Functional linear regression models

edit

Functional linear models can be viewed as an extension of the traditional multivariate linear models that associates vector responses with vector covariates. The traditional linear model with scalar response   and vector covariate   can be expressed as

 
(2)

where   denotes the inner product in Euclidean space,   and   denote the regression coefficients, and   is a zero mean finite variance random error (noise). Functional linear models can be divided into two types based on the responses.

Functional regression models with scalar response

edit

Replacing the vector covariate   and the coefficient vector   in model (2) by a centered functional covariate   and coefficient function   for   and replacing the inner product in Euclidean space by that in Hilbert space  , one arrives at the functional linear model

 
(3)

One ad hoc approach to estimating   and   is to expand the covariate   and the coefficient function   in the same functional basis, such as the B-spline basis or the eigenbasis in (FPCA formula). Specially, consider an orthonormal basis,  , of the function space. Then expanding   and   in this basis leads to  ,   and model (3) is seen to be equivalent to the traditional linear model (2) of the form

 
(4)

where in implementations the sum on r.h.s. is replaced by a finite sum that is truncated at the first K terms. In addition to the basis-expansion approach, a penalized approach using either P-splines or smoothing splines has also been studied[1] can be applied. In special case where the basis functions   are selected as the eigenfunctions   of   (refers to (1)), the basis representation approach in (4) is equivalent to conducting a principal component regression, albeit with an increasing number of principal components.

The simple functional linear model (3) can be extended to multiple functional covariates,  , also including additional vector covariates  , where  , by

 
(5)

where   is regression coefficient for  ,   is the interval where   is defined,   is the centered functional covariate given by  , and   is regression coefficient function for  , for  . Model (3) and (5) have been studied extensively.[2][3][4][5]

Functional regression models with functional response

edit

For a functional response   on   and multiple functional covariates  ,  , two major models have been considered.[6][7] One of these two models is generally referred to as functional linear model (FLM) which can be written as:

 
(6)

where   is the functional intercept, for   ,   is a centered functional covariate on  ,   is the corresponding functional slopes with same domain, respectively, and   is usually a random process with mean zero and finite variance.[6] In this case, at any given time  , the value of  , i.e.,  , depends on the entire trajectories of  . In particular, taking   as a constant function yields a special case of model (6)

 
which is a functional linear model with functional responses and scalar covariates. Model (6) are also studied extensively.[8][9][10][11][12]

Concurrent regression model

edit

The second model assumes that   and is most often referred to as "varying-coefficient" model. It can be written as

 
(7)

where   are multiple functional covariates on  ,   are the coefficient functions defined on the same interval and   is usually assumed to be a random process with mean zero and finite variance.[6] This model assumes that the value of   depends on the current value of   only and not the history   or future value, hence it is a "concurrent regression model". Various estimation methods can be applied to model (7).[13][14][15][16][17][18]

Clustering of functional data

edit

For vector-valued multivariate data, hierarchical clustering and the k-means partitioning methods are two classical and popular approaches. Classical clustering concepts for vector-valued multivariate data can typically be extended to functional data, where various considerations arise, such as discrete approximations of distance measures, and dimension reduction of the infinite-dimensional functional data objects. Generally, k-means type clustering algorithms have been widely applied to functional data, and more popular than hierarchical clustering algorithms.

Mean functions as cluster centers

edit

 It is natural to view cluster mean functions as the clusters in functional clustering. Specifically, for a sample of functional data  , the k-means functional clustering aims to find a set of cluster centers  , assuming there are   clusters, by minimizing the sum of the squared distances between   and the cluster centers that are associated with their cluster labels  , for a suitable functional distance  . The distance   is often chosen as the   norm. The traditional k-means clustering for vector-valued multivariate data has been extended to functional data using mean functions as cluster centers, and one can distinguish two typical approaches, as follows.

Functional clustering via functional basis expansion

edit

Given a set of pre-specified basis functions   of the function space, the first   projections   of the observed trajectories onto the space spanned by the set of basis functions can be used to represent the functional data, where  ,  . The distributional patterns of   then reflect the clusters in function space. Then applying available clustering algorithms for multivariate data, such as the k-means algorithm, to partition the estimated sets of coefficients. Finally, one can obtain cluster centers   on the projected space, and thus the set of cluster centers in the function space  , where  . For good performance, this method requires a judicious choice of the basis function and researches corresponding to different choice of basis function are studied a lot, such as B-spline basis[19], Fourier basis[20], P splines basis[21], Gaussian orthonormalized basis[22] and wavelet basis[23] functions.

Functional clustering via FPCA

edit

In contrast to the functional basis expansion approach that requires a pre-specified set of basis functions, the finite approximation FPCA approach (1) using the FPCs employs data-adaptive basis functions that are determined by the covariance function of the functional data. Then the FPC scores   play a similar role as the basis coefficients   for clustering.[24]

When the mean functions as the cluster centers are sufficient to define the clusters, this step is sufficient. However, when covariance structures also play a role to distinguish clusters, taking mean functions as cluster is not adequate, as will be discussed as followed.

Subspaces as cluster centers

edit

Since functional data are realizations of random functions, it is natural to use differences in the stochastic structure of random functions for clustering. This idea is particularly sensible in functional data clustering, utilizing the truncated Karhunen–Loève representation in (1). Then the subspace spanned by the components of the expansion, the mean function and the set of the eigenfunctions, can be used to characterize clusters. Therefore, clusters of the data set are identified via subspace projection such that cluster centers hinge on the stochastic structure of the random functions, rather than the mean functions only.

The FPC subspace-projected k-centers functional clustering approach uses subspaces as cluster centers.[25] Let   be the cluster membership variable, and the FPC subspace  ,  , assuming that there are   clusters. The projected function of   onto the FPC subspace   can be written as

 
(8)

One aims to find the set of cluster centers  , such that the best cluster membership of  ,  , is determined by minimizing the discrepancy between the projected function   and the observation  ,

 
(9)

In contrast, k-means clustering aims to find the set of cluster sample means as the cluster centers, instead of the subspaces spanned by  . The initial step of the subspace-projected clustering procedure uses only  , which reduces to the k-means functional clustering. In the subsequent iteration steps, the mean function and the set of eigenfunction for each cluster is updated and used to identify the set of cluster subspaces  , until iterations converge. This functional clustering approach simultaneously identifies the structural components of the stochastic representation for each cluster.

Functional clustering with mixture models

edit

Model-based clustering[26] based on mixture models is widely used in clustering vector-valued multivariate data and has been extended to functional data clustering. In this approach, similarly to the k-means type of functional data clustering, typical mixture model-based approaches to functional data clustering in a first step project the infinite dimensional functional data onto low-dimensional subspaces. For example, we can apply functional clustering models based on Gaussian mixture[27][28][29] distributions to the natural cubic spline basis coefficients, with emphasis on clustering sparsely sampled functional data. Furthermore, random effects modeling also provides a model-based clustering approach, using mixed effects models with B-splines or P-splines.[21] For clustering longitudinal data, a linear mixed model for clustering using a penalized normal mixture as random effects distribution has been studied.[30] Bayesian hierarchical clustering also plays an important role in the development of model-based functional clustering.[31][32][33][34]


Clustering of functional data

edit

For vector-valued multivariate data, k-means partitioning methods and hierarchical clustering are two main approaches. Classical clustering concepts for vector-valued multivariate data can typically be extended to functional data, where various considerations arise, such as discrete approximations of distance measures, and dimension reduction of the infinite-dimensional functional data objects. Generally, k-means type clustering algorithms have been widely applied to functional data, and more popular than hierarchical clustering algorithms. For k-means clustering, mean functions are usually viewed as the cluster centers. Specifically, for a sample of functional data  , the k-means functional clustering aims to find a set of cluster centers  , assuming there are   clusters, by minimizing the sum of the squared distances between   and the cluster centers that are associated with their cluster labels  , for a suitable functional distance  . Typically, functional basis expansion[35][36][37][38][39] or FPCA[40] is used for dimension reduction. However, when covariance structures also play a role to distinguish clusters, taking mean functions as center is not adequate and FPC subspace-projected k-centers functional clustering approach, who uses subspaces as cluster centers, is taken into consideration.[41] Specifically, the Let   be the cluster membership variable, and the FPC subspace  ,  , assuming that there are   clusters. The projected function of   onto the FPC subspace   can be written as   One aims to find the set of cluster centers  , such that the best cluster membership of  ,  , is determined by minimizing the discrepancy between the projected function   and the observation  . The initial step of the subspace-projected clustering procedure uses only  , which reduces to the k-means functional clustering. In the subsequent iteration steps, the mean function and the set of eigenfunction for each cluster is updated and used to identify the set of cluster subspaces  , until iterations converge. This functional clustering approach simultaneously identifies the structural components of the stochastic representation for each cluster. Furthermore, model-based clustering[42] based on mixture models is widely used in clustering vector-valued multivariate data and has been extended to functional data clustering. In this approach, similarly to the k-means type of functional data clustering, typical mixture model-based approaches to functional data clustering in a first step project the infinite dimensional functional data onto low-dimensional subspaces. For example, we can apply functional clustering models based on Gaussian mixture[43][44][45] distributions to the natural cubic spline basis coefficients, with emphasis on clustering sparsely sampled functional data. Furthermore, random effects modeling also provides a model-based clustering approach, using mixed effects models with B-splines or P-splines.[37] For clustering longitudinal data, a linear mixed model for clustering using a penalized normal mixture as random effects distribution has been studied.[46] Bayesian hierarchical clustering also plays an important role in the development of model-based functional clustering.[47][48][49][50]

Most recent developments in FDA

edit
  • Functional designs
  • Domain selection problems
  • Dependent functional data, such as functional time series
  • Multivariate functional data
  • Spatially indexed functional data
  • “Second Generation” functional data

Applications

edit
  • Continuous tracking and monitoring of movement and health data
  • Traffic flow data continuously recorded over time by arrays of sensors
  • Continuously recorded climate and weather data
  • Transcription factor count modeling along the genome
  • Analysis of auction data
  • Volatility data


Functional data analysis (FDA) is a branch of statistics that analyzes data providing information about curves, surfaces or anything else varying over a continuum. In its most general form, under an FDA framework, each sample element of functional data is considered to be a function. The physical continuum over which these functions are defined is often time, but may also be spatial location, wavelength, probability, etc. Intrinsically, functional data are infinite dimensional. The high intrinsic dimensionality of these data brings challenges for theory as well as computation, where these challenges vary with how the functional data were sampled. However, the high or infinite dimensional structure of the data is a rich source of information and there are many interesting challenges for research and data analysis.

History

edit

Functional data analysis has roots going back to work by Grenander and Karhunen in the 1940s and 1950s.[51][52][53][54] They considered the decomposition of square-integrable continuous time stochastic process into eigencomponents, now known as the Karhunen-Loève decomposition. A rigorous analysis of functional principal components analysis was done in the 1970s by Kleffe, Dauxois and Pousse including results about the asymptotic distribution of the eigenvalues.[55][56] More recently in the 1990s and 2000s the field has focused more on application and understanding the effects of dense and sparse observations schemes. The term "Functional Data Analysis" was coined by James O. Ramsay.[57]

Mathematical formalism

edit

Random functions can be viewed as random elements taking values in a Hilbert space, or as a stochastic process. The former is mathematically convenient, whereas the latter is somewhat more suitable from an applied perspective. These two approaches coincide if the random functions are continuous and a condition called mean-squared continuity is satisfied. [58]

Hilbertian random variables

edit

In the Hilbert space viewpoint, one considers an  -valued random element  , where   is a separable Hilbert space such as the space of square-integrable functions  . Under the integrability condition that  , one can define the mean of   as the unique element   satisfying

 

This formulation is the Pettis integral but the mean can also be defined as   the Bochner sense. Under the integrability condition that   is finite, the covariance operator of   is a linear operator   that is uniquely defined by the relation

 

or, in tensor form,  . The spectral theorem allows to decompose   as the Karhunen-Loève decomposition

 

where   are eigenvectors of  , corresponding to the nonnegative eigenvalues of  , in a non-increasing order. Truncating this infinite series to a finite order underpins functional principal component analysis.

Stochastic processes

edit

The Hilbertian point of view is mathematically convenient, but abstract; the above considerations do not necessarily even view   as a function at all, since common choices of   like   and Sobolev spaces consist of equivalence classes, not functions. The stochastic process perspective views   as a collection of random variables

 

indexed by the unit interval (or more generally some compact metric space  ). The mean and covariance functions are defined in a pointwise manner as

 

(if   for all  ). We can hope to view   as a random element on the Hilbert function space  . However, additional conditions are required for such a pursuit to be fruitful, since if we let   be Gaussian white noise, i.e.   is standard Gaussian and independent from   for any  , it is clear that we have no hope of viewing this as a square integrable function.

A convenient sufficient condition is mean square continuity, stipulating that   and   are continuous functions. In this case   defines a covariance operator   by

 

The spectral theorem applies to  , yielding eigenpairs  , so that in tensor product notation   writes

 

Moreover, since   is continuous for all  , all the  's are continuous. Mercer's theorem then states that the covariance function   admits an analogous decomposition

 

Finally, under the extra assumption that   has continuous sample paths, namely that with probability one, the random function   is continuous, the Karhunen-Loève expansion above holds for   and the Hilbert space machinery can be subsequently applied. Continuity of sample paths can be shown using Kolmogorov continuity theorem.

Functional data designs

edit

Functional data are considered as realizations of a stochastic process   that is an   process on a bounded and closed interval   with mean function   and covariance function  . The realizations of the process for the i-th subject is  , and the sample is assumed to consist of   independent subjects. The sampling schedule may vary across subjects, denoted as   for the i-th subject. The corresponding i-th observation is denoted as  , where  . In addition, the measurement of   is assumed to have random noise   with   and  , which are independent across   and  .

1. Fully observed functions without noise at arbitrarily dense grid

edit

Measurements   available for all  

Often unrealistic but mathematically convenient.

Real life example: Tecator spectral data.[57]

2. Dense design with noisy measurements

edit

Measurements  , where   are recorded on a regular grid,

 , and   applies to typical functional data.

Real life example: Berkeley Growth Study Data and Stock data

3. Sparse design with noisy measurements (longitudinal data)

edit

Measurements  , where   are random times and their number   per subject is random and finite.

Real life example: CD4 count data for AIDS patients.[59]

Functional principal component analysis

edit

Functional principal component analysis (FPCA) is the most prevalent tool in FDA, partly because FPCA facilitates dimension reduction of the inherently infinite-dimensional functional data to finite-dimensional random vector of  . More specifically, dimension reduction is achieved by expansion the underlying observed random trajectories   in a functional basis that consists of the eigenfunctions of the (auto)-covariance operator on  . Consider the (auto)-covariance operator  ,  , which is a compact operator on Hilbert space. By Mercer's theorem, the kernel of  , i.e., the covariance function  , has spectral decomposition  , where the series convergence is absolute and uniform, and   are real-valued nonnegative eigenvalues in descending order with the corresponding orthonormal eigenfunctions   . By the Karhunen–Loève theorem, the FPCA expansion of an underlying random trajectory is  , where   are the functional principal components (FPCs), sometimes referred to as  . The Karhunen–Loève expansion facilitates dimension reduction in the sense that the partial sum converges uniformly, i.e.,   as   and thus the partial sum with a large enough   yields a good approximation to the infinite sum. Thereby, the information in   is reduced from infinite dimensional to a  -dimensional vector   with the approximated process:

 
(1)

Other popular bases include spline, Fourier series and wavelet bases. Important applications of FPCA include the modes of variation and functional principal component regression.

Functional linear regression models

edit

Functional linear models can be viewed as an extension of the traditional multivariate linear models that associates vector responses with vector covariates. The traditional linear model with scalar response   and vector covariate   can be expressed as

 
(2)

where   denotes the inner product in Euclidean space,   and   denote the regression coefficients, and   is a zero mean finite variance random error (noise). Functional linear models can be divided into two types based on the responses.

Functional regression models with scalar response

edit

Replacing the vector covariate   and the coefficient vector   in model (2) by a centered functional covariate   and coefficient function   for   and replacing the inner product in Euclidean space by that in Hilbert space  , one arrives at the functional linear model

 
(3)

The simple functional linear model (3) can be extended to multiple functional covariates,  , also including additional vector covariates  , where  , by

 
(4)

where   is regression coefficient for  , the domain of   is  ,   is the centered functional covariate given by  , and   is regression coefficient function for  , for  . Model (3) and (4) have been studied extensively.[60][61][62]

Functional regression models with functional response

edit

For a functional response   on   and multiple functional covariates  ,  , two major models have been considered.[63][64] One of these two models is generally referred to as functional linear model (FLM) which can be written as:

 
(5)

where   is the functional intercept, for   ,   is a centered functional covariate on  ,   is the corresponding functional slopes with same domain, respectively, and   is usually a random process with mean zero and finite variance.[63] In this case, at any given time  , the value of  , i.e.,  , depends on the entire trajectories of  . Model (5) are also studied extensively.[65][66][67][68][69]

Function-on-scalar regression

edit

In particular, taking   as a constant function yields a special case of model (5)

 
which is a functional linear model with functional responses and scalar covariates.

Concurrent regression models

edit

This model is given by,

 
(6)

where   are multiple functional covariates on  ,   are the coefficient functions defined on the same interval and   is usually assumed to be a random process with mean zero and finite variance.[63] This model assumes that the value of   depends on the current value of   only and not the history   or future value, hence it is a "concurrent regression model", which also has been referred to as "varying-coefficient" model. Various estimation methods have been proposed.[70][71][72][73][74][75]

Functional nonlinear regression models

edit

Direct nonlinear extensions of the classical functional linear regression models (FLMs) still involve a linear predictor, but combine it with a nonlinear link function, analogous to the idea of generalized linear model from the conventional linear model. Developments towards fully nonparametric regression models for functional data encounter problems such as curse of dimensionality. In order to bypass the “curse” and the metric selection problem, we are motivated to consider nonlinear functional regression models, which are subject to some structural constraints but do not overly infringe flexibility. One desires models that retain polynomial rates of convergence, while being more flexible than, say, functional linear models. Such models are particularly useful when diagnostics for the functional linear model indicate lack of fit, which is often encountered in real life situations. In particular, functional polynomial models, functional single and multiple index models and functional additive models are three special cases of functional nonlinear regression models.

Functional polynomial regression models

edit

Functional polynomial regression models may be viewed as a natural extension of the Functional Linear Models (FLMs) with scalar responses, analogous to extending linear regression model to polynomial regression model. For a scalar response   and a functional covariate   with domain   and the corresponding centered predictor processes  , the simplest and the most prominent member in the family of functional polynomial regression models is the quadratic functional regression[76] given as follows,

 
where   is the centered functional covariate,   is a scalar coefficient,   and   are coefficient functions with domains   and  , respectively. In addition to the parameter function β that the above functional quadratic regression model shares with the FLM, it also features a parameter surface γ. By analogy to FLMs with scalar responses, estimation of functional polynomial models can be obtained through expanding both the centered covariate   and the coefficient functions   and   in an orthonormal basis.[76][77]

Functional single and multiple index models

edit

A functional multiple index model is given as below, with symbols having their usual meanings as formerly described,

 
Here g represents an (unknown) general smooth function defined on a p-dimensional domain. The case   yields a functional single index model while multiple index models correspond to the case  . However, for  , this model is problematic due to curse of dimensionality. With   and relatively small sample sizes, the estimator given by this model often has large variance.[78][79]

Functional additive models (FAMs)

edit

Let   denote an expansion of a functional covariate   with domain   in an orthonormal basis  .


A functional linear model with scalar responses [as shown in model (2) of FLM] can be thus be written as follows,

 
One form of FAMs is obtained by replacing the linear function of   in the above expression ( i.e.,  ) by a general smooth function  , analogous to the extension of multiple linear regression models to additive models and is expressed as,
 
where   satisfies   for  .[80][64] This constraint on the general smooth functions   ensures identifiability in the sense that the estimates of these additive component functions do not interfere with that of the intercept term  .


Another form of FAM is the continuously additive model[81], expressed as,

 
for a bivariate smooth (i.e. twice differentiable) additive surface g :   which is required to satisfy   for all t in   in order to ensure identifiability.

Generalized functional linear model

edit

An obvious and direct extension of FLMs with scalar responses [shown in model (2)] is to add a link function so as to create a generalized functional linear model (GFLM)[82] by analogy to extending linear model to generalized linear model (GLM), of which the three components are:

  1. Linear predictor  ; [systematic component]
  2. Variance function  , where   is the conditional mean; [random component]
  3. Link function   connecting the conditional mean   and the linear predictor   through  . [systematic component]

Clustering and classification of functional data

edit

For vector-valued multivariate data, k-means partitioning methods and hierarchical clustering are two main approaches. These classical clustering concepts for vector-valued multivariate data have been extended to functional data. For clustering of functional data, k-means clustering methods are more popular than hierarchical clustering methods. For k-means clustering on functional data, mean functions are usually regarded as the cluster centers. Covariance structures have also been taken into consideration.[83] Besides k-means type clustering, functional clustering[84] based on mixture models is also widely used in clustering vector-valued multivariate data and has been extended to functional data clustering.[85][86][87][37][88] Furthermore, Bayesian hierarchical clustering also plays an important role in the development of model-based functional clustering.[89][90][91][92] Functional classification assigns a group membership to a new data object either based on functional regression or functional discriminant analysis. Functional data classification methods based on functional regression models use class levels as responses and the observed functional data and other covariates as predictors. For regression based functional classification models, functional generalized linear models or more specifically, functional binary regression, such as functional logistic regression for binary response, are commonly used classification approaches. More generally, the generalized functional linear regression model based on the FPCA approach is used[93]. Functional Linear Discriminant Analysis (FLDA) has also been considered as a classification method for functional data.[94][95][96][97][98]Functional data classification involving density ratios has also been proposed.[21] A study of the asymptotic behavior of the proposed classifiers in the large sample limit shows that under certain conditions the misclassification rate converges to zero, a phenomenon that has been referred to as "perfect classification".

Time warping

edit

Motivations

edit
 
Structures in cross-sectional mean destroyed if time variation is ignored. On the contrary, structures in cross-sectional mean is well-captured after restoring time variation.

In addition to amplitude variation[6], time variation may also be assumed to present in functional data. Time variation occurs when the subject-specific timing of certain events of interest varies among subjects. One classical example is Berkeley Growth Study Data[99], where the amplitude variation is the growth rate and the time variation explains the difference in children's biological age at which the pubertal and the pre-pubertal growth spurt occurred. In the presence of time variation, cross-sectional mean function may be an efficient estimate as peaks and troughs are located randomly and thus meaningful signals may be distorted or hidden.

Time warping, also known as curve registration[100], curve alignment or time synchronization, aims to identify and separate amplitude variation and time variation. If assume both components exist, then the functional data are viewed as a result of an underlying template function "warped" non-linearly in time by a smooth random process, known as time warping function. Let   denote the observed functions with both phase and amplitude variability and   to be the underlying functions where only the amplitude variation is present. The time warping function   removes the time variation such that  , where   are the realizations of an underlying time warping function that transforms the subject-specific time to the time scale of the template. The time warping functions   are assumed to be invertible and have average identity  . In most cases, the time is assumed to flow forward, so   are assumed to be strictly monotonic increasing. One exception where the time can flow backward has been presented in the context of modeling the declines in house price as time reversals[101].

The simplest case of a family of warping functions to specify phase variation is linear transformation, that is  , which warps the time of an underlying template function by subjected-specific shift and scale. More general class of warping functions includes diffeomorphisms of the domain to itself, that is, loosely speaking, a class of invertible functions that maps the compact domain to itself such that both the function and its inverse are smooth. The set of linear transformation is contained in the set of diffeomorphisms[102]. One challenge in time warping is identifiability of amplitude and phase variation. To break the non-identifiability requires specific assumptions.

Methods

edit

Earlier approaches include dynamic time warping (DTW) for applications such as speech recognition[103]. Another traditional method for time warping is landmark registration [104] [105], which aligns special features such as peak locations to an average location. Other relevant warping methods include pairwise warping[106], registration using   distance[102] and elastic warping[107].

Dynamic time warping

edit

The template function is determined through an iteration process, starting from cross-sectional mean, performing registration and recalculating the cross-sectional mean for the warped curves, expecting convergence after a few iterations. DTW minimizes a cost function through dynamic programming algorithm. Problems of non-smooth differentiable warps or greedy computation in DTW can be resolved by adding a regularization term to the cost function.

Landmark registration

edit

Landmark registration (or feature alignment) assumes well-expressed features are present in all sample curves and uses the location of such features as a gold-standard. Special features such as peak or valley locations or derivatives on the observed sample functions are aligned to their average locations on the template function[102]. Then the warping function is introduced through a smooth transformation from the average location to the subject-specific locations. A problem of landmark registration is that the features may be missing or hard to identify due to the noise in the data.

Extensions

edit

So far we considered scalar valued stochastic process,  , defined on one dimensional time domain.

Extension 1: The dimension of time domain

edit

The time domain of   can be extended from one dimension to multiple dimensions.

Extension 2: The range of the stochastic process

edit

The range set of the stochastic process may be extended from scalar values to vector values and further to nonlinear manifolds[108] and then to Hilbert spaces[109] and eventually to metric spaces[110].


R packages

edit

While the presentation of the models above assumes fully-observed functions, software is available for fitting the models with discretely-observed functions in software such as R.

See also

edit

Further reading

edit
  • Ramsay, J. O. and Silverman, B.W. (2005) Functional data analysis, 2nd ed., New York : Springer, ISBN 0-387-40080-X
  • Horvath, L. and Kokoszka, P. (2012) Inference for Functional Data with Applications, New York: Springer, ISBN 978-1-4614-3654-6
  • Hsing, T. and Eubank, R. (2015) Theoretical Foundations of Functional Data Analysis, with an Introduction to Linear Operators, Wiley series in probability and statistics, John Wiley & Sons, Ltd, ISBN 978-0-470-01691-6
  • Morris, J. (2015) Functional Regression, Annual Review of Statistics and Its Application, Vol. 2, 321 - 359, https://doi.org/10.1146/annurev-statistics-010814-020413
  • Wang et al. (2016) Functional Data Analysis, Annual Review of Statistics and Its Application, Vol. 3, 257-295, https://doi.org/10.1146/annurev-statistics-041715-033624

References

edit
  1. ^ Cardot, H; Ferraty, F; Sarda, P. (2003). "Spline estimators for the functional linear model". Statistica Sinica. 13 (3): 571–591.
  2. ^ Hilgert, N; Mas, A; Verzelen, N. (2013). "Minimax adaptive tests for the functional linear model". Annals of Statistics. 41: 838–869.
  3. ^ Kong, D; Xue, K; Yao, F; Zhang, HH. (2016). "Partially functional linear regression in high dimensions". Biometrika. 103 (1): 147–159.
  4. ^ Hu, Z; Wang, N; Carroll, RJ. (2004). "Profile‐kernel versus backfitting in the partially linear models for longitudinal/clustered data". Biometrika. 91 (2): 251–262.
  5. ^ Horváth, L; Kokoszka, P. (2012). Inference for functional data with applications. Springer Series in Statistics. Springer-Verlag.
  6. ^ a b c d Wang, JL; Chiou, JM; Müller, HG. (2016). "Functional data analysis". Annual Review of Statistics and Its Application. 3 (1): 257–295. Cite error: The named reference "wang:16" was defined multiple times with different content (see the help page).
  7. ^ Ramsay, J; Silverman, BW. (2005). Functional Data Analysis, 2nd ed. Springer.
  8. ^ Ramsay, JO; Dalzell, CJ. (1991). "Some tools for functional data analysis". Journal of the Royal Statistical Society: Series B (Methodological). 53 (3): 539–561.
  9. ^ Malfait, N; Ramsay, JO. (2003). "The historical functional linear model". The Canadian Journal of Statistics. 31 (2): 115–128.
  10. ^ He, G; Müller, HG; Wang, JL. (2003). "Functional canonical analysis for square integrable stochastic processes". Journal of Multivariate Analysis. 85 (1): 54–77.
  11. ^ Yao, F; Müller, HG; Wang, JL. (2005). "Functional data analysis for sparse longitudinal data". Journal of the American Statistical Association. 100 (470): 577–590.
  12. ^ He, G; Müller, HG; Wang, JL; Yang, WJ. (2010). "Functional linear regression via canonical analysis". Journal of Multivariate Analysis. 16 (3): 705–729.
  13. ^ Fan, J; Zhang, W. (1999). "Statistical estimation in varying coefficient models". The Annals of Statistics. 27 (5): 1491–1518.
  14. ^ Wu, CO; Yu, KF. (2002). "Nonparametric varying-coefficient models for the analysis of longitudinal data". International Statistical Review. 70 (3): 373–393.
  15. ^ Huang, JZ; Wu, CO; Zhou, L. (2002). "Varying-coefficient models and basis function approximations for the analysis of repeated measurements". Biometrika. 89 (1): 111–128.
  16. ^ Huang, JZ; Wu, CO; Zhou, L. (2004). "Polynomial spline estimation and inference for varying coefficient models with longitudinal data". Statistica Sinica. 14 (3): 763–788.
  17. ^ Şentürk, D; Müller, HG. (2010). "Functional varying coefficient models for longitudinal data". Journal of the American Statistical Association. 105 (491): 1256–1264.
  18. ^ Eggermont, PPB; Eubank, RL; LaRiccia, VN. (2010). "Convergence rates for smoothing spline estimators in varying coefficient models". Journal of Statistical Planning and Inference. 140 (2): 369–381.
  19. ^ Abraham, C; Cornillon, PA; Matzner‐Løber, E; Molinari, N. (2003). "Unsupervised curve clustering using B-splines". Scandinavian Journal of Statistics. 30 (3): 581–595.
  20. ^ Serban, N; Wasserman, L. (2005). "CATS: Clustering after transformation and smoothing". Journal of the American Statistical Association. 100 (471): 990–999.
  21. ^ a b c Coffey, N; Hinde, J; Holian, E. (2014). "Clustering longitudinal profiles using P-splines and mixed effects models applied to time-course gene expression data". Computational Statistics & Data Analysis. 71 (C): 14–29. Cite error: The named reference ":0" was defined multiple times with different content (see the help page).
  22. ^ Kayano, M; Dozono, K; Konishi, S. (2010). "Functional cluster analysis via orthonormalized Gaussian basis expansions and Its application". Journal of Classification. 27 (2): 211–230.
  23. ^ Giacofci, M; Lambert‐Lacroix, S; Marot, G; Picard, F. (2013). "Wavelet-based clustering for mixed-effects functional models in high dimension". Biometrics. 69 (1): 31–40.
  24. ^ Peng, J; Müller, HG. (2008). "Distance-based clustering of sparsely observed stochastic processes, with applications to online auctions". The Annals of Applied Statistics. 2 (3): 1056–1077.
  25. ^ Chiou, JM; Li, PL. (2007). "Functional clustering and identifying substructures of longitudinal data". Journal of the Royal Statistical Society: Series B (Statistical Methodology). 69 (4): 679–699.
  26. ^ Banfield, JD; Raftery, AE. (1993). "Model-based Gaussian and non-Gaussian clustering". Biometrics. 49 (3): 803–821.
  27. ^ James, GM; Sugar, CA. (2003). "Clustering for sparsely sampled functional data". Journal of the American Statistical Association. 98 (462): 397–408.
  28. ^ Jacques, J; Preda, C. (2013). "Funclust: A curves clustering method using functional random variables density approximation". Neurocomputing. 112: 164–171.
  29. ^ Jacques, J; Preda, C. (2014). "Model-based clustering for multivariate functional data". Computational Statistics & Data Analysis. 71 (C): 92–106.
  30. ^ Heinzl, F; Tutz, G. (2014). "Clustering in linear-mixed models with a group fused lasso penalty". Biometrical Journal. 56 (1): 44–68.
  31. ^ Angelini, C; Canditiis, DD; Pensky, M. (2012). "Clustering time-course microarray data using functional Bayesian infinite mixture model". Journal of Applied Statistics. 39 (1): 129–149.
  32. ^ Rodríguez, A; Dunson, DB; Gelfand, AE. (2009). "Bayesian nonparametric functional data analysis through density estimation". Biometrika. 96 (1): 149–162.
  33. ^ Petrone, S; Guindani, M; Gelfand, AE. (2009). "Hybrid Dirichlet mixture models for functional data". Journal of the Royal Statistical Society. 71 (4): 755–782.
  34. ^ Heinzl, F; Tutz, G. (2013). "Clustering in linear mixed models with approximate Dirichlet process mixtures using EM algorithm". Statistical Modelling. 13 (1): 41–67.
  35. ^ Abraham, C; Cornillon, PA; Matzner‐Løber, E; Molinari, N. (2003). "Unsupervised curve clustering using B-splines". Scandinavian Journal of Statistics. 30 (3): 581–595.
  36. ^ Serban, N; Wasserman, L. (2005). "CATS: Clustering after transformation and smoothing". Journal of the American Statistical Association. 100 (471): 990–999.
  37. ^ a b c Coffey, N; Hinde, J; Holian, E. (2014). "Clustering longitudinal profiles using P-splines and mixed effects models applied to time-course gene expression data". Computational Statistics & Data Analysis. 71 (C): 14–29.
  38. ^ Kayano, M; Dozono, K; Konishi, S. (2010). "Functional cluster analysis via orthonormalized Gaussian basis expansions and Its application". Journal of Classification. 27 (2): 211–230.
  39. ^ Giacofci, M; Lambert‐Lacroix, S; Marot, G; Picard, F. (2013). "Wavelet-based clustering for mixed-effects functional models in high dimension". Biometrics. 69 (1): 31–40.
  40. ^ Peng, J; Müller, HG. (2008). "Distance-based clustering of sparsely observed stochastic processes, with applications to online auctions". The Annals of Applied Statistics. 2 (3): 1056–1077.
  41. ^ Chiou, JM; Li, PL. (2007). "Functional clustering and identifying substructures of longitudinal data". Journal of the Royal Statistical Society: Series B (Statistical Methodology). 69 (4): 679–699.
  42. ^ Banfield, JD; Raftery, AE. (1993). "Model-based Gaussian and non-Gaussian clustering". Biometrics. 49 (3): 803–821.
  43. ^ James, GM; Sugar, CA. (2003). "Clustering for sparsely sampled functional data". Journal of the American Statistical Association. 98 (462): 397–408.
  44. ^ Jacques, J; Preda, C. (2013). "Funclust: A curves clustering method using functional random variables density approximation". Neurocomputing. 112: 164–171.
  45. ^ Jacques, J; Preda, C. (2014). "Model-based clustering for multivariate functional data". Computational Statistics & Data Analysis. 71 (C): 92–106.
  46. ^ Heinzl, F; Tutz, G. (2014). "Clustering in linear-mixed models with a group fused lasso penalty". Biometrical Journal. 56 (1): 44–68.
  47. ^ Angelini, C; Canditiis, DD; Pensky, M. (2012). "Clustering time-course microarray data using functional Bayesian infinite mixture model". Journal of Applied Statistics. 39 (1): 129–149.
  48. ^ Rodríguez, A; Dunson, DB; Gelfand, AE. (2009). "Bayesian nonparametric functional data analysis through density estimation". Biometrika. 96 (1): 149–162.
  49. ^ Petrone, S; Guindani, M; Gelfand, AE. (2009). "Hybrid Dirichlet mixture models for functional data". Journal of the Royal Statistical Society. 71 (4): 755–782.
  50. ^ Heinzl, F; Tutz, G. (2013). "Clustering in linear mixed models with approximate Dirichlet process mixtures using EM algorithm". Statistical Modelling. 13 (1): 41–67.
  51. ^ Grenander, U. (1950). "Stochastic processes and statistical inference". Arkiv för Matematik. 1 (3): 195–277.
  52. ^ Rice, JA; Silverman, BW. (1991). "Estimating the mean and covariance structure nonparametrically when the data are curves". Journal of the Royal Statistical Society. 53 (1): 233–243.
  53. ^ Müller, HG. (2016). "Peter Hall, functional data analysis and random objects". Annals of Statistics. 44 (5): 1867–1887.
  54. ^ Karhunen, K (1946). Zur Spektraltheorie stochastischer Prozesse. Annales Academiae scientiarum Fennicae.
  55. ^ Kleffe, J. (1973). "Principal components of random variables with values in a seperable hilbert space". Mathematische Operationsforschung und Statistik. 4 (5): 391–406.
  56. ^ Dauxois, J; Pousse, A; Romain, Y. (1982). "Asymptotic theory for the principal component analysis of a vector random function: Some applications to statistical inference". Journal of Multivariate Analysis. 12 (1): 136–154.
  57. ^ a b c Ramsay, J; Silverman, BW. (2005). Functional Data Analysis, 2nd ed. Springer.
  58. ^ Hsing, T; Eubank, R (2015). Theoretical Foundations of Functional Data Analysis, with an Introduction to Linear Operators. Wiley Series in Probability and Statistics.
  59. ^ Shi, M; Weiss, RE; Taylor, JMG. (1996). "An analysis of paediatric CD4 counts for acquired immune deficiency syndrome using flexible random curves". Journal of the Royal Statistical Society. Series C (Applied Statistics). 45 (2): 151–163.
  60. ^ Hilgert, N; Mas, A; Verzelen, N. (2013). "Minimax adaptive tests for the functional linear model". Annals of Statistics. 41: 838–869.
  61. ^ Kong, D; Xue, K; Yao, F; Zhang, HH. (2016). "Partially functional linear regression in high dimensions". Biometrika. 103 (1): 147–159.
  62. ^ Horváth, L; Kokoszka, P. (2012). Inference for functional data with applications. Springer Series in Statistics. Springer-Verlag.
  63. ^ a b c Wang, JL; Chiou, JM; Müller, HG. (2016). "Functional data analysis". Annual Review of Statistics and Its Application. 3 (1): 257–295.
  64. ^ a b Ramsay, J; Silverman, BW. (2005). Functional Data Analysis, 2nd ed. Springer.
  65. ^ Ramsay, JO; Dalzell, CJ. (1991). "Some tools for functional data analysis". Journal of the Royal Statistical Society: Series B (Methodological). 53 (3): 539–561.
  66. ^ Malfait, N; Ramsay, JO. (2003). "The historical functional linear model". The Canadian Journal of Statistics. 31 (2): 115–128.
  67. ^ He, G; Müller, HG; Wang, JL. (2003). "Functional canonical analysis for square integrable stochastic processes". Journal of Multivariate Analysis. 85 (1): 54–77.
  68. ^ Yao, F; Müller, HG; Wang, JL. (2005). "Functional data analysis for sparse longitudinal data". Journal of the American Statistical Association. 100 (470): 577–590.
  69. ^ He, G; Müller, HG; Wang, JL; Yang, WJ. (2010). "Functional linear regression via canonical analysis". Journal of Multivariate Analysis. 16 (3): 705–729.
  70. ^ Fan, J; Zhang, W. (1999). "Statistical estimation in varying coefficient models". The Annals of Statistics. 27 (5): 1491–1518.
  71. ^ Wu, CO; Yu, KF. (2002). "Nonparametric varying-coefficient models for the analysis of longitudinal data". International Statistical Review. 70 (3): 373–393.
  72. ^ Huang, JZ; Wu, CO; Zhou, L. (2002). "Varying-coefficient models and basis function approximations for the analysis of repeated measurements". Biometrika. 89 (1): 111–128.
  73. ^ Huang, JZ; Wu, CO; Zhou, L. (2004). "Polynomial spline estimation and inference for varying coefficient models with longitudinal data". Statistica Sinica. 14 (3): 763–788.
  74. ^ Şentürk, D; Müller, HG. (2010). "Functional varying coefficient models for longitudinal data". Journal of the American Statistical Association. 105 (491): 1256–1264.
  75. ^ Eggermont, PPB; Eubank, RL; LaRiccia, VN. (2010). "Convergence rates for smoothing spline estimators in varying coefficient models". Journal of Statistical Planning and Inference. 140 (2): 369–381.
  76. ^ a b Yao, F; Müller, HG. (2010). "Functional quadratic regression". Biometrika. 97 (1):49–64.
  77. ^ Horváth, L; Reeder, R. (2013). "A test of significance in functional quadratic regression". Bernoulli. 19 (5A): 2120–2151.
  78. ^ Chen, D; Hall, P; Müller HG. (2011). "Single and multiple index functional regression models with nonparametric link". The Annals of Statistics. 39 (3):1720–1747.
  79. ^ Jiang, CR; Wang JL. (2011). "Functional single index models for longitudinal data". he Annals of Statistics. 39 (1):362–388.
  80. ^ Wang, JL; Chiou, JM; Müller, HG. (2016). "Functional data analysis". Annual Review of Statistics and Its Application. 3 (1): 257–295.
  81. ^ Müller HG; Wu Y; Yao, F. (2013). "Continuously additive models for nonlinear functional regression". Biometrika. 100 (3): 607–622.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  82. ^ Müller HG; Stadmüller, U. (2005). "Generalized Functional Linear Models". The Annals of Statistics. 33 (2): 774–805.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  83. ^ Chiou, JM; Li, PL. (2007). "Functional clustering and identifying substructures of longitudinal data". Journal of the Royal Statistical Society: Series B (Statistical Methodology). 69 (4): 679–699.
  84. ^ Banfield, JD; Raftery, AE. (1993). "Model-based Gaussian and non-Gaussian clustering". Biometrics. 49 (3): 803–821.
  85. ^ James, GM; Sugar, CA. (2003). "Clustering for sparsely sampled functional data". Journal of the American Statistical Association. 98 (462): 397–408.
  86. ^ Jacques, J; Preda, C. (2013). "Funclust: A curves clustering method using functional random variables density approximation". Neurocomputing. 112: 164–171.
  87. ^ Jacques, J; Preda, C. (2014). "Model-based clustering for multivariate functional data". Computational Statistics & Data Analysis. 71 (C): 92–106.
  88. ^ Heinzl, F; Tutz, G. (2014). "Clustering in linear-mixed models with a group fused lasso penalty". Biometrical Journal. 56 (1): 44–68.
  89. ^ Angelini, C; Canditiis, DD; Pensky, M. (2012). "Clustering time-course microarray data using functional Bayesian infinite mixture model". Journal of Applied Statistics. 39 (1): 129–149.
  90. ^ Rodríguez, A; Dunson, DB; Gelfand, AE. (2009). "Bayesian nonparametric functional data analysis through density estimation". Biometrika. 96 (1): 149–162.
  91. ^ Petrone, S; Guindani, M; Gelfand, AE. (2009). "Hybrid Dirichlet mixture models for functional data". Journal of the Royal Statistical Society. 71 (4): 755–782.
  92. ^ Heinzl, F; Tutz, G. (2013). "Clustering in linear mixed models with approximate Dirichlet process mixtures using EM algorithm". Statistical Modelling. 13 (1): 41–67.
  93. ^ Leng, X; Müller, HG. (2006). "Classification using functional data analysis for temporal gene expression data". Bioinformatics. 22 (1): 68–76.
  94. ^ James, GM; Hastie, TJ. (2001). "Functional linear discriminant analysis for irregularly sampled curves". Journal of the Royal Statistical Society. 63 (3): 533–550.
  95. ^ Hall, P; Poskitt, DS; Presnell, B. (2001). "A Functional Data—Analytic Approach to Signal Discrimination". Technometrics. 43 (1): 1–9.
  96. ^ Ferraty, F; Vieu, P. (2003). "Curves discrimination: a nonparametric functional approach". Computational Statistics & Data Analysis. 44 (1–2): 161–173.
  97. ^ Chang, C; Chen, Y; Ogden, RT. (2014). "Functional data classification: a wavelet approach". Computational Statistics. 29 (6): 1497–1513.
  98. ^ Zhu, H; Brown, PJ; Morris, JS. (2012). "Robust Classification of Functional and Quantitative Image Data Using Functional Mixed Models". Biometrics. 68 (4): 1260–1268.
  99. ^ Gasser, T; Müller, HG; Kohler, W; Molinari, L; Prader, A. (1984). "Nonparametric regression analysis of growth curves". The Annals of Statistics. 12 (1): 210 -- 229.
  100. ^ Ramsay, JO; Li, X. (1998). "Curve registration". Journal of the Royal Statistical Society: Series B. 60 (2): 351–363.
  101. ^ Peng, J; Paul, D; Müller, HG. (2014). "Time-warped growth processes, with applications to the modeling of boom-bust cycles in house prices". The Annals of Applied Statistics. 8 (3): 1561–1582.
  102. ^ a b c Marron, JS; Ramsay, JO; Sangalli, LM; Srivastava, A (2015). "Functional data analysis of amplitude and phase variation". Statistical Science. 30 (4): 468–484.
  103. ^ Sakoe, H; Chiba, S. (1978). "Dynamic programming algorithm optimization for spoken word recognition". IEEE Transactions on Acoustics, Speech, and Signal Processing. 26: 43--49.
  104. ^ Kneip, A; Gasser, T (1992). "Statistical tools to analyze data representing a sample of curves". Annals of Statistics. 20: 1266–1305.
  105. ^ Gasser, T; Kneip, A (1995). "Searching for structure in curve sample". Journal of the American Statistical Association. 90 (432): 1179–1188.
  106. ^ Tang, R; Müller, HG. (2008). "Pairwise curve synchronization for functional data". Biometrika. 95: 875–889.
  107. ^ a b Anirudh, R; Turaga, P; Su, J; Srivastava, A (2015). "Elastic functional coding of human actions: From vector-fields to latent variables". Proceedings of the IEEE Conference on Computer Vision and Pattern Recognition: 3147-3155.
  108. ^ Dai, X; Müller, HG (2018). "Principal component analysis for functional data on Riemannian manifolds and spheres". The Annals of Statistics. 46 (6B): 3334–3361.
  109. ^ Chen, K; Delicado, P; Müller, HG (2017). "Modelling function-valued stochastic processes, with applications to fertility dynamics". Journal of the Royal Statistical Society. Series B (Statistical Methodology). 79 (1): 177–196.
  110. ^ Dubey, P; Müller, HG (2021). "Modeling Time-Varying Random Objects and Dynamic Networks". Journal of the American Statistical Association. 0 (0): 1–16.
  111. ^ Yao, F; Müller, HG; Wang, JL. (2005). "Functional data analysis for sparse longitudinal data". Journal of the American Statistical Association. 100 (470): 577–590.

Cite error: A list-defined reference named "Ramsay2005" is not used in the content (see the help page).
Cite error: A list-defined reference named "Grenander1950" is not used in the content (see the help page).
Cite error: A list-defined reference named "Müller2006" is not used in the content (see the help page).
Cite error: A list-defined reference named "Karhunen1946" is not used in the content (see the help page).
Cite error: A list-defined reference named "Rice2001" is not used in the content (see the help page).
Cite error: A list-defined reference named "Dauxois1982" is not used in the content (see the help page).
Cite error: A list-defined reference named "Refund" is not used in the content (see the help page).
Cite error: A list-defined reference named "FDboost" is not used in the content (see the help page).

Cite error: A list-defined reference named "Hsing2015" is not used in the content (see the help page).

Category:Regression analysis