forked from duvenaud/phd-thesis
-
Notifications
You must be signed in to change notification settings - Fork 0
/
additive.tex
699 lines (557 loc) · 44.2 KB
/
additive.tex
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
\input{common/header.tex}
\inbpdocument
%\chapter{Dropout in Gaussian processes}
\chapter{Additive Gaussian Processes}
\label{ch:additive}
%\begin{quotation}
%``I am in [Gaussian processes] stepped in so far that, should I wade no more, Returning were as tedious as go o'er.''
%\hspace*{\fill}--MacBeth
%\end{quotation}
\Cref{ch:grammar} showed how to learn the structure of a kernel by building it up piece-by-piece.
This chapter presents an alternative approach: starting with many different types of structure in a kernel, adjusting kernel parameters to discard whatever structure is \emph{not} present in the current dataset.
The advantage of this approach is that we do not need to run an expensive discrete-and-continuous search in order to build a structured model, and implementation is simpler.
This model, which we call \emph{additive Gaussian processes}, is a sum of functions of all possible combinations of input variables.
This model can be specified by a weighted sum of all possible products of one-dimensional kernels.
There are $2^D$ combinations of $D$ objects, so na\"{i}ve computation of this kernel is intractable.
Furthermore, if each term has different kernel parameters, fitting or integrating over so many parameters is difficult.
To address these problems, we introduce a restricted parameterization of the kernel which allows efficient evaluation of all interaction terms, while still allowing a different weighting of each order of interaction.
Empirically, this model has good predictive performance in regression tasks, and its parameters are relatively interpretable.
This model also has an interpretation as an approximation to \emph{dropout}, a recently-introduced regularization method for neural networks.
The work in this chapter was done in collaboration with Hannes Nickisch and Carl Rasmussen, who derived and coded up the additive kernel.
My role in the project was to examine the properties of the resulting model, clarify the connections to existing methods, to create all figures and run all experiments.
That work was published in \citet{duvenaud2011additive11}.
The connection to dropout regularization in \cref{sec:dropout-gps} is an independent contribution.
\section{Different types of multivariate additive {\mbox structure}}
\Cref{ch:kernels} showed how additive structure in a \gp{} prior enabled extrapolation in multivariate regression problems.
In general, models of the form
%
\begin{align}
f(\vx) = g \big( f(x_1) + f(x_2) + \dots + f(x_D) \big)
\label{eq:general-additive}
\end{align}
%
are widely used in machine learning and statistics, partly for this reason, and partly because they are relatively easy to fit and interpret.
Examples include logistic regression, linear regression, generalized linear models~\citep{nelder1972generalized} and generalized additive models~\citep{hastie1990generalized}.
%, are typically easy to fit and interpret.
%Some extensions of this family, such as smoothing-splines ANOVA \citep{wahba1990spline}, also consider terms depending on more than one variable.
%However, such models often become intractable and difficult to fit as the number of terms increases.
At the other end of the spectrum are models which allow the response to depend on all input variables simultaneously, without any additive decomposition:
%
\begin{align}
f(\vx) = f(x_1, x_2, \dots, x_D)
\label{eq:highest-order}
\end{align}
%
An example would be a \gp{} with an \seard{} kernel.
Such models are much more flexible than those having the form \eqref{eq:general-additive}, but this flexibility can make it difficult to generalize to new combinations of input variables.
In between these extremes are function classes depending on pairs or triplets of inputs, such as
%
\begin{align}
f(x_1, x_2, x_3) = f_{12}(x_1, x_2) + f_{23}(x_2, x_3) + f_{13}(x_1, x_3) .
\label{eq:second-order-additive}
\end{align}
%
We call the number of input variables appearing in each term the \emph{order} of that term.
Models containing terms only of intermediate order such as \eqref{eq:second-order-additive} allow more flexibility than models of form \eqref{eq:highest-order} (first-order), but have more structure than those of form \eqref{eq:general-additive} ($D$-th order).
Capturing the low-order additive structure present in a function can be expected to improve predictive accuracy.
However, if the function being learned depends in some way on an interaction between all input variables, a $D$th-order term is required in order for the model to be consistent.
%However, if the function also contains lower-order interactions, capturing that structure will still improve the predioctive performance on finite datasets.
%To the extent that a set of variables is relevant to the output, this model must be highly uncertain about any novel combination of those variables.
\section{Defining additive kernels}
To define the additive kernels introduced in this chapter, we first assign each dimension $i \in \{1 \dots D\}$ a one-dimensional \emph{base kernel} $k_i(x_i, x'_i)$.
Then the first order, second order and $n$th order additive kernels are defined as:
%
\begin{align}
k_{add_1}({\bf x, x'}) & = \sigma_1^2 \sum_{i=1}^D k_i(x_i, x_i') \\
k_{add_2}({\bf x, x'}) & = \sigma_2^2 \sum_{i=1}^D \sum_{j = i + 1}^D k_i(x_i, x_i') k_j(x_j, x_j') \\
k_{add_n}({\bf x, x'}) & = \sigma_n^2 \sum_{1 \leq i_1 < i_2 < ... < i_n \leq D} \left[ \prod_{d=1}^n k_{i_d}(x_{i_d}, x_{i_d}') \right] \\
k_{add_D}({\bf x, x'}) & = \sigma_D^2 \sum_{1 \leq i_1 < i_2 < ... < i_D \leq D} \left[ \prod_{d=1}^D k_{i_d}(x_{i_d}, x_{i_d}') \right] = \sigma_D^2 \prod_{d=1}^D k_{d}(x_{d}, x_{d}')
\end{align}
%
where $D$ is the dimension of the input space, and $\sigma_n^2$ is the variance assigned to all $n$th order interactions.
The $n$th-order kernel is a sum of ${D \choose n}$ terms.
In particular, the $D$th-order additive kernel has ${D \choose D} = 1$ term, a product of each dimension's kernel.
In the case where each base kernel is a one-dimensional squared-exponential kernel, the $D$th-order term corresponds to the multivariate squared-exponential kernel, also known as \seard{}:
%
\begin{align}
%k_{add_D}({\bf x, x'}) =
%\sigma_D^2 \prod_{d=1}^D k_{d}(x_{d}, x_{d}') =
\prod_{d=1}^D \kSE(x_d, x_d') =
\prod_{d=1}^D \sigma_d^2\exp \left( -\frac{ ( x_{d} - x_{d}')^2}{2 \ell^2_d} \right) =
\sigma_D^2 \exp \left( -\sum_{d=1}^D \frac{ ( x_{d} - x_{d}')^2}{2 \ell^2_d} \right)
\end{align}
%
%also commonly known as the Gaussian kernel.
The full additive kernel is a sum of the additive kernels of all orders.
%\subsubsection{Parameterization}
The only design choice necessary to specify an additive kernel is the selection of a one-dimensional base kernel for each input dimension.
Parameters of the base kernels (such as length-scales $\ell_1, \ell_2, \dots, \ell_D$) can be learned as per usual by maximizing the marginal likelihood of the training data.
\subsection{Weighting different orders of interaction}
In addition to the parameters of each dimension's kernel, additive kernels are equipped with a set of $D$ parameters $\sigma_1^2 \dots \sigma_D^2$.
%, controlling how much variance is assigned to each order of interaction.
These \emph{order variance} parameters have a useful interpretation: the $d$th order variance parameter specifies how much of the target function's variance comes from interactions of the $d$th order.
%\input{\additivetablesdir/r_concrete_500_hypers_table.tex}
%Table \ref{tbl:concrete} gives an example of parameters learnt by both a SE-GP, and by those learnt by an additive GP whose base kernels are one-dimensional squared exponential kernels. In this case, the 1st, 2nd and 3rd-order interactions contribute most to the final function.
%
%
\Cref{tbl:all_orders} shows examples of the variance contributed by different orders of interaction, estimated on real datasets.
These datasets are described in \cref{sec:additive-datasets}.
% --- Automatically generated by hypers_to_latex.m ---
% Exported at 03-Jun-2011 00:22:23
\begin{table}[h]
\caption[Relative variance contributed by each order of the additive model]
{Percentage of variance contributed by each order of interaction of the additive model on different datasets.
The maximum order of interaction is set to the input dimension or 10, whichever is smaller.
}
\begin{center}
\begin{tabular}{r | r r r r r r r r r r}
\multicolumn{1}{c}{} & \multicolumn{10}{c}{Order of interaction} \\
Dataset & \nth{1} & \nth{2} & \nth{3} & \nth{4} & \nth{5} & \nth{6} & \nth{7} & \nth{8} & \nth{9} & \nth{10} \\ \hline
pima & $0.1 $ & $0.1 $ & $0.1 $ & $0.3 $ & $1.5 $ & ${\bf 96.4}$ & $1.4 $ & $0.0 $ & & \\
liver & $0.0 $ & $0.2 $ & ${\bf 99.7 } $ & $0.1 $ & $0.0 $ & $0.0 $ & & & & \\
heart & ${\bf 77.6} $ & $0.0 $ & $0.0 $ & $0.0 $ & $0.1 $ & $0.1 $ & $0.1 $ & $0.1 $ & $0.1 $ & $22.0 $ \\
concrete & ${\bf 70.6 } $ & $13.3 $ & $13.8 $ & $2.3 $ & $0.0 $ & $0.0 $ & $0.0 $ & $0.0 $ & & \\
pumadyn-8nh & $0.0 $ & $0.1 $ & $0.1 $ & $0.1 $ & $0.1 $ & $0.1 $ & $0.1 $ & ${\bf 99.5 } $ & & \\
servo & ${\bf 58.7 }$ & $27.4 $ & $0.0 $ & $13.9 $ & & & & & & \\
housing & $0.1 $ & $0.6 $ & ${\bf 80.6 }$ & $1.4 $ & $1.8 $ & $0.8 $ & $0.7 $ & $0.8 $ & $0.6 $ & $12.7 $ \\
\end{tabular}
\end{center}
\label{tbl:all_orders}
\end{table}
% End automatically generated LaTeX
On different datasets, the dominant order of interaction estimated by the additive model varies widely.
In some cases, the variance is concentrated almost entirely onto a single order of interaction.
This may may be a side-effect of using the same lengthscales for all orders of interaction; lengthscales appropriate for low-dimensional regression might not be appropriate for high-dimensional regression.
%A re-scaling of lengthscales which enforces similar distances between datapoints might improve the model.
%An additive \gp{} with all of its variance coming from the 1st order is equivalent to a sum of one-dimensional functions.
%An additive \gp{} with all its variance coming from the $D$th order is equivalent to a \gp{} with an \seard{} kernel.
%
% We can see that the length-scales learnt are not necessarily the same. This is because, in a normal ARD-style kernel, the length-scale of a dimension is conflated with the variance of the function along that dimension. The additive kernel separates these two parameters.
%
%This table lets us examine the relative contribution of each order of interaction to the total function variance. Again, the ARD kernel is equivalent to an additive kernel in which all the variance is constrained to arise only from the highest-order interactions.
%Because the variance parameters can specify which degrees of interaction are important, the additive \gp{} can capture many different types of structure.
%By optimizing these parameters, we are approximately performing weighted model selection over a set of $D$ models.
%If the function we are modeling is decomposable into a sum of low-dimensional functions, our model can discover this fact and exploit it.
%The marginal likelihood will usually favor using lower orders if possible, since the $|K|^{-\frac{1}{2}}$ term in the \gp{} marginal likelihood (\cref{eq:gp_marg_lik}) will usually be larger for less flexible model classes.
%Low-order structure sometimes allows extrapolation, as shown in \cref{sec:additivity-extrapolation}.
%If low-dimensional additive structure is not present, the kernel parameters can specify a suitably flexible model, with interactions between as many variables as necessary.
%The expected number of zero crossings is $O(k^d)$ for product kernels (sqexp), $O(kd)$ for first-order additive, and $O\left(k {d \choose r}\right)$ for $r$th order additive functions. [Conjecture]
%To push the analogy, using a product kernel is analogous to performing polynomial regression when we are forced to use a high-degree polynomial, even when the data do not support it. One could also view the Gaussian kernel as one which makes the pessimistic assumption that the function we are model can vary independently for each new combination of variables observed.
\subsection{Efficiently evaluating additive kernels}
An additive kernel over $D$ inputs with interactions up to order $n$ has $O(2^n)$ terms.
Na\"{i}vely summing these terms is intractable.
One can exactly evaluate the sum over all terms in $O(D^2)$, while also weighting each order of interaction separately.
To efficiently compute the additive kernel, we exploit the fact that the $n$th order additive kernel corresponds to the $n$th \textit{elementary symmetric polynomial}~\citep{macdonald1998symmetric}
of the base kernels, which we denote $e_n$.
For example, if $\vx$ has 4 input dimensions ($D = 4$), and if we use the shorthand notation $k_d = k_d(x_d, x_d')$, then
%
\begin{align}
k_{\textnormal{add}_0}({\bf x, x'}) & = e_0( k_1, k_2, k_3, k_4 ) = 1 \\
k_{\textnormal{add}_1}({\bf x, x'}) & = e_1( k_1, k_2, k_3, k_4 ) = k_1 + k_2 + k_3 + k_4 \\
k_{\textnormal{add}_2}({\bf x, x'}) & = e_2( k_1, k_2, k_3, k_4 ) = k_1 k_2 + k_1 k_3 + k_1k_4 + k_2 k_3 + k_2 k_4 + k_3 k_4 \\
k_{\textnormal{add}_3}({\bf x, x'}) & = e_3( k_1, k_2, k_3, k_4 ) = k_1 k_2 k_3 + k_1 k_2 k_4 + k_1 k_3 k_4 + k_2 k_3 k_4 \\
k_{\textnormal{add}_4}({\bf x, x'}) & = e_4( k_1, k_2, k_3, k_4 ) = k_1 k_2 k_3 k_4
\end{align}
%
The Newton-Girard formulas give an efficient recursive form for computing these polynomials.
%If we define $s_k$ to be the $k$th power sum: $s_k(k_1,k_2,\dots,k_D) = \sum_{i=1}^Dk_i^k$, then
%
\begin{align}
k_{\textnormal{add}_n}({\bf x, x'}) = e_n(k_1,k_2,\dots,k_D) = \frac{1}{n} \sum_{a=1}^n (-1)^{(a-1)} e_{n-a}(k_1, k_2, \dots,k_D) \sum_{i=1}^Dk_i^a
\end{align}
%
Each iteration has cost $\mathcal{O}(D)$, given the next-lowest polynomial.
%\begin{equation}
%$s_k(k_1,k_2,\dots,k_n) = k_1^k + k_2^k + \dots + k_D^k = \sum_{i=1}^Dk_i^k$.
%\end{equation}
%These formulas have time complexity $\mathcal{O}( D^2 )$, while computing a sum over an exponential number of terms.
\subsubsection{Evaluation of derivatives}
Conveniently, we can use the same trick to efficiently compute the necessary derivatives of the additive kernel with respect to the base kernels.
This can be done by removing the base kernel of interest, $k_j$, from each term of the polynomials:
%
\begin{align}
\frac{\partial k_{add_n}}{\partial k_j} =
\frac{\partial e_{n}(k_1,k_2,\dots, k_D)}{\partial k_j} =
e_{n-1}(k_1, k_2, \dots,k_{j-1},k_{j+1}, \dots k_D)
% & = \frac{1}{n-1} \sum_{k=1}^{n-1} (-1)^{(k-1)} e_{n-k-1}(k_1,\dots,k_{j-1},k_{j+1}, \dots k_D)s_k(k_1,\dots,k_{j-1},k_{j+1}, \dots k_D)
\label{eq:additive-derivatives}
\end{align}
%
\Cref{eq:additive-derivatives} gives all terms that $k_j$ is multiplied by in the original polynomial, which are exactly the terms required by the chain rule.
These derivatives allow gradient-based optimization of the base kernel parameters with respect to the marginal likelihood.
% The final derivative is a sum of multilinear terms, so if only one term depends on the hyperparameter under consideration, we can factorise it out and compute the sum with one degree less.
\subsubsection{Computational cost}
The computational cost of evaluating the Gram matrix $k(\vX, \vX)$ of a product kernel such as the \seard{} scales as $\mathcal{O}(N^2D)$, while the cost of evaluating the Gram matrix of the additive kernel scales as $\mathcal{O}(N^2DR)$, where $R$ is the maximum degree of interaction allowed (up to $D$).
In high dimensions this can be a significant cost, even relative to the $\mathcal{O}(N^3)$ cost of inverting the Gram matrix.
However, \cref{tbl:all_orders} shows that sometimes only the first few orders of interaction contribute much variance.
In those cases, one may be able to limit the maximum degree of interaction in order to save time, without losing much accuracy.
\section{Additive models allow non-local interactions}
%The SE-GP model relies on local smoothing to make predictions at novel locations.
Commonly-used kernels such as the $\kSE$, $\kRQ$ or Mat\'{e}rn kernels are \emph{local} kernels, depending only on the scaled Euclidean distance between two points, all having the form:
\begin{align}
k({\bf x, x'}) = g \!\left( \; \sum_{d=1}^D \left( \frac{ x_{d} - x_{d}' }{ \ell_d } \right)^2 \right)
\end{align}
%\begin{eqnarray}
%k_{se}({\bf x, x'}) & = & v_D^2 \exp \left( -\frac{r^2}{2} \right) \\
%k_{\nu}({\bf x, x'}) & = & \frac{2^{1-\nu}}{\Gamma(\nu)}\left(\sqrt{2\nu}r\right) K_\nu \left(\sqrt{2\nu}r\right)
%\end{eqnarray}
%Where
%\begin{equation}
%$r = \sqrt{\sum_{d=1}^D \left( x_{d} - x_{d}' \right)^2 / l_d^2 }$.
%\end{equation}
for some function $g(\cdot)$.
\citet{bengio2006curse} argued that models based on local kernels are particularly susceptible to the \emph{curse of dimensionality}~\citep{bellman1956dynamic}, and are generally unable to extrapolate away from the training data.
%They emphasize that the locality of the kernels means that these models cannot capture non-local structure.
%They argue that many functions that we care about have such structure.
Methods based solely on local kernels sometimes require training examples at exponentially-many combinations of inputs.
In contrast, additive kernels can allow extrapolation away from the training data.
For example, additive kernels of second order give high covariance between function values at input locations which are similar in any two dimensions.
\begin{figure}[h!]
\centering
\renewcommand{\tabcolsep}{0pt}
\begin{tabular}{cccc}
1st-order terms: & 2nd-order terms: & 3rd-order terms: & All interactions: \\
$k_1 + k_2 + k_3$ & $k_1k_2 + k_2k_3 + k_1k_3$ & $k_1k_2k_3$ & \\
& & \seard{} kernel & Additive kernel\\
\includegraphics[trim=1em 0em 1em 3em, clip, width=0.25\textwidth]{\additivefigsdir/3d-kernel/3d_add_kernel_1} &
\includegraphics[trim=1em 0em 1em 3em, clip, width=0.25\textwidth]{\additivefigsdir/3d-kernel/3d_add_kernel_2} &
\includegraphics[trim=1em 0em 1em 3em, clip, width=0.25\textwidth]{\additivefigsdir/3d-kernel/3d_add_kernel_3} &
\includegraphics[trim=1em 0em 1em 3em, clip, width=0.25\textwidth]{\additivefigsdir/3d-kernel/3d_add_kernel_321}\\
$\vx - \vx'$ & $\vx - \vx'$ & $\vx - \vx'$ & $\vx - \vx'$\\[0.5em]
\end{tabular}
\caption[Isocontours of additive kernels in 3 dimensions]
{Isocontours of additive kernels in $D = 3$ dimensions.
The $D$th-order kernel only considers nearby points relevant, while lower-order kernels allow the output to depend on distant points, as long as they share one or more input value.}
\label{fig:kernels3d}
\end{figure}
\Cref{fig:kernels3d} provides a geometric comparison between squared-exponential kernels and additive kernels in 3 dimensions.
\Cref{sec:additivity-multiple-dimensions} contains an example of how additive kernels extrapolate differently than local kernels.
%This is one possible reason why low-order interactions might improve predictive accuracy.
\section{Dropout in Gaussian processes}
\label{sec:dropout-gps}
\emph{Dropout} is a recently-introduced method for regularizing neural networks~\citep{hinton2012improving, srivastava2013improving}.
Training with dropout entails independently setting to zero (``dropping'') some proportion $p$ of features or inputs, in order to improve the robustness of the resulting network, by reducing co-dependence between neurons.
To maintain similar overall activation levels, the remaining weights are divided by $p$.
% at test time. Alternatively, feature activations are divided by $p$ during training.
Predictions are made by approximately averaging over all possible ways of dropping out neurons.
\citet{baldi2013understanding} and \citet{wang2013fast} analyzed dropout in terms of the effective prior induced by this procedure in several models, such as linear and logistic regression.
In this section, we perform a similar analysis for \gp{}s, examining the priors on functions that result from performing dropout in the one-hidden-layer neural network implicitly defined by a \gp{}.
Recall from \cref{sec:relating} that some \gp{}s can be derived as infinitely-wide one-hidden-layer neural networks, with fixed activation functions $\feat(\vx)$ and independent random weights $\vw$ having zero mean and finite variance $\sigma^2_{\vw}$:
%
\begin{align}
f(\vx)
%= \frac{1}{K}{\mathbf \vnetweights}\tra \hPhi(\vx)
= \frac{1}{K} \sum_{i=1}^K \netweights_i \hphi_i(\vx)
\implies f \distas{K \to \infty} \GPt{0}{\sigma^2_{\vw} \feat(\vx)\tra\feat(\vx')} .
\label{eq:one-layer-gp-two}
\end{align}
%
%Mercer's theorem implies that we can write any \gp{} prior equivalently in this way,
%where $k(\vx, \vx') = \sigma^2_{\vw} \feat(\vx)\tra\feat(\vx'$).
%Having expressed a \gp{} as a neural network, we can examine the prior obtained by performing dropout in this network.
\subsection{Dropout on infinitely-wide hidden layers has no effect}
First, we examine the prior obtained by dropping features from $\hPhi(\vx)$ by setting weights in $\vnetweights$ to zero independently with probability $p$.
For simplicity, we assume that $\expectargs{}{\vnetweights} = \vzero$.
If the weights $w_i$ initially have finite variance $\sigma^2_{\netweights}$ before dropout, then the weights after dropout (denoted by $r_i w_i$, where $r_i$ is a Bernoulli random variable) will have variance:
%
\begin{align}
r_i \simiid \textnormal{Ber}(p) \qquad
\varianceargs{}{r_i \netweights_i} = p \sigma_{\netweights}^2 \;.
\end{align}
%
Because \cref{eq:one-layer-gp-two} is a result of the central limit theorem, it does not depend on the exact form of the distribution on $\vnetweights$, but only on its mean and variance.
Thus the central limit theorem still applies.
Performing dropout on the features of an infinitely-wide \MLP{} does not change the resulting model at all, except to rescale the output variance.
%If we were to rescale the weights by a factor other than $p^{-1/2}$, we would only rescale the output variance of the model, leaving all other aspects identical.
%Is there a better way to drop out features that would lead to robustness?
Indeed, dividing all weights by $\sqrt p$ restores the initial variance:
%
\begin{align}
%\expectargs{}{ \frac{1}{p} r_i \alpha_i } = \frac{p}{p}\mu = \mu, \quad
\varianceargs{}{ \frac{1}{p^\frac{1}{2}} r_i \netweights_i} = \frac{p}{p} \sigma_{\netweights}^2 = \sigma_{\netweights}^2
\end{align}
%
in which case dropout on the hidden units has no effect at all.
Intuitively, this is because no individual feature can have more than an infinitesimal contribution to the network output.
This result does not hold in neural networks having a finite number of hidden features with Gaussian-distributed weights, another model class that also gives rise to \gp{}s.
\subsection{Dropout on inputs gives additive covariance}
One can also perform dropout on the $D$ inputs to the \gp{}.
For simplicity, consider a stationary product kernel ${k(\vx, \vx') = \prod_{d=1}^D k_d(x_d, x_d')}$ which has been normalized such that $k(\vx, \vx) = 1$, and a dropout probability of $p = \nicefrac{1}{2}$.
In this case, the generative model can be written as:
%
\begin{align}
\vr = [r_1, r_2, \dots, r_D], \quad \textnormal{each} \;\; r_i \simiid \textnormal{Ber} \left( \frac{1}{2} \right), \quad f(\vx) | \vr \sim \GP \left( 0, \prod_{d=1}^D k_d(x_d, x_d')^{r_d} \right)
\end{align}
%
This is a mixture of $2^D$ \gp{}s, each depending on a different subset of the inputs:
%
\begin{align}
p \left( f(\vx) \right) =
\sum_{\vr} p \left( f(\vx) | \vr \right) p( \vr) =
\frac{1}{2^D} \sum_{\vr \in \{0,1\}^D} \GP \left(f(\vx) \,\Big|\, 0, \prod_{d=1}^D k_d(x_d, x_d')^{r_d} \right)
\label{eq:dropout-mixture}
\end{align}
We present two results which might give intuition about this model.
First, if the kernel on each dimension has the form ${k_d(x_d, x_d') = g \left( \frac{x_d - x_d'}{\ell_d} \right)}$, as does the \kSE{} kernel, then any input dimension can be dropped out by setting its lengthscale $\ell_d$ to $\infty$.
In this case, performing dropout on the inputs of a \gp{} corresponds to putting independent spike-and-slab priors on the lengthscales, with each dimension's distribution independently having ``spikes'' at $\ell_d = \infty$ with probability mass of $\nicefrac{1}{2}$.
Another way to understand the resulting prior is to note that the dropout mixture (\cref{eq:dropout-mixture}) has the same covariance as an additive \gp{}, scaled by a factor of $2^{-D}$:
\begin{align}
%f(\vx) \sim \textnormal{\gp} \left(0, \frac{1}{2^{-2D}} \sum_{\vr \in \{0,1\}^D} \prod_{d=1}^D k_d(x_d, x_d')^{r_d} \right)
%f(\vx) \sim \textnormal{\gp} \left(0, \frac{1}{2^{D}} \sum_{\vr \in \{0,1\}^D} \prod_{d=1}^D k_d(x_d, x_d')^{r_d} \right)
\cov\left( \colvec{f(\vx)}{f(\vx')} \right) = \frac{1}{2^{D}} \sum_{\vr \in \{0,1\}^D} \prod_{d=1}^D k_d(x_d, x_d')^{r_d}
\label{eq:dropout-mixture-covariance}
\end{align}
%TODO: Make this equation more precise
%
For dropout rates $p \neq \nicefrac{1}{2}$, the $d$th order terms will be weighted by $p^{(D - d)}(1-p)^d$.
Therefore, performing dropout on the inputs of a \gp{} gives a distribution with the same first two moments as an additive \gp{}.
This suggests an interpretation of additive \gp{}s as an approximation to a mixture of models where each model only depends on a subset of the input variables.
\section{Related work}
Since additive models are a relatively natural and easy-to-analyze model class, the literature on similar model classes is extensive.
This section attempts to provide a broad overview.
\subsubsection{Previous examples of additive \sgp{}s}
The additive models considered in this chapter are axis-aligned, but transforming the input space allows one to recover non-axis aligned additivity.
This model was explored by \citet{gilboa2013scaling}, who developed a linearly-transformed first-order additive \gp{} model, called projection-pursuit \gp{} regression.
They showed that inference in this model was possible in $\mathcal{O}(N)$ time.
\citet{durrande2011additive} also examined properties of additive \gp{}s, and proposed a layer-wise optimization strategy for kernel hyperparameters in these models.
\citet{plate1999accuracy} constructed an additive \gp{} having only first-order and $D$th-order terms, motivated by the desire to trade off the interpretability of first-order models with the flexibility of full-order models.
However, \cref{tbl:all_orders} shows that sometimes the intermediate degrees of interaction contribute most of the variance.
\citet{kaufman2010bayesian} used a closely related procedure called Gaussian process \ANOVA{} to perform a Bayesian analysis of meteorological data using 2nd and 3rd-order interactions.
They introduced a weighting scheme to ensure that each order's total contribution sums to zero.
It is not clear if this weighting scheme permits the use of the Newton-Girard formulas to speed computation of the Gram matrix.
%[TODO: read and mention \citep{friedman1981projection} from Hannes]
\subsubsection{Hierarchical kernel learning}
A similar model class was recently explored by \citet{Bach_HKL} called hierarchical kernel learning (\HKL{}).
\HKL{} uses a regularized optimization framework to build a weighted sum of an exponential number of kernels that can be computed in polynomial time.
%
\begin{figure}
\centering
\begin{tabular}{c|c}
Hierarchical kernel learning & All-orders additive \gp{} \\
\includegraphics[height=0.35\textwidth]{\additivefigsdir/compare_models/hkl.pdf} &
\includegraphics[height=0.35\textwidth]{\additivefigsdir/compare_models/additive.pdf}
\\ \hline \\
\gp{} with product kernel & First-order additive \gp{} \\
\includegraphics[height=0.35\textwidth]{\additivefigsdir/compare_models/ard.pdf} &
\includegraphics[height=0.35\textwidth]{\additivefigsdir/compare_models/gam.pdf} \\
\end{tabular}
\caption[A comparison of different additive model classes]
{A comparison of different additive model classes of 4-dimensional functions.
Circles represent different interaction terms, ranging from first-order to fourth-order interactions.
Shaded boxes represent the relative weightings of different terms.
\emph{Top left: }\HKL{} can select a hull of interaction terms, but must use a pre-determined weighting over those terms.
\emph{Top right:} the additive \gp{} model can weight each order of interaction separately, but weights all terms equally within each order.
\emph{Bottom row:} \gp{}s with product kernels (such as the \seard{} kernel) and first-order additive \gp{} models are special cases of the all-orders additive \gp{}, with all variance assigned to a single order of interaction.}
\label{hulls-figure}
\end{figure}
%
This method chooses among a \emph{hull} of kernels, defined as a set of terms such that if $\prod_{j \in J} k_j(\vx, \vx')$ is included in the set, then so are all products of strict subsets of the same elements:
$\prod_{j \in J / i} k_j(\vx, \vx')$, for all $i \in J$.
\HKL{} does not estimate a separate weighting parameter for each order.
%\HKL{} computes the sum over all orders in $\mathcal{O}(D)$ time by the formula:
%
%\begin{align}
%k_a(\vx, \vx') = v^2 \prod_{d=1}^D \left(1 + \alpha k_d( x_d, x_d') \right)
%\label{eqn:uniform}
%\end{align}
%
%which forces the weight of all $n$th order terms to be weighted by $\alpha^n$.
\Cref{hulls-figure} contrasts the \HKL{} model class with the additive \gp{} model.
Neither model class encompasses the other.
The main difficulty with this approach is that its parameters are hard to set other than by cross-validation.
\subsubsection{Support vector machines}
\citet{vapnik1998statistical} introduced the support vector \ANOVA{} decomposition, which has the same form as our additive kernel.
They recommend approximating the sum over all interactions with only one set of interactions ``of appropriate order'', presumably because of the difficulty of setting the parameters of an \SVM{}.
This is an example of a model choice which can be automated in the \gp{} framework.
\citet{stitson1999support} performed experiments which favourably compared the predictive accuracy of the support vector \ANOVA{} decomposition against polynomial and spline kernels.
They too allowed only one order to be active, and set parameters by cross-validation.
%
% The order of the kernel, kernel parameters, the allowable regression error $\epsilon$, and the cost hyperparameter $c$ were all set by a lengthy cross-validation process.
%\subsection{Smoothing Splines ANOVA}
\subsubsection{Other related models}
A closely related procedure from \citet{wahba1990spline} is smoothing-splines \ANOVA{} (\SSANOVA{}).
An \SSANOVA{} model is a weighted sum of splines along each dimension, splines over all pairs of dimensions, all triplets, etc, with each individual interaction term having a separate weighting parameter.
Because the number of terms to consider grows exponentially in the order, only terms of first and second order are usually considered in practice.
%Learning in SS-ANOVA is usually done via penalized-maximum likelihood with a fixed sparsity hyperparameter.
%\citet{shawe2004kernel} define ANOVA kernels thusly: "The ANOVA kernel of degree d is like the all-subsets kernel except that it is restricted to subsets of the given cardinality $d$."
This more general model class, in which each interaction term is estimated separately, is known in the physical sciences as high dimensional model representation (\HDMR{}).
\citet{rabitz1999general} review some properties and applications of this model class.
The main benefits of the model setup and parameterization proposed in this chapter are the ability to include all $D$ orders of interaction with differing weights, and the ability to learn kernel parameters individually per input dimension, allowing automatic relevance determination to operate.
\section{Regression and classification experiments}
\label{sec:additive-experiments}
\subsubsection{Choosing the base kernel}
%A $D$-dimensional \kSE-\ARD{} kernel has $D$ lengthscale parameters and one output variance parameter.
%A first-order additive $\kSE$ model has $D$ lengthscale parameters and $D$ output variance parameters.
An additive \gp{} using a separate $\kSE$ kernel on each input dimension will have ${3 \kerntimes D}$ effective parameters.
Because each additional parameter increases the tendency to overfit, %we recommend allowing only one kernel parameter per input dimension.
%
%Since there always exists a parameterization of the additive function corresponding to the product kernel, we initialize the hyperparameter search for the additive kernel by first learning parameters for the equivalent product kernel. This ensures that we start our search in a reasonable area of hyperparameter space.
%
in our experiments we fixed each one-dimensional kernel's output variance to be 1, and only estimated the lengthscale of each kernel.
\subsubsection{Methods}
We compared six different methods.
In the results tables below, \gp{} Additive refers to a \gp{} using the additive kernel with squared-exp base kernels.
For speed, we limited the maximum order of interaction to 10.
\gp{}-1st denotes an additive \gp{} model with only first-order interactions: a sum of one-dimensional kernels.
\gp{} Squared-exp is a \gp{} using an \seard{} kernel.
\HKL{} was run using the all-subsets kernel, which corresponds to the same set of interaction terms considered by \gp{} Additive.
For all \gp{} models, we fit kernel parameters by the standard method of maximizing training-set marginal likelihood, using \LBFGS{}~\citep{nocedal1980updating} for 500 iterations, allowing five random restarts.
In addition to learning kernel parameters, we fit a constant mean function to the data.
In the classification experiments, approximate \gp{} inference was performed using expectation propagation~\citep{minka2001expectation}.
The regression experiments also compared against the structure search method from \cref{ch:grammar}, run up to depth 10, using only the \SE{} and \RQ{} base kernels.
\subsection{Datasets}
\label{sec:additive-datasets}
We compared the above methods on regression and classification datasets from the \UCI{} repository~\citep{UCI}.
Their size and dimension are given in \cref{table:regression-dataset-stats,table:classification-dataset-stats}:
\begin{table}[h]
\caption{Regression dataset statistics}
\label{tbl:Regression Dataset Statistics}
\begin{center}
\begin{tabular}{l | lllll}
Method & bach & concrete & pumadyn & servo & housing \\ \hline
Dimension & 8 & 8 & 8 & 4 & 13 \\
Number of datapoints & 200 & 500 & 512 & 167 & 506
\end{tabular}
\end{center}
\label{table:regression-dataset-stats}
\end{table}
%
\begin{table}[h]
\caption{Classification dataset statistics}
\label{tbl:Classification Dataset Statistics}
\begin{center}
\begin{tabular}{l | llllll}
Method & breast & pima & sonar & ionosphere & liver & heart\\ \hline
Dimension & 9 & 8 & 60 & 32 & 6 & 13 \\
Number of datapoints & 449 & 768 & 208 & 351 & 345 & 297
\end{tabular}
\end{center}
\label{table:classification-dataset-stats}
\end{table}
\subsubsection{Bach synthetic dataset}
In addition to standard UCI repository datasets, we generated a synthetic dataset using the same recipe as \citet{Bach_HKL}.
This dataset was presumably designed to demonstrate the advantages of \HKL{} over a \gp{} using an \seard{} kernel.
%From a covariance matrix drawn from a Wishart distribution with 1024 degrees of freedom, we select 8 variables.
It is generated by passing correlated Gaussian-distributed inputs $x_1, x_2, \dots, x_8$ through a quadratic function
%
\begin{align}
f(\vx) = \sum_{i=1}^4 \sum_{j=i+1}^4 x_i x_j + \epsilon \qquad \epsilon \sim \Nt{0}{\sigma_\epsilon}.
\end{align}
%
%which sums all 2-way products of the first 4 variables, and adds Gaussian noise $\epsilon$.
This dataset will presumably be well-modeled by an additive kernel which includes all two-way interactions over the first 4 variables, but does not depend on the extra 4 correlated nuisance inputs or the higher-order interactions.%, as well as the higher-order interactions.
%TODO: Move description to grammar experiments?
%If the dataset is large enough, \HKL{} can construct a hull around only the 16 cross-terms optimal for predicting the output.
%\gp{}-\SE{}, in contrast, can learn to ignore the noisy copy variables, but cannot learn to ignore the higher-order interactions between dimensions.
%However, a \gp{} with an additive kernel can learn both to ignore irrelevant variables, and to ignore missing orders of interaction.
%With enough data, the marginal likelihood will favour hyperparameters specifying such a model.
%In this example, the additive \gp{} is able to recover the relevant structure.
\subsection{Results}
\Cref{tbl:Regression Mean Squared Error,tbl:Regression Negative Log Likelihood,tbl:Classification Percent Error,tbl:Classification Negative Log Likelihood} show mean performance across 10 train-test splits.
Because \HKL{} does not specify a noise model, it was not included in the likelihood comparisons.
% --- Automatically generated by resultsToLatex2.m ---
% Exported at 19-Jan-2012 10:55:19
\begin{table}
\caption[Comparison of predictive error on regression problems]
{Regression mean squared error}
\label{tbl:Regression Mean Squared Error}
\begin{center}
\begin{tabular}{l | r r r r r}
Method & \rotatebox{0}{ bach } & \rotatebox{0}{ concrete } & \rotatebox{0}{ pumadyn-8nh } & \rotatebox{0}{ servo } & \rotatebox{0}{ housing } \\ \hline
Linear Regression & $1.031$ & $0.404$ & $0.641$ & $0.523$ & $0.289$ \\
\gp{}-1st & $1.259$ & $0.149$ & $0.598$ & $0.281$ & $0.161$ \\
\HKL{} & $\mathbf{0.199}$ & $0.147$ & $0.346$ & $0.199$ & $0.151$ \\
\gp{} Squared-exp & $\mathbf{0.045}$ & $0.157$ & $0.317$ & $0.126$ & $\mathbf{0.092}$ \\
\gp{} Additive & $\mathbf{0.045}$ & $\mathbf{0.089}$ & $\mathbf{0.316}$ & $\mathbf{0.110}$ & $0.102$ \\
Structure Search & $\mathbf{0.044}$ & $\mathbf{0.087}$ & $\mathbf{0.315}$ & $\mathbf{0.102}$ & $\mathbf{0.082}$
\end{tabular}
\end{center}
%\end{table}
% End automatically generated LaTeX
%
% --- Automatically generated by resultsToLatex2.m ---
% Exported at 19-Jan-2012 10:55:19
%\begin{table}[h!]
\caption[Comparison of predictive likelihood on regression problems]
{Regression negative log-likelihood}
\label{tbl:Regression Negative Log Likelihood}
\begin{center}
\begin{tabular}{l | r r r r r}
Method & \rotatebox{0}{ bach } & \rotatebox{0}{ concrete } & \rotatebox{0}{ pumadyn-8nh } & \rotatebox{0}{ servo } & \rotatebox{0}{ housing } \\ \hline
Linear Regression & $2.430$ & $1.403$ & $1.881$ & $1.678$ & $1.052$ \\
\gp{}-1st & $1.708$ & $0.467$ & $1.195$ & $0.800$ & $0.457$ \\
GP Squared-exp & $\mathbf{-0.131}$ & $0.398$ & $0.843$ & $0.429$ & $0.207$ \\
GP Additive & $\mathbf{-0.131}$ & $\mathbf{0.114}$ & $\mathbf{0.841}$ & $\mathbf{0.309}$ & $0.194$ \\
Structure Search & $\mathbf{-0.141}$ & $\mathbf{0.065}$ & $\mathbf{0.840}$ & $\mathbf{0.265}$ & $\mathbf{0.059}$
\end{tabular}
\end{center}
\end{table}
% End automatically generated LaTeX
%
% --- Automatically generated by resultsToLatex2.m ---
% Exported at 03-Jun-2011 00:23:25
\begin{table}
\caption[Comparison of predictive error on classification problems]
{Classification percent error}
\label{tbl:Classification Percent Error}
\begin{center}
\begin{tabular}{l | r r r r r r}
Method & \rotatebox{0}{ breast } & \rotatebox{0}{ pima } & \rotatebox{0}{ sonar } & \rotatebox{0}{ ionosphere } & \rotatebox{0}{ liver } & \rotatebox{0}{ heart } \\ \hline
Logistic Regression & $7.611$ & $24.392$ & $26.786$ & $16.810$ & $45.060$ & $\mathbf{16.082}$ \\
\gp{}-1st & $\mathbf{5.189}$ & $\mathbf{22.419}$ & $\mathbf{15.786}$ & $\mathbf{8.524}$ & $\mathbf{29.842}$ & $\mathbf{16.839}$ \\
\HKL{} & $\mathbf{5.377}$ & $24.261$ & $\mathbf{21.000}$ & $9.119$ & $\mathbf{27.270}$ & $\mathbf{18.975}$ \\
\gp{} Squared-exp & $\mathbf{4.734}$ & $\mathbf{23.722}$ & $\mathbf{16.357}$ & $\mathbf{6.833}$ & $\mathbf{31.237}$ & $\mathbf{20.642}$ \\
\gp{} Additive & $\mathbf{5.566}$ & $\mathbf{23.076}$ & $\mathbf{15.714}$ & $\mathbf{7.976}$ & $\mathbf{30.060}$ & $\mathbf{18.496}$ \\
\end{tabular}
\end{center}
%\end{table}
% End automatically generated LaTeX
%
% --- Automatically generated by resultsToLatex2.m ---
% Exported at 03-Jun-2011 00:23:28
%\begin{table}[h!]
\caption[Comparison of predictive likelihood on classification problems]
{Classification negative log-likelihood}
\label{tbl:Classification Negative Log Likelihood}
\begin{center}
\begin{tabular}{l | r r r r r r}
Method & \rotatebox{0}{ breast } & \rotatebox{0}{ pima } & \rotatebox{0}{ sonar } & \rotatebox{0}{ ionosphere } & \rotatebox{0}{ liver } & \rotatebox{0}{ heart } \\ \hline
Logistic Regression & $0.247$ & $0.560$ & $4.609$ & $0.878$ & $0.864$ & $0.575$ \\
\gp{}-1st & $\mathbf{0.163}$ & $\mathbf{0.461}$ & $\mathbf{0.377}$ & $\mathbf{0.312}$ & $\mathbf{0.569}$ & $\mathbf{0.393}$ \\
\gp{} Squared-exp & $\mathbf{0.146}$ & $0.478$ & $\mathbf{0.425}$ & $\mathbf{0.236}$ & $\mathbf{0.601}$ & $0.480$ \\
\gp{} Additive & $\mathbf{0.150}$ & $\mathbf{0.466}$ & $\mathbf{0.409}$ & $\mathbf{0.295}$ & $\mathbf{0.588}$ & $\mathbf{0.415}$ \\
\end{tabular}
\end{center}
\end{table}
% End automatically generated LaTeX
On each dataset, the best performance is in boldface, along with all other performances not significantly different under a paired $t$-test.
%The additive model never performs significantly worse than any other model, and sometimes performs significantly better than all other models.
The additive and structure search methods usually outperformed the other methods, especially on regression problems.
%The difference between all methods is larger in the case of regression experiments.
The structure search outperforms the additive \gp{} at the cost of a slower search over kernels.
Structure search was on the order of 10 times slower than the additive \gp{}, which was on the order of 10 times slower than \gp{}-\kSE{}.
The additive \gp{} performed best on datasets well-explained by low orders of interaction, and approximately as well as the \SE{}-\gp{} model on datasets which were well explained by high orders of interaction (see \cref{tbl:all_orders}).
Because the additive \gp{} is a superset of both the \gp{}-1st model and the \gp{}-\kSE{} model, instances where the additive \gp{} performs slightly worse are presumably due to over-fitting, or due to the hyperparameter optimization becoming stuck in a local maximum. % Absence of over-fitting may explain the relatively strong performance of GP-GAM on classification tasks.
Performance of all \gp{} models could be expected to benefit from approximately integrating over kernel parameters.
The performance of \HKL{} is consistent with the results in \citet{Bach_HKL}, performing competitively but slightly worse than \SE{}-\gp{}.
\subsubsection{Source code}
%Additive Gaussian processes are particularly appealing in practice because their use requires only the specification of the base kernel; all other aspects of \gp{} inference remain the same.
%Note that we are also free to choose a different covariance function along each dimension.
All of the experiments in this chapter were performed using the standard \GPML{} toolbox, available at \url{http://wwww.gaussianprocess.org/gpml/code}.
The additive kernel described in this chapter is included in \GPML{} as of version 3.2.
Code to perform all experiments in this chapter is available at \url{http://www.github.com/duvenaud/additive-gps}.
\section{Conclusions}
This chapter presented a tractable \gp{} model consisting of a sum of exponentially-many functions, each depending on a different subset of the inputs.
Our experiments indicate that, to varying degrees, such additive structure is useful for modeling real datasets.
When it is present, modeling this structure allows our model to perform better than standard \gp{} models.
In the case where no such structure exists, the higher-order interaction terms present in the kernel can recover arbitrarily flexible models.
The additive \gp{} also affords some degree of interpretability: the variance parameters on each order of interaction indicate which sorts of structure are present the data, although they do not indicate which particular interactions explain the dataset.
The model class considered in this chapter is a subset of that explored by the structure search presented in \cref{ch:grammar}.
Thus additive \gp{}s can be considered a quick-and-dirty structure search, being strictly more limited in the types of structure that it can discover, but much faster and simpler to implement.
Closely related model classes have previously been explored, most notably smoothing-splines \ANOVA{} and the support vector \ANOVA{} decomposition.
However, these models can be difficult to apply in practice because their kernel parameters, regularization penalties, and the relevant orders of interaction must be set by hand or by cross-validation.
This chapter illustrates that the \gp{} framework allows these model choices to be performed automatically.
\outbpdocument{
\bibliographystyle{plainnat}
\bibliography{references.bib}
}