https://github.com/cran/fields
Raw File
Tip revision: 6c8b30169bba182a68765ee3cb9b4e2ef7d38332 authored by Doug Nychka on 16 November 2011, 00:00:00 UTC
version 6.6.3
Tip revision: 6c8b301
Tps.Rd
% fields, Tools for spatial data
% Copyright 2004-2011, Institute for Mathematics Applied Geosciences
% University Corporation for Atmospheric Research
% Licensed under the GPL -- www.gpl.org/licenses/gpl.html

\name{Tps}
\alias{Tps}
\alias{fastTps}
\title{
  Thin plate spline regression 
}
\description{
Fits a thin plate spline surface to irregularly spaced data. The 
smoothing parameter is chosen by generalized cross-validation. The assumed 
model is additive  Y = f(X) +e  where f(X) is a d dimensional surface. 
This function also works for just a single dimension  and is a special case of a spatial process estimate
(Kriging).  A "fast" version of this function uses a compactly supported Wendland covariance and computes the estimate for a fixed smoothing parameter.
}
\usage{
Tps(x, Y, m = NULL, p = NULL, scale.type = "range", lon.lat = FALSE, miles = TRUE, ...)

fastTps(x, Y, m = NULL, p = NULL,  theta, lon.lat=FALSE,   ...)
}
\arguments{
To be helpful, a more complete list of arguments are described that are the 
same as those for the Krig function.  

\item{x}{
Matrix of independent variables. Each row is a location or a set of 
independent covariates. 
}
\item{Y}{
Vector of dependent variables. 
}
\item{m}{
A polynomial function of degree (m-1) will be  
included in the model as the drift (or spatial trend) component. 
Default is the value such that 2m-d is greater than zero where d is the 
dimension of x. 
}

\item{p}{
Polynomial power for Wendland radial basis functions. Default is 2m-d 
where d is the dimension of x.  
}

\item{scale.type}{
The independent variables and knots are scaled to the specified 
scale.type. 
By default the scale type is "range", whereby 
the locations are transformed  
 to the interval (0,1) by forming (x-min(x))/range(x) for each x. 
Scale type of "user" allows specification of an x.center and x.scale by 
the 
user. The default for "user" is mean 0 and standard deviation 1. Scale 
type of "unscaled" does not scale the data.  
}
\item{theta}{The tapering range that is passed to the Wendland compactly 
supported covariance. The covariance (i.e. the radial basis function) is 
zero beyond range theta.}
\item{lon.lat}{If TRUE locations are interpreted as lognitude and latitude and great circle distance is used to find distances among locations. The theta scale parameter for \code{fast.Tps}
(setting the compact support of the Wendland function) in this case is in units of miles (see example and caution below). }
\item{miles}{If TRUE great circle distances are in miles if FALSE distances are in kilometers}

\item{\dots}{
For \code{Tps} any argument that is valid for the \code{Krig} function. Some of the main ones
are listed below. 
For \code{fastTps} any argument that is suitable for the \code{mKrig} function 
see help on mKrig for these choices.

\describe{

\item{lambda}{
Smoothing parameter that is the ratio of the error variance (sigma**2) 
to the scale parameter of the  
covariance function. If omitted this is estimated by GCV. 
}

\item{Z}{Linear covariates to be included in fixed part of the model that are distinct
        from the default low order polynomial in \code{x}}

\item{df}{
The effective number of parameters for the fitted surface. Conversely,
N- df, where N is the total number of observations is the degrees of
freedom associated with the residuals.
This is an alternative to specifying lambda and much more interpretable.
}

\item{cost}{
Cost value used in GCV criterion. Corresponds to a penalty for
increased number of parameters. The default is 1.0 and corresponds to the
usual GCV.
}


\item{weights}{
Weights are proportional to the reciprocal variance of the measurement  
error. The default is no weighting i.e. vector of unit weights. 
}
\item{nstep.cv}{
Number of grid points for minimum GCV search. 
}
\item{x.center}{
Centering values are subtracted from each column of the x matrix. 
Must 
have scale.type="user".
}
\item{x.scale}{
Scale values that divided into each column after centering. 
Must 
have scale.type="user".
}
\item{rho}{
Scale factor for covariance. 
}
\item{sigma2}{
Variance of errors or if weights are not equal to 1 the variance is
sigma**2/weight. 
}

\item{method}{
Determines what "smoothing" parameter should be used. The default
is to estimate standard GCV
Other choices are: GCV.model, GCV.one, RMSE, pure error and REML. The
differences are explained below.
}
\item{verbose}{
If true will print out all kinds of intermediate stuff.  
}
\item{mean.obj}{
Object to predict the mean of the spatial process. 
}
\item{sd.obj}{
Object to predict the marginal standard deviation of the spatial process. 
}

\item{null.function}{
An R function that creates the matrices for the null space model.
The default is fields.mkpoly, an R function that creates a polynomial
regression matrix with all terms up to degree m-1. (See Details)
}

\item{offset}{
The offset to be used in the GCV criterion. Default is 0. This would be 
used when Krig/Tps is part of a backfitting algorithm and the offset has
to be included  to reflect other model degrees of freedom. 
}
}
}
}

\value{
A list of class Krig. This includes the 
fitted values, the predicted surface evaluated at the 
observation locations, and the residuals. The results of the grid 
search minimizing the generalized cross validation function are
returned in gcv.grid. Note that the GCV/REML optimization is 
done even if lambda or df is given. 
Please see the documentation on Krig for details of the returned 
arguments.  
}
\details{
Both of these functions are special cases of using the 
\code{Krig} and \code{mKrig} functions. See the help on each of these
for more information on the calling arguments and what is returned. 

A thin plate spline is result of minimizing the residual sum of 
squares subject to a constraint that the function have a certain 
level of smoothness (or roughness penalty). Roughness is 
quantified by the integral of squared m-th order derivatives. For one 
dimension and m=2 the roughness penalty is the integrated square of 
the second derivative of the function. For two dimensions the 
roughness penalty is the integral of  

        (Dxx(f))**22 + 2(Dxy(f))**2 + (Dyy(f))**22 

(where Duv denotes the second partial derivative with respect to u 
and v.) Besides controlling the order of the derivatives, the value of 
m also determines the base polynomial that is fit to the data. 
The degree of this polynomial is (m-1). 

The smoothing parameter controls the amount that the data is 
smoothed. In the usual form this is denoted by lambda, the Lagrange 
multiplier of the minimization problem. Although this is an awkward 
scale, lambda =0 corresponds to no smoothness constraints and the data 
is interpolated.  lambda=infinity corresponds to just fitting the 
polynomial base model by ordinary least squares.  

This estimator is implemented by passing the right generalized covariance
function based on radial basis functions to the more general function
Krig.  One advantage of this implementation is that once a Tps/Krig object
is created the estimator can be found rapidly for other data and smoothing
parameters provided the locations remain unchanged. This makes simulation
within R efficient (see example below). Tps does not currently support the
knots argument where one can use a reduced set of basis functions. This is
mainly to simplify the code and a good alternative using knots would be to use a
valid covariance from the Matern family and a large range parameter. 

CAUTION about \code{lon.lat=TRUE}: The option to use great circle distance
 to define the radial basis functions (\code{lon.lat=TRUE}) is very useful
 for small geographic domains where the spherical geometry is well approximated by a plane. However, for large domains the spherical distortion be large enough that the basis function no longer define a positive definite system and Tps will report a numerical error. An alternative is to switch to a three
dimensional thin plate spline the locations being the direction cosines. This will 
give approximate great circle distances for locations that are close and also the numerical methods will always have a positive definite matrices.

Here is an example using this idea for \code{RMprecip} and also some 
examples of building grids and evaluating the Tps results on them:
\preformatted{
# a useful function:
  dircos<- function(x1){
             coslat1 <- cos((x1[, 2] * pi)/180)
             sinlat1 <- sin((x1[, 2] * pi)/180)
             coslon1 <- cos((x1[, 1] * pi)/180)
             sinlon1 <- sin((x1[, 1] * pi)/180)
             cbind(coslon1*coslat1, sinlon1*coslat1, sinlat1)}
# fit in 3-d to direction cosines
  out<- Tps(dircos(RMprecip$x),RMprecip$y)
  xg<-make.surface.grid(fields.x.to.grid(RMprecip$x))
  fhat<- predict( out, dircos(xg))
# coerce to image format from prediction vector and grid points.
  out.p<- as.surface( xg, fhat)
  surface( out.p)
# compare to the automatic
  out0<- Tps(RMprecip$x,RMprecip$y, lon.lat=TRUE)
  surface(out0)
}

The function \code{fastTps} is really a convenient wrapper function that 
calls \code{mKrig} with the Wendland covariance function. This is 
experimental and some care needs to exercised in specifying the taper 
range and power ( \code{p}) which describes the polynomial behavior of 
the Wendland at the origin. Note that unlike Tps the locations are not 
scaled to unit range and this can cause havoc in smoothing problems with 
variables in very different units. So rescaling the locations \code{ x<- scale(x)} 
is a good idea for putting the variables on a common scale for smoothing.  
This function does have the potential to approximate estimates of Tps 
for very large spatial data sets. See \code{wendland.cov} and help on 
the SPAM package for more background.

See also the mKrig function for handling larger data sets and also for an example
of combining Tps and mKrig for evaluation on a huge grid. 

 }

\section{References}{
See "Nonparametric Regression and Generalized Linear Models"  
by Green and Silverman. 
See "Additive Models" by Hastie and Tibshirani. 
}
\seealso{
Krig, summary.Krig, predict.Krig, predict.se.Krig, plot.Krig, mKrig  
\code{\link{surface.Krig}}, 
\code{\link{sreg}} 
}
\examples{
#2-d example 

fit<- Tps(ozone$x, ozone$y)  # fits a surface to ozone measurements. 

set.panel(2,2)
plot(fit) # four diagnostic plots of  fit and residuals. 
set.panel()

# summary of fit and estiamtes of lambda the smoothing parameter
summary(fit)

surface( fit) # Quick image/contour plot of GCV surface.

# NOTE: the predict function is quite flexible:

     look<- predict( fit, lambda=2.0)
#  evaluates the estimate at lambda =2.0  _not_ the GCV estimate
#  it does so very efficiently from the Krig fit object.

     look<- predict( fit, df=7.5)
#  evaluates the estimate at the lambda values such that 
#  the effective degrees of freedom is 7.5
 

# compare this to fitting a thin plate spline with 
# lambda chosen so that there are 7.5 effective 
# degrees of freedom in estimate
# Note that the GCV function is still computed and minimized
# but the lambda values used correpsonds to 7.5 df.

fit1<- Tps(ozone$x, ozone$y,df=7.5)

set.panel(2,2)
plot(fit1) # four diagnostic plots of  fit and residuals.
          # GCV function (lower left) has vertical line at 7.5 df.
set.panel()

# The basic matrix decompositions are the same for 
# both fit and fit1 objects. 

# predict( fit1) is the same as predict( fit, df=7.5)
# predict( fit1, lambda= fit$lambda) is the same as predict(fit) 


# predict onto a grid that matches the ranges of the data.  

out.p<-predict.surface( fit)
image( out.p) 

# the surface function (e.g. surface( fit))  essentially combines 
# the two steps above

# predict at different effective 
# number of parameters 
out.p<-predict.surface( fit,df=10)


#A 1-d example  with confidence intervals
  out<-Tps( rat.diet$t, rat.diet$trt) # lambda found by GCV 
  out
  plot( out$x, out$y)
  xgrid<- seq(  min( out$x), max( out$x),,100)
  fhat<- predict( out,xgrid)
  lines( xgrid, fhat,)
  SE<- predict.se( out, xgrid)
  lines( xgrid,fhat + 1.96* SE, col="red", lty=2)
  lines(xgrid, fhat - 1.96*SE, col="red", lty=2)

# 
# compare to the ( much faster) B spline algorithm 
#  sreg(rat.diet$t, rat.diet$trt) 

# Here is a 1-d example with 95 percent  CIs  where sreg would not 
# work:
#  sreg would give the right estimate here but not the right CI's
  x<- seq( 0,1,,8)
  y<- sin(3*x)
  out<-Tps( x, y) # lambda found by GCV 
  plot( out$x, out$y)
  xgrid<- seq(  min( out$x), max( out$x),,100)
  fhat<- predict( out,xgrid)
  lines( xgrid, fhat, lwd=2)
  SE<- predict.se( out, xgrid)
  lines( xgrid,fhat + 1.96* SE, col="red", lty=2)
  lines(xgrid, fhat - 1.96*SE, col="red", lty=2)

# Adding a covariate to the fixed part of model
# Note: this is a fairly big problem numerically (850+ locations)

set.panel( 1,2)
# without elevation covariate
  Tps( RMprecip$x,RMprecip$y)-> out0 
  surface( out0)
  US( add=TRUE, col="grey")

# with elevation covariate
  Tps( RMprecip$x,RMprecip$y, Z=RMprecip$elev)-> out
# NOTE: out$d[4] is the estimated elevation coeficient
  out.p<-predict.surface( out, drop.Z=TRUE)
  surface( out.p)
  US( add=TRUE, col="grey")
  
# decomposing into the different pieces:
  fit<- predict( out)
  fit0<- predict( out, drop.Z=TRUE, just.fixed=TRUE)
  fit.elev<- predict( out, just.fixed=TRUE) - fit0
  fit1<- predict(out, drop.Z=TRUE) - fit0

  set.panel( 2,2)
  quilt.plot( out$x, fit0)
  title("spatial drift")
  quilt.plot( out$x, fit.elev)
  title("elevation")
  quilt.plot( out$x, fit1)
  title("spatial nonparametric")
  quilt.plot( out$x, fit)
  US( add=TRUE)
  title("full prediction")
  set.panel()

### 
### fast Tps
# m=2   p= 2m-d= 2
#
# Note: theta =3 degrees is a very generous taper range. 
# Use some trial theta value with rdist.nearest to determine a
# a useful taper. Some empirical studies suggest that in the 
# interpolation case in 2 d the taper should be large enough to 
# about 20 non zero nearest neighbors for every location.

set.panel( 1,2)
fastTps( RMprecip$x,RMprecip$y,m=2,lambda= 1e-2, theta=3.0) -> out2

# note that fastTps produces an mKrig object so one can use all the 
# the overloaded functions that are defined for the mKrig class. 
# summary of what happened note estimate of effective degrees of 
# freedom

print( out2)
surface( out2)
#
# now use great circle distance for this smooth 
# note the different  "theta" for the taper support  ( there are
# about 70 miles in one degree of latitude).
#
fastTps( RMprecip$x,RMprecip$y,m=2,lambda= 1e-2,lon.lat=TRUE, theta=210) -> out3
print( out3)  # note the effective degrees of freedom is different.
surface(out3)

set.panel()


#
# simulation reusing Tps/Krig object
#
fit<- Tps( rat.diet$t, rat.diet$trt)
true<- fit$fitted.values
N<-  length( fit$y)
temp<- matrix(  NA, ncol=50, nrow=N)
sigma<- fit$shat.GCV
for (  k in 1:50){
ysim<- true + sigma* rnorm(N) 
temp[,k]<- predict(fit, y= ysim)
}
matplot( fit$x, temp, type="l")


# 
#4-d example 
fit<- Tps(BD[,1:4],BD$lnya,scale.type="range") 

# plots fitted surface and contours 
# default is to hold 3rd and 4th fixed at median values 

surface(fit)   



}
\keyword{smooth}
% docclass is function
% Converted by Sd2Rd version 1.21.
back to top