Gaussian process

Last updated

In probability theory and statistics, a Gaussian process is a stochastic process (a collection of random variables indexed by time or space), such that every finite collection of those random variables has a multivariate normal distribution. The distribution of a Gaussian process is the joint distribution of all those (infinitely many) random variables, and as such, it is a distribution over functions with a continuous domain, e.g. time or space.

Contents

The concept of Gaussian processes is named after Carl Friedrich Gauss because it is based on the notion of the Gaussian distribution (normal distribution). Gaussian processes can be seen as an infinite-dimensional generalization of multivariate normal distributions.

Gaussian processes are useful in statistical modelling, benefiting from properties inherited from the normal distribution. For example, if a random process is modelled as a Gaussian process, the distributions of various derived quantities can be obtained explicitly. Such quantities include the average value of the process over a range of times and the error in estimating the average using sample values at a small set of times. While exact models often scale poorly as the amount of data increases, multiple approximation methods have been developed which often retain good accuracy while drastically reducing computation time.

Definition

A time continuous stochastic process is Gaussian if and only if for every finite set of indices in the index set

is a multivariate Gaussian random variable. [1] That is the same as saying every linear combination of has a univariate normal (or Gaussian) distribution.

Using characteristic functions of random variables with denoting the imaginary unit such that , the Gaussian property can be formulated as follows: is Gaussian if and only if, for every finite set of indices , there are real-valued , with such that the following equality holds for all ,

or . The numbers and can be shown to be the covariances and means of the variables in the process. [2]

Variance

The variance of a Gaussian process is finite at any time , formally [3] :p. 515

Stationarity

For general stochastic processes strict-sense stationarity implies wide-sense stationarity but not every wide-sense stationary stochastic process is strict-sense stationary. However, for a Gaussian stochastic process the two concepts are equivalent. [3] :p. 518

A Gaussian stochastic process is strict-sense stationary if and only if it is wide-sense stationary.

Example

There is an explicit representation for stationary Gaussian processes. [4] A simple example of this representation is

where and are independent random variables with the standard normal distribution.

Covariance functions

A key fact of Gaussian processes is that they can be completely defined by their second-order statistics. [5] Thus, if a Gaussian process is assumed to have mean zero, defining the covariance function completely defines the process' behaviour. Importantly the non-negative definiteness of this function enables its spectral decomposition using the Karhunen–Loève expansion. Basic aspects that can be defined through the covariance function are the process' stationarity, isotropy, smoothness and periodicity. [6] [7]

Stationarity refers to the process' behaviour regarding the separation of any two points and . If the process is stationary, the covariance function depends only on . For example, the Ornstein–Uhlenbeck process is stationary.

If the process depends only on , the Euclidean distance (not the direction) between and , then the process is considered isotropic. A process that is concurrently stationary and isotropic is considered to be homogeneous; [8] in practice these properties reflect the differences (or rather the lack of them) in the behaviour of the process given the location of the observer.

Ultimately Gaussian processes translate as taking priors on functions and the smoothness of these priors can be induced by the covariance function. [6] If we expect that for "near-by" input points and their corresponding output points and to be "near-by" also, then the assumption of continuity is present. If we wish to allow for significant displacement then we might choose a rougher covariance function. Extreme examples of the behaviour is the OrnsteinUhlenbeck covariance function and the squared exponential where the former is never differentiable and the latter infinitely differentiable.

Periodicity refers to inducing periodic patterns within the behaviour of the process. Formally, this is achieved by mapping the input to a two dimensional vector .

Usual covariance functions

The effect of choosing different kernels on the prior function distribution of the Gaussian process. Left is a squared exponential kernel. Middle is Brownian. Right is quadratic. Gaussian process draws from prior distribution.png
The effect of choosing different kernels on the prior function distribution of the Gaussian process. Left is a squared exponential kernel. Middle is Brownian. Right is quadratic.

There are a number of common covariance functions: [7]

Here . The parameter is the characteristic length-scale of the process (practically, "how close" two points and have to be to influence each other significantly), is the Kronecker delta and the standard deviation of the noise fluctuations. Moreover, is the modified Bessel function of order and is the gamma function evaluated at . Importantly, a complicated covariance function can be defined as a linear combination of other simpler covariance functions in order to incorporate different insights about the data-set at hand.

The inferential results are dependent on the values of the hyperparameters (e.g. and ) defining the model's behaviour. A popular choice for is to provide maximum a posteriori (MAP) estimates of it with some chosen prior. If the prior is very near uniform, this is the same as maximizing the marginal likelihood of the process; the marginalization being done over the observed process values . [7] This approach is also known as maximum likelihood II, evidence maximization, or empirical Bayes . [9]

Continuity

For a Gaussian process, continuity in probability is equivalent to mean-square continuity, [10] :145 and continuity with probability one is equivalent to sample continuity. [11] :91 "Gaussian processes are discontinuous at fixed points." The latter implies, but is not implied by, continuity in probability. Continuity in probability holds if and only if the mean and autocovariance are continuous functions. In contrast, sample continuity was challenging even for stationary Gaussian processes (as probably noted first by Andrey Kolmogorov), and more challenging for more general processes. [12] :Sect. 2.8 [13] :69,81 [14] :80 [15] As usual, by a sample continuous process one means a process that admits a sample continuous modification. [16] :292 [17] :424

Stationary case

For a stationary Gaussian process some conditions on its spectrum are sufficient for sample continuity, but fail to be necessary. A necessary and sufficient condition, sometimes called Dudley–Fernique theorem, involves the function defined by

(the right-hand side does not depend on due to stationarity). Continuity of in probability is equivalent to continuity of at When convergence of to (as ) is too slow, sample continuity of may fail. Convergence of the following integrals matters:

these two integrals being equal according to integration by substitution The first integrand need not be bounded as thus the integral may converge () or diverge (). Taking for example for large that is, for small one obtains when and when In these two cases the function is increasing on but generally it is not. Moreover, the condition

(*)  there exists such that is monotone on

does not follow from continuity of and the evident relations (for all ) and

Theorem 1  Let be continuous and satisfy (*). Then the condition is necessary and sufficient for sample continuity of

Some history. [17] :424 Sufficiency was announced by Xavier Fernique in 1964, but the first proof was published by Richard M. Dudley in 1967. [16] :Theorem 7.1 Necessity was proved by Michael B. Marcus and Lawrence Shepp in 1970. [18] :380

There exist sample continuous processes such that they violate condition (*). An example found by Marcus and Shepp [18] :387 is a random lacunary Fourier series

where are independent random variables with standard normal distribution; frequencies are a fast growing sequence; and coefficients satisfy The latter relation implies

whence almost surely, which ensures uniform convergence of the Fourier series almost surely, and sample continuity of

Autocorrelation of a random lacunary Fourier series Autocorrelation of a random lacunary Fourier series.svg
Autocorrelation of a random lacunary Fourier series

Its autocovariation function

is nowhere monotone (see the picture), as well as the corresponding function

Brownian motion as the integral of Gaussian processes

A Wiener process (also known as Brownian motion) is the integral of a white noise generalized Gaussian process. It is not stationary, but it has stationary increments.

The Ornstein–Uhlenbeck process is a stationary Gaussian process.

The Brownian bridge is (like the Ornstein–Uhlenbeck process) an example of a Gaussian process whose increments are not independent.

The fractional Brownian motion is a Gaussian process whose covariance function is a generalisation of that of the Wiener process.

Driscoll's zero-one law

Driscoll's zero-one law is a result characterizing the sample functions generated by a Gaussian process.

Let be a mean-zero Gaussian process with non-negative definite covariance function . Let be a Reproducing kernel Hilbert space with positive definite kernel .

Then

where and are the covariance matrices of all possible pairs of points, implies

Moreover,

implies [19]

This has significant implications when , as

As such, almost all sample paths of a mean-zero Gaussian process with positive definite kernel will lie outside of the Hilbert space .

Linearly constrained Gaussian processes

For many applications of interest some pre-existing knowledge about the system at hand is already given. Consider e.g. the case where the output of the Gaussian process corresponds to a magnetic field; here, the real magnetic field is bound by Maxwell's equations and a way to incorporate this constraint into the Gaussian process formalism would be desirable as this would likely improve the accuracy of the algorithm.

A method on how to incorporate linear constraints into Gaussian processes already exists: [20]

Consider the (vector valued) output function which is known to obey the linear constraint (i.e. is a linear operator)

Then the constraint can be fulfilled by choosing , where is modelled as a Gaussian process, and finding such that

Given and using the fact that Gaussian processes are closed under linear transformations, the Gaussian process for obeying constraint becomes

Hence, linear constraints can be encoded into the mean and covariance function of a Gaussian process.

Applications

An example of Gaussian Process Regression (prediction) compared with other regression models. Regressions sine demo.svg
An example of Gaussian Process Regression (prediction) compared with other regression models.

A Gaussian process can be used as a prior probability distribution over functions in Bayesian inference. [7] [22] Given any set of N points in the desired domain of your functions, take a multivariate Gaussian whose covariance matrix parameter is the Gram matrix of your N points with some desired kernel, and sample from that Gaussian. For solution of the multi-output prediction problem, Gaussian process regression for vector-valued function was developed. In this method, a 'big' covariance is constructed, which describes the correlations between all the input and output variables taken in N points in the desired domain. [23] This approach was elaborated in detail for the matrix-valued Gaussian processes and generalised to processes with 'heavier tails' like Student-t processes. [24]

Inference of continuous values with a Gaussian process prior is known as Gaussian process regression, or kriging; extending Gaussian process regression to multiple target variables is known as cokriging. [25] Gaussian processes are thus useful as a powerful non-linear multivariate interpolation tool.

Gaussian processes are also commonly used to tackle numerical analysis problems such as numerical integration, solving differential equations, or optimisation in the field of probabilistic numerics.

Gaussian processes can also be used in the context of mixture of experts models, for example. [26] [27] The underlying rationale of such a learning framework consists in the assumption that a given mapping cannot be well captured by a single Gaussian process model. Instead, the observation space is divided into subsets, each of which is characterized by a different mapping function; each of these is learned via a different Gaussian process component in the postulated mixture.

In the natural sciences, Gaussian processes have found use as probabilistic models of astronomical time series and as predictors of molecular properties. [28]

Gaussian process prediction, or Kriging

Gaussian Process Regression (prediction) with a squared exponential kernel. Left plot are draws from the prior function distribution. Middle are draws from the posterior. Right is mean prediction with one standard deviation shaded. Gaussian Process Regression.png
Gaussian Process Regression (prediction) with a squared exponential kernel. Left plot are draws from the prior function distribution. Middle are draws from the posterior. Right is mean prediction with one standard deviation shaded.

When concerned with a general Gaussian process regression problem (Kriging), it is assumed that for a Gaussian process observed at coordinates , the vector of values is just one sample from a multivariate Gaussian distribution of dimension equal to number of observed coordinates . Therefore, under the assumption of a zero-mean distribution, , where is the covariance matrix between all possible pairs for a given set of hyperparameters θ. [7] As such the log marginal likelihood is:

and maximizing this marginal likelihood towards θ provides the complete specification of the Gaussian process f. One can briefly note at this point that the first term corresponds to a penalty term for a model's failure to fit observed values and the second term to a penalty term that increases proportionally to a model's complexity. Having specified θ, making predictions about unobserved values at coordinates x* is then only a matter of drawing samples from the predictive distribution where the posterior mean estimate A is defined as

and the posterior variance estimate B is defined as:

where is the covariance between the new coordinate of estimation x* and all other observed coordinates x for a given hyperparameter vector θ, and are defined as before and is the variance at point x* as dictated by θ. It is important to note that practically the posterior mean estimate of (the "point estimate") is just a linear combination of the observations ; in a similar manner the variance of is actually independent of the observations . A known bottleneck in Gaussian process prediction is that the computational complexity of inference and likelihood evaluation is cubic in the number of points |x|, and as such can become unfeasible for larger data sets. [6] Works on sparse Gaussian processes, that usually are based on the idea of building a representative set for the given process f, try to circumvent this issue. [29] [30] The kriging method can be used in the latent level of a nonlinear mixed-effects model for a spatial functional prediction: this technique is called the latent kriging. [31]

Often, the covariance has the form , where is a scaling parameter. Examples are the Matérn class covariance functions. If this scaling parameter is either known or unknown (i.e. must be marginalized), then the posterior probability, , i.e. the probability for the hyperparameters given a set of data pairs of observations of and , admits an analytical expression. [32]

Bayesian neural networks as Gaussian processes

Bayesian neural networks are a particular type of Bayesian network that results from treating deep learning and artificial neural network models probabilistically, and assigning a prior distribution to their parameters. Computation in artificial neural networks is usually organized into sequential layers of artificial neurons. The number of neurons in a layer is called the layer width. As layer width grows large, many Bayesian neural networks reduce to a Gaussian process with a closed form compositional kernel. This Gaussian process is called the Neural Network Gaussian Process (NNGP). [7] [33] [34] It allows predictions from Bayesian neural networks to be more efficiently evaluated, and provides an analytic tool to understand deep learning models.

Computational issues

In practical applications, Gaussian process models are often evaluated on a grid leading to multivariate normal distributions. Using these models for prediction or parameter estimation using maximum likelihood requires evaluating a multivariate Gaussian density, which involves calculating the determinant and the inverse of the covariance matrix. Both of these operations have cubic computational complexity which means that even for grids of modest sizes, both operations can have a prohibitive computational cost. This drawback led to the development of multiple approximation methods.

See also

Related Research Articles

<span class="mw-page-title-main">Multivariate normal distribution</span> Generalization of the one-dimensional normal distribution to higher dimensions

In probability theory and statistics, the multivariate normal distribution, multivariate Gaussian distribution, or joint normal distribution is a generalization of the one-dimensional (univariate) normal distribution to higher dimensions. One definition is that a random vector is said to be k-variate normally distributed if every linear combination of its k components has a univariate normal distribution. Its importance derives mainly from the multivariate central limit theorem. The multivariate normal distribution is often used to describe, at least approximately, any set of (possibly) correlated real-valued random variables each of which clusters around a mean value.

In statistics, maximum likelihood estimation (MLE) is a method of estimating the parameters of an assumed probability distribution, given some observed data. This is achieved by maximizing a likelihood function so that, under the assumed statistical model, the observed data is most probable. The point in the parameter space that maximizes the likelihood function is called the maximum likelihood estimate. The logic of maximum likelihood is both intuitive and flexible, and as such the method has become a dominant means of statistical inference.

<span class="mw-page-title-main">Wiener process</span> Stochastic process generalizing Brownian motion

In mathematics, the Wiener process is a real-valued continuous-time stochastic process named in honor of American mathematician Norbert Wiener for his investigations on the mathematical properties of the one-dimensional Brownian motion. It is often also called Brownian motion due to its historical connection with the physical process of the same name originally observed by Scottish botanist Robert Brown. It is one of the best known Lévy processes and occurs frequently in pure and applied mathematics, economics, quantitative finance, evolutionary biology, and physics.

In mathematics, a Gaussian function, often simply referred to as a Gaussian, is a function of the base form

<span class="mw-page-title-main">Radon transform</span> Integral transform

In mathematics, the Radon transform is the integral transform which takes a function f defined on the plane to a function Rf defined on the (two-dimensional) space of lines in the plane, whose value at a particular line is equal to the line integral of the function over that line. The transform was introduced in 1917 by Johann Radon, who also provided a formula for the inverse transform. Radon further included formulas for the transform in three dimensions, in which the integral is taken over planes. It was later generalized to higher-dimensional Euclidean spaces and more broadly in the context of integral geometry. The complex analogue of the Radon transform is known as the Penrose transform. The Radon transform is widely applicable to tomography, the creation of an image from the projection data associated with cross-sectional scans of an object.

In mathematical statistics, the Fisher information is a way of measuring the amount of information that an observable random variable X carries about an unknown parameter θ of a distribution that models X. Formally, it is the variance of the score, or the expected value of the observed information.

<span class="mw-page-title-main">Cross-correlation</span> Covariance and correlation

In signal processing, cross-correlation is a measure of similarity of two series as a function of the displacement of one relative to the other. This is also known as a sliding dot product or sliding inner-product. It is commonly used for searching a long signal for a shorter, known feature. It has applications in pattern recognition, single particle analysis, electron tomography, averaging, cryptanalysis, and neurophysiology. The cross-correlation is similar in nature to the convolution of two functions. In an autocorrelation, which is the cross-correlation of a signal with itself, there will always be a peak at a lag of zero, and its size will be the signal energy.

In Bayesian statistics, a maximum a posteriori probability (MAP) estimate is an estimate of an unknown quantity, that equals the mode of the posterior distribution. The MAP can be used to obtain a point estimate of an unobserved quantity on the basis of empirical data. It is closely related to the method of maximum likelihood (ML) estimation, but employs an augmented optimization objective which incorporates a prior distribution over the quantity one wants to estimate. MAP estimation can therefore be seen as a regularization of maximum likelihood estimation.

<span class="mw-page-title-main">Rice distribution</span> Probability distribution

In probability theory, the Rice distribution or Rician distribution is the probability distribution of the magnitude of a circularly-symmetric bivariate normal random variable, possibly with non-zero mean (noncentral). It was named after Stephen O. Rice (1907–1986).

In Bayesian probability, the Jeffreys prior, named after Sir Harold Jeffreys, is a non-informative prior distribution for a parameter space; its density function is proportional to the square root of the determinant of the Fisher information matrix:

<span class="mw-page-title-main">Ornstein–Uhlenbeck process</span> Stochastic process modeling random walk with friction

In mathematics, the Ornstein–Uhlenbeck process is a stochastic process with applications in financial mathematics and the physical sciences. Its original application in physics was as a model for the velocity of a massive Brownian particle under the influence of friction. It is named after Leonard Ornstein and George Eugene Uhlenbeck.

The cross-entropy (CE) method is a Monte Carlo method for importance sampling and optimization. It is applicable to both combinatorial and continuous problems, with either a static or noisy objective.

Covariance matrix adaptation evolution strategy (CMA-ES) is a particular kind of strategy for numerical optimization. Evolution strategies (ES) are stochastic, derivative-free methods for numerical optimization of non-linear or non-convex continuous optimization problems. They belong to the class of evolutionary algorithms and evolutionary computation. An evolutionary algorithm is broadly based on the principle of biological evolution, namely the repeated interplay of variation and selection: in each generation (iteration) new individuals are generated by variation, usually in a stochastic way, of the current parental individuals. Then, some individuals are selected to become the parents in the next generation based on their fitness or objective function value . Like this, over the generation sequence, individuals with better and better -values are generated.

In statistics, the inverse Wishart distribution, also called the inverted Wishart distribution, is a probability distribution defined on real-valued positive-definite matrices. In Bayesian statistics it is used as the conjugate prior for the covariance matrix of a multivariate normal distribution.

In operator theory, a branch of mathematics, a positive-definite kernel is a generalization of a positive-definite function or a positive-definite matrix. It was first introduced by James Mercer in the early 20th century, in the context of solving integral operator equations. Since then, positive-definite functions and their various analogues and generalizations have arisen in diverse parts of mathematics. They occur naturally in Fourier analysis, probability theory, operator theory, complex function-theory, moment problems, integral equations, boundary-value problems for partial differential equations, machine learning, embedding problem, information theory, and other areas.

A product distribution is a probability distribution constructed as the distribution of the product of random variables having two other known distributions. Given two statistically independent random variables X and Y, the distribution of the random variable Z that is formed as the product is a product distribution.

Within bayesian statistics for machine learning, kernel methods arise from the assumption of an inner product space or similarity structure on inputs. For some such methods, such as support vector machines (SVMs), the original formulation and its regularization were not Bayesian in nature. It is helpful to understand them from a Bayesian perspective. Because the kernels are not necessarily positive semidefinite, the underlying structure may not be inner product spaces, but instead more general reproducing kernel Hilbert spaces. In Bayesian probability kernel methods are a key component of Gaussian processes, where the kernel function is known as the covariance function. Kernel methods have traditionally been used in supervised learning problems where the input space is usually a space of vectors while the output space is a space of scalars. More recently these methods have been extended to problems that deal with multiple outputs such as in multi-task learning.

In the study of artificial neural networks (ANNs), the neural tangent kernel (NTK) is a kernel that describes the evolution of deep artificial neural networks during their training by gradient descent. It allows ANNs to be studied using theoretical tools from kernel methods.

A Neural Network Gaussian Process (NNGP) is a Gaussian process (GP) obtained as the limit of a certain type of sequence of neural networks. Specifically, a wide variety of network architectures converges to a GP in the infinitely wide limit, in the sense of distribution. The concept constitutes an intensional definition, i.e., a NNGP is just a GP, but distinguished by how it is obtained.

In statistics and machine learning, Gaussian process approximation is a computational method that accelerates inference tasks in the context of a Gaussian process model, most commonly likelihood evaluation and prediction. Like approximations of other models, they can often be expressed as additional assumptions imposed on the model, which do not correspond to any actual feature, but which retain its key properties while simplifying calculations. Many of these approximation methods can be expressed in purely linear algebraic or functional analytic terms as matrix or function approximations. Others are purely algorithmic and cannot easily be rephrased as a modification of a statistical model.

References

  1. MacKay, David, J.C. (2003). Information Theory, Inference, and Learning Algorithms (PDF). Cambridge University Press. p. 540. ISBN   9780521642989. The probability distribution of a function is a Gaussian processes if for any finite selection of points , the density is a Gaussian{{cite book}}: CS1 maint: multiple names: authors list (link)
  2. Dudley, R.M. (1989). Real Analysis and Probability. Wadsworth and Brooks/Cole. ISBN   0-534-10050-3.
  3. 1 2 Amos Lapidoth (8 February 2017). A Foundation in Digital Communication. Cambridge University Press. ISBN   978-1-107-17732-1.
  4. Kac, M.; Siegert, A.J.F (1947). "An Explicit Representation of a Stationary Gaussian Process". The Annals of Mathematical Statistics. 18 (3): 438–442. doi: 10.1214/aoms/1177730391 .
  5. Bishop, C.M. (2006). Pattern Recognition and Machine Learning. Springer. ISBN   978-0-387-31073-2.
  6. 1 2 3 Barber, David (2012). Bayesian Reasoning and Machine Learning. Cambridge University Press. ISBN   978-0-521-51814-7.
  7. 1 2 3 4 5 6 Rasmussen, C.E.; Williams, C.K.I (2006). Gaussian Processes for Machine Learning. MIT Press. ISBN   978-0-262-18253-9.
  8. Grimmett, Geoffrey; David Stirzaker (2001). Probability and Random Processes. Oxford University Press. ISBN   978-0198572220.
  9. Seeger, Matthias (2004). "Gaussian Processes for Machine Learning". International Journal of Neural Systems. 14 (2): 69–104. CiteSeerX   10.1.1.71.1079 . doi:10.1142/s0129065704001899. PMID   15112367. S2CID   52807317.
  10. Dudley, R. M. (1975). "The Gaussian process and how to approach it" (PDF). Proceedings of the International Congress of Mathematicians. Vol. 2. pp. 143–146.
  11. Dudley, R. M. (2010). "Sample Functions of the Gaussian Process". Selected Works of R.M. Dudley. Vol. 1. pp. 66–103. doi:10.1007/978-1-4419-5821-1_13. ISBN   978-1-4419-5820-4.{{cite book}}: |journal= ignored (help)
  12. Talagrand, Michel (2014). Upper and lower bounds for stochastic processes: modern methods and classical problems. Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge / A Series of Modern Surveys in Mathematics. Springer, Heidelberg. ISBN   978-3-642-54074-5.
  13. Ledoux, Michel (1996), "Isoperimetry and Gaussian analysis", in Dobrushin, Roland; Groeneboom, Piet; Ledoux, Michel (eds.), Lectures on Probability Theory and Statistics: Ecole d'Eté de Probabilités de Saint-Flour XXIV–1994, Lecture Notes in Mathematics, vol. 1648, Berlin: Springer, pp. 165–294, doi:10.1007/BFb0095676, ISBN   978-3-540-62055-6, MR   1600888
  14. Adler, Robert J. (1990). An Introduction to Continuity, Extrema, and Related Topics for General Gaussian Processes. Vol. 12. Hayward, California: Institute of Mathematical Statistics. ISBN   0-940600-17-X. JSTOR   4355563. MR   1088478.{{cite book}}: |journal= ignored (help)
  15. Berman, Simeon M. (1992). "Review of: Adler 1990 'An introduction to continuity...'". Mathematical Reviews. MR   1088478.
  16. 1 2 Dudley, R. M. (1967). "The sizes of compact subsets of Hilbert space and continuity of Gaussian processes". Journal of Functional Analysis. 1 (3): 290–330. doi: 10.1016/0022-1236(67)90017-1 .
  17. 1 2 Marcus, M.B.; Shepp, Lawrence A. (1972). "Sample behavior of Gaussian processes". Proceedings of the sixth Berkeley symposium on mathematical statistics and probability, vol. II: probability theory. Vol. 6. Univ. California, Berkeley. pp. 423–441.
  18. 1 2 Marcus, Michael B.; Shepp, Lawrence A. (1970). "Continuity of Gaussian processes". Transactions of the American Mathematical Society . 151 (2): 377–391. doi: 10.1090/s0002-9947-1970-0264749-1 . JSTOR   1995502.
  19. Driscoll, Michael F. (1973). "The reproducing kernel Hilbert space structure of the sample paths of a Gaussian process". Zeitschrift für Wahrscheinlichkeitstheorie und Verwandte Gebiete. 26 (4): 309–316. doi: 10.1007/BF00534894 . ISSN   0044-3719. S2CID   123348980.
  20. Jidling, Carl; Wahlström, Niklas; Wills, Adrian; Schön, Thomas B. (2017-09-19). "Linearly constrained Gaussian processes". arXiv: 1703.00787 [stat.ML].
  21. The documentation for scikit-learn also has similar examples.
  22. Liu, W.; Principe, J.C.; Haykin, S. (2010). Kernel Adaptive Filtering: A Comprehensive Introduction. John Wiley. ISBN   978-0-470-44753-6. Archived from the original on 2016-03-04. Retrieved 2010-03-26.
  23. Álvarez, Mauricio A.; Rosasco, Lorenzo; Lawrence, Neil D. (2012). "Kernels for vector-valued functions: A review" (PDF). Foundations and Trends in Machine Learning. 4 (3): 195–266. doi:10.1561/2200000036. S2CID   456491.
  24. Chen, Zexun; Wang, Bo; Gorban, Alexander N. (2019). "Multivariate Gaussian and Student-t process regression for multi-output prediction". Neural Computing and Applications. 32 (8): 3005–3028. arXiv: 1703.04455 . doi: 10.1007/s00521-019-04687-8 .
  25. Stein, M.L. (1999). Interpolation of Spatial Data: Some Theory for Kriging. Springer.
  26. Platanios, Emmanouil A.; Chatzis, Sotirios P. (2014). "Gaussian Process-Mixture Conditional Heteroscedasticity". IEEE Transactions on Pattern Analysis and Machine Intelligence. 36 (5): 888–900. doi:10.1109/TPAMI.2013.183. PMID   26353224. S2CID   10424638.
  27. Chatzis, Sotirios P. (2013). "A latent variable Gaussian process model with Pitman–Yor process priors for multiclass classification". Neurocomputing. 120: 482–489. doi:10.1016/j.neucom.2013.04.029.
  28. Griffiths, Ryan-Rhys (2022). Applications of Gaussian Processes at Extreme Lengthscales: From Molecules to Black Holes (PhD thesis). University of Cambridge. arXiv: 2303.14291 . doi:10.17863/CAM.93643.
  29. Smola, A.J.; Schoellkopf, B. (2000). "Sparse greedy matrix approximation for machine learning". Proceedings of the Seventeenth International Conference on Machine Learning: 911–918. CiteSeerX   10.1.1.43.3153 .
  30. Csato, L.; Opper, M. (2002). "Sparse on-line Gaussian processes". Neural Computation. 14 (3): 641–668. CiteSeerX   10.1.1.335.9713 . doi:10.1162/089976602317250933. PMID   11860686. S2CID   11375333.
  31. Lee, Se Yoon; Mallick, Bani (2021). "Bayesian Hierarchical Modeling: Application Towards Production Results in the Eagle Ford Shale of South Texas". Sankhya B. 84: 1–43. doi: 10.1007/s13571-020-00245-8 .
  32. Ranftl, Sascha; Melito, Gian Marco; Badeli, Vahid; Reinbacher-Köstinger, Alice; Ellermann, Katrin; von der Linden, Wolfgang (2019-12-31). "Bayesian Uncertainty Quantification with Multi-Fidelity Data and Gaussian Processes for Impedance Cardiography of Aortic Dissection". Entropy. 22 (1): 58. Bibcode:2019Entrp..22...58R. doi: 10.3390/e22010058 . ISSN   1099-4300. PMC   7516489 . PMID   33285833.
  33. Novak, Roman; Xiao, Lechao; Hron, Jiri; Lee, Jaehoon; Alemi, Alexander A.; Sohl-Dickstein, Jascha; Schoenholz, Samuel S. (2020). "Neural Tangents: Fast and Easy Infinite Neural Networks in Python". International Conference on Learning Representations. arXiv: 1912.02803 .
  34. Neal, Radford M. (2012). Bayesian Learning for Neural Networks. Springer Science and Business Media.

Literature

Software

Video tutorials