation of ‘nnet’ using 10-fold cross validation: 0.2024174
Time difference of 15.9804 secs
> tune.nnet(QUDOF ~ MAT + Precip, data = data.Set2U,

+    size = 12)
Error estimation of ‘nnet’ using 10-fold cross validation: 0.1943285
Time difference of 22.1304 secs

 The code to calculate the elapsed time is not shown. These results provide evidence that for this particular problem there is only a very limited advantage to increasing the number of cells. We can use repeated evaluation to pick the prediction that provides the best fit. We will try twenty different random starting configurations to see the distribution of error rates (keeping the same cutoff value for all of them).

> seeds <- numeric(20)
> error.rate <- numeric(length(seeds))
> conv <- numeric(length(seeds))
> seed <- 0
> i <- 1
> data.set <- Set2.Test
> while (i <= length(seeds)){
+    seed <- seed + 1
+    set.seed(seed)
+    md <- nnet(QUDO ~ MAT + Precip,
+      data = data.Set2U,
size = 4, maxit = 1000,
+      trace = FALSE)

+    if (md$convergence == 0){
+       seeds[i] <- seed
+       pred <- predict(md)
+       QP <- numeric(length(pred))
+       QP[which(pred > cutoff)] <- as.factor(1)
+       error.rate[i] <-
+          length(which(QP != data.Set2U$QUDOF)) /
+              nrow(data.Set2U)

+       i <- i + 1
+       }
+    }
> seeds
[1]  1  3  4  5  6  7  8  9 10 11 12 13 14 15 16 17 18 19 20 21
> sort(error.rate)

[1] 0.1782091 0.1804146 0.1817380 0.1821791 0.1830613 0.1839435 0.1843846
[8] 0.1843846 0.1848258 0.1857080 0.1865902 0.1874724 0.1887958 0.1905602
[15] 0.1932069 0.1954124 0.1971769 0.2015880 0.2037936 0.2042347
> which.min(error.rate)

[1] 17   All but one of the seeds converged, and the best error rate is obtained using the seventeenth random number seed. The error rate here is not the same as the cross-validation error rate calculated above, as reflected in the different values. A plot of the predictions using this seed is similar to that shown in Fig. 16b. Looking at that figure, one might suspect that the model may overfit the data. The concept of overfitting and underfitting in terms of the bias-variance tradeoff is discussed in a linear regression context in SDA2 Section 8.2.1 and in a classification context in the Additional Topic on Support Vector Machines. In summary, a model that fits a training data set too closely (overfits) may do a worse job of fitting a different data set and, as discussed in the two sources, has a high variance. A simpler model that does not fit the training data as well has a higher bias, and the decision of which model to choose is a matter of balancing these two factors. Ten-fold cross validation is commonly used to test where a model stands the bias-variance tradeoff (again see the two sources given above). The following code carries out such a cross-validation.

> best.seed <- seeds[which.min(error.rate)]
> ntest <- trunc(nrow(data.Set2U) / 10) * 10
> data.set <- data.Set2U[1:ntest,]
> n.rand <- sample(1:nrow(data.set))
> n <- nrow(data.set) / 10
> folds <- matrix(n.rand, nrow = n, ncol = 10)
> SSE <- 0
> for (i in 1:10){
+    test.fold <- i
+    if (i == 1) train.fold <- 2:10
+    if (i == 10) train.fold <- 1:9
+    if ((i > 1) & (i < 10)) train.fold <-
+         c(1:(i-1), (i+1):10)

+    train <-
+       data.frame(data.set[folds[,train.fold],])

+    test <-
+         data.frame(data.set[folds[,test.fold],])

+    true.sp <- data.set$QUDO[folds[,test.fold]]
+    set.seed(best.seed)
+    modf.nnet <- nnet(QUDOF ~ MAT + Precip,
+       data = train, size = 8, maxit = 10000,
+           Trace = FALSE)

+    pred.1 <- predict(modf.nnet, test)
+    test$QUDO <- 0
+    test$QUDO[which(pred.1 >= 0.52)] <- 1
+    E <- length(which(test$QUDO != true.sp)) /
+        length(true.sp)

+    SSE <- SSE + E^2
+    }
> print(RMSE <- sqrt(SSE / 10))
[1] 0.1930847   The RMSE seems aligned with the error rate. Venables and Ripley (2002) and Ripley (1996) discuss some of the options available with the function nnet(). Some of these are explored in Exercises (4) and (5). Now that we have some experience with an ANN having a single hidden layer, we will move on to the package neuralnet, in which multiple hidden layers are possible. This concludes the first installment. Part 2 will be published subsequently. Again, R code and data are available in the Additional Topics section of SDA2, http://psfaculty.plantsciences.ucdavis.edu/plant/sda2.htm  

Numerical Partial Derivative Estimation – the {NNS} package

NNS (v0.5.5) now on CRAN has an updated partial derivative routine dy.d_() . This function estimates true average partial derivatives, as well as ceteris paribus conditions for points of interest.

Example below on the syntax for estimating first derivatives of the function y = x_1^2 * x_2^2 , for the points x_1 = 0.5 and x_2 = 0.5, and for both regressors x_1 and x_2.

set.seed(123)
x_1 = runif(1000)
x_2 = runif(1000)
y = x_1 ^ 2 * x_2 ^ 2
dy.d_(cbind(x_1, x_2), y, wrt = 1:2, eval.points = t(c(.5,.5)))["First",]
[[1]]
[1] 0.2454744

[[2]]
[1] 0.2439307


The analytical solution for both regressors at x_1 = x_2 = 0.5 is 0.25.

The referenced paper gives many more examples, comparing dy.d_() to kernel regression gradients and OLS coefficients.

For even more NNS capabilities, check out the examples at GitHub:
https://github.com/OVVO-Financial/NNS/blob/NNS-Beta-Version/examples/index.md


Reference Paper:
Vinod, Hrishikesh D. and Viole, Fred, Comparing Old and New Partial Derivative Estimates from Nonlinear Nonparametric Regressions
https://ssrn.com/abstract=3681104 

Supplemental Materials:
https://ssrn.com/abstract=3681436


Most popular on Netflix. Daily Tops for last 60 days

Everyday, around 9 pm, I get fresh portion of the Netflix Top movies / TV shows. I’ve been doing this for more than two months and decided to show the first results answering following questions:

How many movies / TV shows make the Top?
What movie was the longest #1 on Netflix?
What movie was the longest in Tops?
For how many days movies / TV shows stay in Tops and as #1?
To have a try to plot all this up and down zigzags…

I took 60 days span (almost all harvest so far) and Top5 overall, not Top10, in each category to talk about really the most popular and trendy.

So, let`s count how many movies made the Top5, I mean it is definitely less than 5 *60…
library(tidyverse)
library (stats)
library (gt)
fjune <- read_csv("R/movies/fjune.csv")
#wrangle raw data - reverse (fresh date first), take top 5, take last 60 days
fjune_dt %>% rev () %>% slice (1:5) %>% select (1:60)

#gather it together and count the number of unique titles 
fjune_dt_gathered <- gather (fjune_dt)
colnames (fjune_dt_gathered) <- c("date", "title")
unique_fjune_gathered <- unique (fjune_dt_gathered$title)

str (unique_fjune_gathered)
#chr [1:123] "Unknown Origins" "The Debt Collector 2" "Lucifer" "Casino Royale"
OK, it is 123 out of 300.
Now, it is good to have distribution in each #1 #2 #3 etc.
t_fjune_dt <- as.data.frame (t(fjune_dt),stringsAsFactors = FALSE)
# list of unique titles in each #1 #2 #3etc. for each day
unique_fjune <- sapply (t_fjune_dt, unique)

# number of unique titles in each #1 #2 #3 etc.
n_unique_fjune <- sapply (unique_fjune, length)
n_unique_fjune <- setNames (n_unique_fjune, c("Top1", "Top2", "Top3", "Top4", "Top5"))

n_unique_fjune
Top1 Top2 Top3 Top4 Top5
32   45   45   52   49
What movie was the longest in Tops / #1?
# Top5
table_fjune_dt5 <- sort (table (fjune_dt_gathered$title), decreasing = T)
# Top1
table_fjune_dt1 <- sort (table (t_fjune_dt$V1), decreasing = T)

# plotting the results
bb5 <- barplot (table_fjune_dt5 [1:5], ylim=c(0,22), main = "Days in Top5", las = 1, col = 'light blue')
text(bb5,table_fjune_dt5 [1:5] +1.2,labels=as.character(table_fjune_dt5 [1:5]))
bb1 <- barplot (table_fjune_dt1 [1:5], main = "Days as Top1", las = 1, ylim=c(0,6), col = 'light green')
text(bb1,table_fjune_dt1 [1:5] +0.5, labels=as.character(table_fjune_dt1 [1:5]))


Let`s weight and rank (1st = 5, 2nd = 4, 3rd = 3 etc.) Top 5 movies / TV shows during last 60 days.

i<- (5:1)
fjune_dt_gathered5 <- cbind (fjune_dt_gathered, i) 
fjune_dt_weighted % group_by(title) %>% summarise(sum = sum (i)) %>% arrange (desc(sum))
top_n (fjune_dt_weighted, 10) %>% gt () 


As we see the longer movies stays in Top the better rank they have with simple weighting meaning “lonely stars”, which got the most #1, draw less top attention through the time span.

Average days in top
av_fjune_dt5 <- round (nrow (fjune_dt_gathered) / length (unique_fjune_gathered),1) # in Top5
av_fjune_dt1 <- round (nrow (t_fjune_dt) / length (unique_fjune$V1),1) #as Top1

cat("Average days in top5: ", av_fjune_dt5, "\n") 
cat("Average days as top1: ", av_fjune_dt1)
#Average days in top5: 2.4
#Average days as top1: 1.9 
Average days in top distribution
as5 <- as.data.frame (table_fjune_dt5, stringsAsFActrors=FALSE) 

as1 <- as.data.frame (table_fjune_dt1, stringsAsFActrors=FALSE) 
par (mfcol = c(1,2)) 
boxplot (as5$Freq, ylim=c(0,7), main = "Days in Top5") 
boxplot (as1$Freq, ylim=c(0,7), main = "Days as Top1")

And now, let`s try to plot Top5 movies / TV shows daily changes.
I played with different number of days and picked 4 – with longer span I had less readable plots since every new day brings new comers into the Top with its own line, position in the legend, and, ideally its own color.

fjune_dt_gathered55 <- fjune_dt_gathered5 %>% slice (1:20) %>% rev ()
ggplot(fjune_dt_gathered55)+ aes (x=date, y=i, group = title) + geom_line (aes (color = title), size = 2) +
geom_point (aes (color = title), size = 5)+ theme_classic()+
theme(legend.position="right")+ylab("top")+ ylim ("5", "4", "3", "2", "1")+
geom_line(aes(colour = title), arrow = arrow(angle = 15, ends = "last", type = "closed"))+scale_color_brewer(palette="Paired")