0% found this document useful (0 votes)
31 views373 pages

Computational Statistics in Climatology

The document is a preface to the book 'Computational Statistics in Climatology' by Ilya Polyak, which discusses the application of statistical methods in climatology and the importance of computational statistics for climate studies. It outlines the structure of the book, including topics such as digital filters, averaging, random processes, and modeling of climatic data, while emphasizing the need for accurate statistical descriptions in understanding climate variability. The book aims to serve as a resource for specialists in statistical software development and climatologists, highlighting the ongoing challenges in the field of climatology.

Uploaded by

lbarreto
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
31 views373 pages

Computational Statistics in Climatology

The document is a preface to the book 'Computational Statistics in Climatology' by Ilya Polyak, which discusses the application of statistical methods in climatology and the importance of computational statistics for climate studies. It outlines the structure of the book, including topics such as digital filters, averaging, random processes, and modeling of climatic data, while emphasizing the need for accurate statistical descriptions in understanding climate variability. The book aims to serve as a resource for specialists in statistical software development and climatologists, highlighting the ongoing challenges in the field of climatology.

Uploaded by

lbarreto
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 373

Computational Statistics in Climatology

This page intentionally left blank


Computational
Statistics in
Climatology

Ilya Polyak

New York Oxford


OXFORD UNIVERSITY PRESS
1996
Oxford University Press
Oxford New York
Athens Auckland Bangkok Bogota Bombay
Buenos Aires Calcutta Cape Town Dar es Salaam
Delhi Florence Hong Kong Istanbul Karachi
Kuala Lumpur Madras Madrid Melbourne
Mexico City Nairobi Paris Singapore
Taipei Tokyo Toronto
and associated companies in
Berlin Ibadan

Copyright © 1996 Oxford University Press, Inc.


Published by Oxford University Press, Inc.,
198 Madison Avenue, New York, New York 10016
Oxford is a registered trademark of Oxford University Press
All rights reserved. No part of this publication may be reproduced,
stored in a retrieval system, or transmitted, in any form or by any means
electronic, mechanical, photocopying, recording, or otherwise,
without the prior permission of Oxford University Press.

Library of Congress Cataloging-in-Publication Data


Polyak, Ilya.
Computational statistics in climatology / Ilya Polyak.
p. cm.
Includes bibliographical references and index.
ISBN 0-19-509999-0
1. Climatology—Statistical methods. I. Title.
QC981.P63 1996
551.5'01'176—dc20 95-35132

987654321
Printed in the United States of America
on acid-free paper
Preface

The range of statistical applications in the atmospheric sciences is so wide


and diverse that it is difficult to mention even the most important areas of
study. The proceedings of the Conferences on Probability and Statistics in
the Atmospheric Sciences and of the Meetings on Statistical Climatology
number in the dozens, and several books were published more than a decade
ago (Gandin, 1965; Panofsky and Brier, 1968; Kagan, 1979; Epstein, 1985).
Recently published textbooks (Thiebaux, 1994; Wilks, 1995) basically con-
tain an introduction to the statistical methods, and an interesting mono-
graph by R. Daley (1991) shows the possibility of the mutual physical and
statistical analysis and modeling of atmospheric data. There are also a
monograph, Applied Statistics in Atmospheric Sciences, by Essenwanger
(1989) and a book by Preisendorfer and Mobley (1988) devoted solely to the
principal component analysis in meteorology and oceanography.
Although much work has been done in developing statistical climatol-
ogy, many important topics remain to be studied—for instance, topics in
the field of computational statistics and its climatological applications.
This book presents the advanced course on computational statistics
that I have taught for some years to Ph.D. students and scientists in the
Main Geophysical Observatory, Hydrological Institute, and Civil Engi-
neering Institute at St. Petersburg, Russia. The main objective of the course
is to progress from univariate modeling and spectral and correlation analy-
sis to multivariate modeling and multidimensional spectral and correlation
analysis. The course developed as a result of my many years of involvement
in a variety of projects dealing with design and development of software for
climate change studies by using statistic and random functions methods.
The elaboration of the corresponding algorithms, their application to ana-
lyzing and modeling climatic data, and consulting in the widespread area of
such methodologies have made it possible to gain an understanding of the
importance of various statistical disciplines for climate study. The two
major areas of application presented in this book are multidimensional
vi Preface

spectral and correlation analysis and multivariate autoregressive modeling.


Because the spectrum estimation methodology is based on the choice of
smoothing window, the book opens with a consideration of digital filters
(Chapter 1). Questions of averaging (as the dominant statistical procedure
in climatology) and fitting simple linear models are given principal attention
in Chapter 2. Estimations of spectral and correlation functions of random
processes and fields are discussed in Chapter 3. Algorithms of univariate
(Chapter 4) and multivariate (Chapter 5) modeling, potentially the most
important methodologies for climate study, conclude the consideration of
statistics.
Chapters 6, 7, and 8 are devoted to the analysis and modeling of cli-
matological data. (The climatological results given are limited to those
that were obtained with my participation.)
The goal of the methodological sections is to present numerical algo-
rithms in a form that can immediately be used for software development.
For example, the formulas for several classes of one- and two-dimension-
al smoothing and differentiating digital filters, each of which depends on
several parameters (width, length, order), can be easily programmed and
presented as a system of corresponding routines. The numerical schemes
for one-, two-, and multidimensional spectral and correlation analyses,
gathered in Tables 3.1 and 3.2, make it possible to design corresponding
software with a general focus on the application of some classes of filters
of different dimensionality. In many respects this work can serve as a
handy reference book for software development.
One of the features of the approach used is a numerical analysis of the
algorithms to gain insight into the theoretical accuracy (and limitations) of
different methods and to obtain maximum preliminary information for the
particular application. Such an approach is especially effective for linear
numerical procedures when the influence of data transformation, which
determines the standard error value to within some constant factor, on the
estimate accuracy can be analyzed before dealing with real observations.
The computational orientation of the text has necessitated that theo-
retical-statistical questions are only briefly outlined. Thus this special-
ized book cannot substitute for the study of more general theoretical
courses. Moreover, a certain level of knowledge in statistics and random
function theory has been assumed, and the material presented can help
the reader design software applications and efficiently use them. For
many of the algorithms considered, I have developed software in Fortran
and the Interactive Data Language (IDL) and have used them for cli-
mate studies.
Multiple examples with illustrations developed while teaching have a
special methodological meaning, not only as appropriate examples, but
Preface vii

also as general results that represent the accuracy of some classes of


methods (see, for instance, the figures in Chapter 4 with normalized stan-
dard deviations of the forecast errors for the entire class of first- and sec-
ond-order ARMA processes). Actually, most of the book consists of meth-
odological or climatological examples.
Statistical descriptions of climate represent an important aspect of
multipurpose scientific studies and provide a means for finding solutions
to many climatological problems. Historically, descriptions of climate have
been conducted by studying the average values (norms) of meteorological
observations from different points on the Earth. As climatology has devel-
oped, it has become necessary to consider not only averages but also nu-
merous other statistics that characterize spatial and temporal climate
variability. The most important topics in statistical climatology are prob-
lems that are (or can be) stated within the strict frameworks of theoreti-
cally founded statistical methods. With this in mind, the following group
of studies should be mentioned.
1. Climate modeling presents the very substantial problem of verify-
ing and confirming GCM (General Circulation Model) experiments as part
of the statistical identification of complicated stochastic systems. The
obvious requirement (to have identical probability distributions of the
observed and simulated data) led to the necessity of the estimation and
multiple comparison of the corresponding first- and second-order mo-
ments. The second moments can be presented in the form of the multi-
dimensional spectra and correlation functions or multivariate stochastic
models. Subsequent hypothesis testing and statistical inferences can help
in evaluating the closeness of the observed and simulated climate and
reveal the validity of GCMs. The subsequent physical analysis and inter-
pretation of the parameter matrices of the fitted models can become a
major diagnostic tool and can help solve some of the inverse problems for
hydrodynamic equations. In other words, at this stage of study, the inter-
pretation of the model parameter matrices as a system of interactions and
feedbacks of the analyzed processes is as important as the forecast. In
principle, according to system identification theory, statistical and physi-
cal modeling must result in a general model of the climate system.
2. Some engineering and practical projects require an analogous com-
parison of the probability distributions of data produced either by differ-
ent types of measurements (such as point gauges, radar, and satellite)
or different data collecting and averaging methodologies. For instance,
the ground truth problem for the Tropical Rainfall Measuring Mission
(TRMM) from satellite or the evaluation of the validity of different sur-
face air temperature data bases necessitates the estimation and compari-
son of second-moment statistics. Generally, most of the questions sur-
viii Preface

rounding retrieval, calibration, beam filling error analysis, and so on, are
the nonlinear statistical problems.
3. Another important direction of statistical modeling in climatology
that is an inseparable part of the GCM development arises in a wide vari-
ety of parametrization problems. It seems that the replacement of the
existing, pure empirical parametrization schemes by the multivariate and
multidimensional stochastic models can help to create scientifically sound,
more accurate and reliable physico-stochastic GCMs corresponding to the
real physico-stochastic character of the climate system.
4. Statistical (as well as physical) modeling of the climatic system is nec-
essary for developing climatology as a scientific discipline and, in particu-
lar, for describing the causes of climate change. The formulation of such
problems is based on the study of the Earth's history, which shows that cli-
mate changes have taken place on global and regional scales: ice-ages fol-
lowed by warming periods, damp regions turned into deserts and semi-
deserts, significantly fluctuating levels of spacious internal reservoirs,
growth and decay of glaciers, and so on. The dependence of every sphere of
human life on the climate makes it obvious that statistical descriptions of
climate constitute some of the most important scientific work of our time.
5. There are some other interesting global and climate change prob-
lems; for example, verification of the astronomical (Milankovitch) climate
theory, and the paleoclimate and tree-ring data analysis, which require
application of different statistical procedures.
Estimating the statistical moments of atmospheric processes (as the
components of a stochastic climate system) provides a means for develop-
ing stochastic models for any of their subsystems. The problem of such mod-
eling is to describe (theoretically and experimentally) the temporal evolu-
tion of the climatic system (or its statistical structure). But the methods of
computational statistics used for such modeling are based, as a rule, on the
assumptions that the variability of the corresponding stochastic system is
spatially and temporally stable (stationary and homogeneous) and that it
possesses a number of other theoretic properties (such as ergodicity and
normality). Of course, real processes of nature are far more complicated
than the above theoretical abstractions. Consequently, the results of the
estimation (for example, of the second moments) obtained within the frame-
work of any mathematical condition are only approximate, which might not
necessarily reflect reality in its details and complexity. Absolutely accurate
stochastic (as well as physical) models of natural processes are impossible
in principle. We can only design and fit some images, the goodness of fit of
which is, among other things, determined by the applied technique.
Nevertheless, conducting investigations using standard statistical
methodologies remains an important step in the development of climatol-
Preface ix

ogy. It also allows an orientation among multiple empirical approaches and


nonscientific interpretations of the results of estimation. Generally, the for-
mulation of climatological problems within the strict framework of funda-
mental statistical methods promotes the natural progress of climatology
along with the simultaneous theoretical development of such methods.
I do not pretend to present the entire range of statistical methods
used in climatology. Several very interesting, and, in some cases, more
advanced, branches of applied statistics (for example, nonlinear modeling,
principal component analysis, methods of objective analysis of meteoro-
logical fields, and others) are not considered here at all.
Of course, the reliability of climate statistics depends primarily on the
quality and volume of data. The climatological results of this book were
obtained using the most complete climate data bases available. These
include the surface air temperature of the Northern Hemisphere for 100
years of observations, special kinds of precipitation measurements such
as GATE rain rate data and PRE-STORM precipitation, historical records
(unique in duration) of temperature (315 years of observations) and pre-
cipitation (250 years of observations) in Central England, atmospheric
pressure in Basel (205 years of observations), the simulation data of one
of the best GCMs, and many others.
The second half of the book presents wide-ranging statistical results,
which describe the temporal and spatial statistical structure of the large-
scale climate fluctuations for both point gauges and spatially averaged
data. In many cases, the closeness of climate fluctuations to white noise or
to first-order multivariate AR models is discussed. We expect that limiting
the order of models (as well as their types) roughens our understanding of
climate, but this stage of climatology development cannot be bypassed.
Such an approach may either manifest the most general regularities of cli-
mate or give the first very rough approximation of climate as a stochastic
system with the possibility of analyzing the interactions and feedbacks of the
diverse climatological processes.
Presentation of the spectral and correlation estimates of different cli-
matological fields is restricted to the two-dimensional case (three-dimen-
sional figures) because pictorial representation of the four- (and more)
dimensional structures is too technically complex, given the book's frame-
work. The best medium for multidimensional illustrations is film, which
enables us to demonstrate variations of climate in the spaces of time, lags,
and frequencies (wave numbers) simultaneously.
In presenting the illustrations, I also tried to show a little bit of beau-
ty, as statistics sometimes seems to the uninitiated a dry and tedious sci-
ence. In this aspect an excellent IDL software developed by Research
Systems, Inc., was very helpful in drawing three-dimensional graphics.
x Preface

This book is intended for specialists in statistical software develop-


ment and application and for climatologists. I also hope that this book
(and especially its inevitable errors and mistakes) will attract the atten-
tion of specialists in theoretical statistics to climatological problems, the
solutions of which, in many cases, are a continuing challenge to modern
science.

I am grateful to Dr. Gerald North for his support, which made this work
possible. I thank Phyllis McBride for help with technical editing of the
first version of the book, and Stephen Cronin for help with the LaTex rep-
resentation of the formulas and the typing of several chapters.
I also appreciate the permission of the American Meteorological
Society to reproduce materials from the papers Polyak et al., 1994; Polyak
and North, 1995; and Polyak, 1996.
This work was made possible in part by NASA grants.

College Station, Texas I. E


December 1995
Contents

1. DIGITAL FILTERS 3
1.1 Gauss-Markov Theory 3
1.2 Polynomial Approximations 6
1.3 Variance of the Point Estimates 15
1.4 The Fourier Set 21
1.5 Examples 23
1.6 Smoothing Digital Filters 24
1.7 Regressive Filters 28
1.8 Harmonic Filters 39
1.9 Applications of Digital Filters 40
1.10 Differentiating Filters 44
1.11 Two-Dimensional Filters 51
1.12 Multidimensional Filters 58
2. AVERAGING AND SIMPLE MODELS 61
2.1 Correlated Observations 61
2.2 The Mean and the Linear Trend 66
2.3 Nonoptimal Estimation 75
2.4 Spatial Averaging 79
2.5 Smoothing of Correlated Observations 87
2.6 Filters of Finite Differences 91
2.7 Regression and Instrumental Variable 99
2.8 Nonlinear Processes 107
3. RANDOM PROCESSES AND FIELDS 110
3.1 Stationary Process 111
3.2 Cross-Statistical Analysis 117
3.3 Nonstationary Processes 122
3.4 Nonergodic Stationary Process 127
xii Contents

3.5 Time Series with Missing Data 129


8.6 Two-Dimensional Fields 131
3.7 Multidimensional Fields 138
3.8 Examples of Climatological Fields 147
3.9 Anisotropy of Climatological Fields 160
4. VAEIABILITY OF ARMA PROCESSES 162
4.1 Fundamental ARMA Processes 163
4.2 AR Processes 165
4.3 AR(1) and AR(2) Processes 167
4.4 Order of the AR Process 174
4.5 MA(1) and MA(2) Processes 179
4.6 ARMA(1,1) Process 183
4.7 Comments 186
4.8 Signal-Plus-White-Noise Type Processes 190
4.9 Process with Stationary Increments 198
4.10 Modeling the Five-Year Mean Surface Air Temperature 201
4.11 Nonstationary and Nonlinear Models 204
5. MULTIVARIATE AR PROCESSES 208
5.1 Fundamental Multivariate AR Processes 208
5.2 Multivariate AR(1) Process 212
5.3 Algorithm for Multivariate Model 218
5.4 AR(2) Process 223
5.5 Examples of Climate Models 225
5.6 Climate System Identification 234
6. HISTORICAL RECORDS 239
6.1 Linear Trends 239
6.2 Climate Trends over Russia 245
6.3 Periodograms 247
6.4 Spectral and Correlation Analysis 251
6.5 Univariate Modeling 259
6.6 Statistics and Climate Change 263
7. THE GCM VALIDATION 266
7.1 Objectives 267
7.2 Data 268
7.3 Zonal Time Series 271
7.4 Multivariate Models 279
7.5 The Diffusion Process 285
Contents xui

7.6 Latitude-Temporal Fields 291


7.7 Conclusion 299
8. SECOND MOMENTS OF RAIN 302
8.1 GATE Observations 303
8.2 PRE-STORM Precipitation 331
8.3 Final Remarks 344
References 348
Index 355
This page intentionally left blank
Computational Statistics in Climatology
This page intentionally left blank
1

Digital Filters

In this chapter, several systems of digital filters are presented. The


first system consists of regressive smoothing filters, which are a di-
rect consequence of the least squares polynomial approximation to
equally spaced observations. Descriptions of some particular univari-
ate cases of these filters have been published and applied (see, for
example, Anderson, 1971; Berezin and Zhidkov, 1965; Kendall and
Stuart, 1963; Lanczos, 1956), but the study presented in this chapter
is more general, more elaborate in detail, and more fully illustrated.
It gives exhaustive information about classical smoothing, differen-
tiating, one- and two-dimensional filtering schemes with their repre-
sentation in the spaces of time, lags, and frequencies. The results
are presented in the form of algorithms, which can be directly used
for software development as well as for theoretical analysis of their
accuracy in the design of an experiment. The second system consists
of harmonic filters, which are a direct consequence of a Fourier ap-
proximation of the observations. These filters are widely used in the
spectral and correlation analysis of time series.

1.1 Gauss-Markov Theory


The foundation for developing regressive filters is the least squares
polynomial approximation (of equally spaced observations), a prin-
cipal notion that will be considered briefly. We look first at the
independent and equally accurate observations

with conditions

3
0 / Digital filters

where /?; are unknown parameters which must be estimated and the
X{j are determined by the structure of the approximation function.
Independence and equal accuracy mean that

where V denotes variance, Y = {Yi}£_0, I is the identity matrix of


size (k + 1) X (k + 1), and a2 is the variance of the observations. In
matrix notation,

and the system of conditions can be written

To estimate the parameters (3j by the least squares method, the


quadratic form

must be minimized over f3j.


Differentiating (1.6) with respect to (3j yields a system of equa-
tions for finding the unknown parameters:

This system is called the normal system (or the normal equations),
and the matrix

is called the matrix of the normal system. C is a symmetric matrix


of size (ra + 1) X (m+1). Let us assume that matrix C is not singular
and that the solution of system (1.7) is

where
1.1 Gauss-Markov Theory 5

Vector 3 is a random vector (as a linear function of Y) with a


covariance matrix M^. Introducing the notation /3 = FY, where
F = (X T X)~ 1 X T , it is possible to obtain

If observations (1.1) are from a normal distribution, then an estima-


tor a2 of variance a1 (1.3) is given by the expression

The vector Y = {KJiU = Y - X/3 is called the vector of the point


estimates.^
The covariance matrix Mp of the random vector Y is

Although observations (1.1) are independent and equally accurate,


the point estimates are statistically dependent with covariance ma-
trix (1.13). Consider, for instance, the normalized (standardized)
matrix

and its diagonal elements

The elements of matrix (1.14) [which are proportional to the elements


of covariance matrix (1.13)] do not depend on either the observations
or their statistical characteristics. The elements are determined only
by the structure of the system of conditions (i.e., by the elements
of matrix X). For example, the polynomial approximations produce
a special class of conditional matrices. Therefore, during the design
0

stage of an experiment it is possible to analyze the matrices ( X X )


and X(X T X)-1 XT, which determine the statistical characteristics of
the parameters and point estimates (to within a factor a2). This pre-
liminary information enables us to make the necessary computations
6 I Digital Filters

more efficiently. Such an analysis can even provide enough infor-


mation to make a decision about the value of the suggested model
without the need for calculations.
The variances of the point estimates [the diagonal elements of
(1.13)] look like ff2of, where <r2 is a constant for given observations
but of depends on i and is determined by the matrix X. Exper-
imenting with different systems of conditions allows us to analyze
the dependence of of on i and to obtain information about the ac-
curacy of the point estimates Y = X/3 before any calculations are
conducted. The values of of are called normalized (or standardized)
variances of the point estimates; they differ from the variances of the
point estimates only by the factor a2.
Sometimes, we need to consider the statistical characteristics of
the residuals R = Y — X/3. The covariance matrix of R is

To check the accuracy of the computations, we can use the following


expression:

where 5 > ir (Mp) is a trace of the matrix Mp.


If observations VJ- are normal, the statistical significance of the
estimates of the parameters (3i is determined by the Student statistic
(Rao, 1973)

which follows a t distribution with k — m degrees of freedom.

1.2 Polynomial Approximations


In this section the least squares polynomial approximations of the
observations, given on the equally spaced grid of points, are consid-
ered. One convenient way to construct such polynomials is based on
a system of orthogonal polynomials (see Bolshev and Smirnov, 1965).
But for some practical purposes, such as designing digital filters, the
direct construction of the polynomials is preferred.
Consider the observations given on the sequence of equally spaced
points
1.2 Polynomial Approximations 1

The transformation

where r = k/2, (k + 1 = 2r + 1), changes this sequence to the


symmetrical grid points

If the number of observations is odd (i.e., k is even), the grid points


are

If the number of points is even (i.e., k is odd), the grid points are

The notation /;_r = Y{ (i = 0, 1,..., k) leads to the symmetrical


indexing of the observations:

If k is odd, the values /,-_,. have fractional subscripts, but these will
not cause any difficulties in our presentation.
The coefficients of the polynomial

under the condition

must then be estimated.


It is clear that

The normal equations in this case are


8 1 Digital Filters

afafkaflffsfsfjfklasjfsjfklsjfasjfklasjf

where

and
1.2 Polynomial Approximations 9

it is alsogvaelaskjastioaslasjfasjfklasot

and the normal system is separated into two independent systems of


the following orders: INT[(m + 2)/2] and INT[(m + l)/2]. l For
the same observations, polynomials of degrees m and m + 1 have
INT[(m + 2)/2] identical coefficients. Polynomial approximations
(to the same observations) with subsequent increasing (by one) de-
gree lead to the necessity of finding the solution to two independent
systems, of orders INT [(m + 2)/2] and INT [(m + l)/2], instead of
one system of order m + 1, as is necessary in general cases. In other
words, such polynomials with degrees 2p and 2p + 1 (p — 0,1,...)
have identical coefficients with even numbers. Polynomials with de-
grees 2p ~ f and 2p (p = 1,2,...) have identical coefficients with odd
numbers. This is one of the advantages of the symmetrical system
of points (1.20). To compute the elements of the normal system's
matrix (1.27), it is necessary to find the sums

if p is even. These sums can be calculated by the formula

l
INT(a) means the integer part of a. For example, INT(5/2) = 2.
10 1 Digital Filters

where j9 p + i(r+ 1) are the Bernoulli polynomials.2


The definition of the Bernoulli polynomials produces the equali-
ties

The Bernoulli polynomials Bp(x) are determined by

where B p are the Bernoulli numbers, defined by the formula

Values for the first few Bernoulli numbers can be found, for instance, in Berezin
and Zhidkov (1965).
Note the following properties of the Bernoulli polynomials:

Now we can find a sum (if p is even)

where the first equality is conditioned by (*), the third by (**), and the fourth
by (***).
1.2 Polynomial Approximations 11

which determine the matrix elements of the normal systems if


m = 0,1,2,3,4,5.
Therefore, the elements of the normal system's matrices are the
Bernoulli polynomials with argument r + 1 = k/2 + 1. This enables
us to find the analytical expressions for the elements of the inverse
matrix C"1 as functions of r. The inversion of the matrix C amounts
to the inversion of the two matrices

and

Note the formulas for the elements of the matrices


12 1 Digital Filters

(m = 0,1,2,3,4,5; m is the superscript, not the power)


that satisfy the requirements of practical applications:

where

where

and

where
1.2 Polynomial Approximations 13

it is alsovaetakltas

For example, for the second- and fourth-degree polynomials, the ma-
trices C"1 can be composed from the above elements as follows:
14 1 Digital Filters

For polynomials with m > 5, the matrices C ! can be found by the


method of dividing matrices into blocks [see Berezin and Zhidkov
(1965)], using the inverse matrices that have already been found for
the smaller values of m.
Sometimes, the least squares polynomial is used to estimate the
derivatives of the observed function. In this case, the polynomial
building can be carried out on the system of points

where h is the step length along the t axis. The elements of the
inverse matrices of the normal equations are determined by the for-
mula

where c"j is given by (1.35)-(1.41).


Concluding this section, it is natural to ask: for what do we really
need these simple, particular results?
First, when we work with a very long time series of observations,
the values of the elements of the normal systems' matrices can be
large enough either to cause overflow or to produce inaccuracy. This
must be taken into account when developing computer programs.
Second, the formulas save computer time whenever it is necessary
to find a large number of polynomials that approximate many dif-
ferent time series. Third, the analytical expressions for the elements
of the inverse matrices make it possible to investigate before the ex-
periment the accuracy (and correlations) for both the polynomial
coefficient and the point estimates. Finally, the availability of an-
alytical expressions for the elements of the inverse matrices of the
normal equations makes it possible to build classic smoothing and
differentiating filters, which are widely applied.
Consider the expressions for the polynomial coefficients needed
to obtain such digital filters. Because the polynomials of even and
odd degrees have a set of identical coefficients, we need formulas for
1.3 Variance of the Point Estimates 15

estimating only parameters /3p, if m is odd. From (1.28) and (1.35)-


(1.38) it follows that

it is alosfafklasjflkasdfklasfasas

where

Replacing the c*-"j in (1.46) by their expressions from (1.35)-(1.38)


produces simple formulas for parameters a"1'''p, for m = 0 , 1 , . . . , 5.

1.3 Variance of the Point Estimates


Expression (1.14) makes it possible to investigate the correlation of
the point estimates. However, the practical interpretation of this
correlation is often cumbersome. It is much simpler to determine
only the variances (1.15) and to draw the corresponding graphics,
which illustrate the accuracy of the point estimates.
Placing (1.24) in formula (1.15) yields the expression for comput-
ing the normalized variance as a function of the number of points.
16 1 Digital Filters

For example, formulas

where the values c^ are taken from (1.35) and (1.36), determine
the diagonal elements of the covariance matrix (1.14) of the point
estimates, if m = 1,2, and 3 respectively.
For the degree m polynomial, the general formula is

Formula (1.48) was derived by replacing the elements of the ma-


trices C"1 and X in (1.15) by (1.24), because the polynomials are
built on the system of points (1.20). From (1.47) and (1.48), it follows
that the normalized variance of any point estimate depends on this
point value, on the number of observations, and on the polynomial
degree. For different m and r, one can calculate the values of such
variances for all the points, then draw the corresponding graphics.
Figure 1.1 illustrates the shape of the variance curve of the point
estimates on the observational interval, when the degree is fixed
(m — 3) and the number of points varies (k = 10,12,14). (For a
visual demonstration, the discrete points in Fig. 1.1, as well as in
other figures, are joined by the continuous curves.) Figure 1.1 shows
that the variance curves are symmetrical relative to the vertical axis
and that they have identical shapes. The fewer the number of obser-
vations, the higher the curve location. If k —> m, the curves converge
into the straight line cr^_r = 1.
Figure 1.2 illustrates the shapes of the curves of variance (1.48)
for different polynomial degrees (m = 0 , 1 , . . . , 7) and for the fixed
number of points k (30). The curves have different shapes, and they
are symmetric relative to the vertical axis. The curves show that the
variance of the point estimates increases as the polynomial degree
increases. The values of the variance of point estimates close to the
ends of the interval are large (close to one); it is only by moving
1.3 Variance of the Point Estimates 17

Figure 1.1: Normalized variances of the point estimates when the polynomial
degree is fixed (m = 3) and the number of observations k varies.

toward the center of the interval that one can obtain estimates with
a small variance.
For the first- and third-degree polynomials, it is possible to show
[with the aid of formula (1.47) and the expressions for the correspond-
ing elements c\. and c^A that the minimal variance corresponds to
the center point of the interval. It is very likely that this assertion
will prove correct for any odd m. If m is even, the minimal variance
corresponds to some of the points that are close to the origin and
that are located symmetrically relative to the center of the interval.
If t — 0, the points on the graphics (Figs. 1.1 and 1.2) present
the value CQQ — a% /a2 (the normalized variance of the estimate of
flo). If the polynomial degree m and the number of points 2r + 1 are
known, one can find the value of CQ-Q and draw a conclusion (with the
aid of Fig. 1.2) about the shape and location of the variance curve
of the point estimates.
The graph in Figure 1.1 resembles a menorah, and Figure 1.2
and equations (1.47)-(1.48) can be used for designing menorahs of
18 1 Digital Filters

Figure 1.2: Normalized variances of the point estimates, when the number of
observations is fixed (A; = 30) and the polynomial degree m varies.

many types and shapes. Therefore, formula (1.48) can be called the
menorah equation.
Figure 1.3 illustrates the dependence of CQQ on the number of
points k for different degrees m. As this illustration clearly shows, in-
creasing the number of observations to more than 30 does not greatly
decrease the point estimate variance (when m < 5). Having rep-
resented the observation variance value, one can use Figure 1.3 to
determine the sample size, which is necessary to obtain estimate (]0
with the required accuracy.
The results of this section enable us to evaluate (within a constant
factor cr 2 ) the accuracy of the point estimates of the least squares
polynomials before making any calculations and estimations. For
example, if one wants to build a second-degree polynomial having 11
observations, one can refer to Figure 1.3 [or formula (1.36)], which
gives Cg0 = a2. /a2 « 0.2. Therefore, the point estimate variance is
Po
1.3 Variance of the Point Estimates 19

Figure 1.3: Dependence of the normalized variance of the /?o estimate on the
number of points and on the polynomial degree.

approximately one fifth of the observation variance. (Of course, the


result for the end points is worse.) These calculations provide some
quantitative results, which previously had been clear only intuitively:
1. The highest accuracy of the point estimates corresponds to the
central point of the approximation interval (if k + 1 is odd) or
to some points close to the center (if A; + 1 is even).
2. Formula (1.48) reveals that, outside the approximation inter-
val, the variance a^ of the extrapolated value of ft increases
as t2m with increasing t. Therefore, from a purely statistical
viewpoint, the least squares polynomial cannot be used for the
extrapolation.
3. Within the framework of the conditions outlined in this chap-
ter, for the variance a\ (&,m) (1.48) of any point i of the
approximational interval, the following inequalities hold true:
20 1 Digital Filters

If the inequality (1.50) is correct for at least one point, then it


is correct for all of the other points.
4. In spite of the fact that the observations have equal accuracy,
the point estimates have different variances that characterize
the nonstatioriary dependence. If the point estimates are used
for further processing, this fact must be taken into account,
especially when there are few observations.
5. The estimates of the coefficients fa with even and odd numbers
are not correlated because the system of normal equations is
separated into two independent systems.
Now let us consider the identification of the polynomial degree.
The least squares estimators have been derived under the following
condition:

If (1.51) is not satisfied, the above least squares estimates will be


biased. Condition (1.51) can be fulfilled by increasing the polyno-
mial degree. However, as has been shown, the higher the degree, the
greater the variance of the parameters and of the point estimates.
Moreover, increasing the polynomial degree increases the volume of
calculations. Finding the lowest polynomial degree for which condi-
tion (1.51) is satisfied is the basis for obtaining unbiased estimates.
An elaboration of the theoretical approach to this question has been
given by Anderson (1971) and Tutubalin (1972).
The test of the statistical significance of fa can be carried out
with the aid of (1.18). If one is interested in comparing fa with zero,
(1.18) and (1.35) --(1.38) yield

If k is large (k > 30), the ^-distribution is close to the normal


distribution, and the inequality

(where t K, 2 for the 5% or 3 for the 1% significance levels) must be


tested.
1.4 The Fourier Set 21

1.4 The Fourier Set


In this section the trigonometrical polynomials (Fourier sets) are con-
sidered briefly. For the Fourier approximation, the principal least
squares relationships, derived in Section 1.1, can be significantly sim-
plified, and all the formulas can be presented by simple and clear ex-
pressions without applying the matrix apparatus. This simplicity is
possible because of the orthogonal properties of the trigonometrical
polynomials.
Consider the principal formulas for estimating the Fourier coeffi-
cients and their statistical characteristics.
Let us assume that the observations

are given on the equally spaced system of points

where

The system of conditions is

where up = up; v is an integer (v < fc/2); and AQ, Ap, Bp (p =


1,2, . . . f V ) are unknown parameters that must be estimated. The
matrix C of the normal equations is

The trigonometric relationships that are necessary to prove this for-


mula can be found in Berezin and Zhidkov (1965). From (1.9) one
can obtain the formulas for the Fourier coefficients
22 1 Digital Filters

The elements Yt (t, = 0, ! , . . . , & — 1) of the vector of the point


estimates are

The parameter covariance matrix <7 2 C L reveals that the estimates


of the Fourier coefficients are statistically independent. From (1-11)
and (1.58) we get

The normalized variances of the Fourier coefficients do not depend


upon the polynomial degree. With the aid of (1.14), one can deter-
mine the elements of the covariance matrix of the point estimates
(1.60) by introducing the notation

Then
1.5 Examples 23

From (1.63) we get

where T = t — j.
Equation (1.64) shows that for the given k and y, the elements
of the covariancc matrix of the point estimates depend only upon
the difference t — j. Therefore, expression (1.64) determines the
covariance function of the stationary random process. This function
is an even periodic function of r = t — j with period k. Stationary
processes will be examined in more detail in Chapters 3 and 4.
The normalized variances of the point estimates (1.60), equal to
^^-, are identical for the given grid (1.56) and value of v; they do not
change from point to point as they do in the algebraic polynomial
approximation. Formula (1.64) shows that, for a fixed number of
observations k, an increase in the degree of v leads to an increase in
the variance of the point estimates. If A; becomes equal to the number
of parameters, this variance equals the observed variance value.

1.5 Examples
Now let us consider several examples of the graphs of the normalized
variances of the point estimates.
1. Figure 1.4 presents normalized variances of the point estimates
for the polynomial approximation of the unequally spaced ob-
servations. The number of points is 31, and the polynomial
degree rn varies from zero to six. The number of observations
to the right of zero is twice that of those to the left of zero. The
minimum value of the normalized variance is located to the
right of zero. The location and shape of the variance curves
depend upon the location of the points with observations on
the interval. It is apparent that the curves in Figure 1.4 are
the deformed curves in Figure 1.2.
2. Figure 1.5a shows graphs of the normalized variances of the
point estimates obtained for the approximation of equally spaced
observations by the function
24 1 Digital Filters

Figure 1.5b shows the analogous variances for the function

Assuming that x = el in (1.65) and that x = t l in (1.66), we


come to the problem of the polynomial approximation of the
observations given on the unequally spaced grid.

1.6 Smoothing Digital Filters


In general, when the observed curve has a complicated shape and
the number of observations is large, the problem of finding the ap-
propriate approximation function is not a simple one. Its solution
is theoretically attainable but practically cumbersome. One must
have some assumptions not only about statistical characteristics of
observations but also about the deterministic (geometric) structure
of the curve. The large diversity of observed natural processes and
of techniques that have different geometrical properties can lead to
a situation in which purely formal approximation cannot satisfy the
physical content of the problem. Moreover, sometimes the neces-
sity of smoothing arises in the process of computations (for example,
smoothing of the periodogram for the spectrum estimation of the
stationary random function) when the empirical curve shape is com-
pletely unknown. This circumstance requires setting a smoothing
problem in a more practically suitable but theoretically less rigorous
form.
First, let us not attempt to approximate the entire set of obser-
vations by one function, but rather consider its subsets. The sixes of
such subsets can be chosen small enough that initially there would be
no question about finding a specific type of approximation function.
This assumption can be best explained with the aid of formulas.
Consider the time series of the independent, equally spaced, and
equally accurate (with variance a 2 ) observations

given at the points


Figure 1.4: Normalized variances of Figure 1.5: Normalized variances of the point estimates for function (1.65) (left)
the point estimates for the unequally and for function (1.66) (right) with fixed k — 30.
spaced observations (k=30).
26 / Digital Filters

The mean value of Yt will be estimated by the formula

where t = r, r — 1,..., N - r; 2r + 1 = k; k « N; a_j — a,-.


The set of the coefficients (weights) {ttj}j__ r . will be called the
digital filter. The graph of the dependence of a.j on j is referred
to as the smoothing window (or the filter window), and the value
k = 2r + 1 is the window width (or the filter width). The smoothing
procedure [the process of finding estimates by (1.68)] is referred to
as the moving average procedure. The smoothing procedure (with
the specific weights a,-) makes it possible to decrease or to filter out
the high-frequency component of the fluctuations. This approach
is the particular case of the process of filtration, which is widely
used for modifying the Fourier components (of the Fourier expansion
of the observations) at any frequencies. The filtration procedure
is related to nonparametric methods of estimation. Its application
differs from the least squares approximation, which has a precise and
elegant theory. Smoothing does not (and possibly, cannot) have such
a theory, ft is a purely empirical approach to data processing, but the
application of digital filters is one of the important elements in the
methodology of computational statistics (see, for example, Anderson,
1971; Box and Jenkins, 1976; Kendall and Stuart, 1963; and Lanczos,
1956).
The filter parameters «j can be obtained from different physical
and statistical assumptions.
The scheme (1.68) cannot give the estimates for the first and
last r points of the set of observations (1.67). But this disadvantage
can be easily overcome, by, for example, the periodic extension of
the data (1.67) in both directions. It is also possible to work out
procedures, similar to (1.68) that compute estimates for the first
and last r points of the interval.
Let us study the statistical and harmonic structure of the set of
point estimates

obtained by (1.68).
The moving average procedure (1.68) is a linear transformation
of the random variables Yt:

where
1.6 Smoothing Digital Filters 27

and

The covariance matrix of the random vector Y is My = a FF .


This matrix has identical elements located on the diagonals, which
are parallel to the main diagonal.
By performing the necessary rearrangements in My = <r 2 FF 7 ,
the covariances MT of the point estimates (1.69) can be presented as

where r — q — js is the difference between the subscripts of the esti-


mates Ys and Yq, for which the covariance is computed.
Therefore, the moving average procedure changes the statistical
structure of the observations. The values of point estimates Ys and
Yq are statistically dependent (if \q - s\ — \T\ < 2r -f 1), with the
covariance determined by (1.72).
The variance a2 of the point estimates (1.69) is

which is less than the variance of the original observations (a1 < <r 2 ).
By combining (1.73) with the covariance function (1.72), it is
possible to determine the correlation function of the point estimates:
28 1 Digital Filters

The moving average procedure changes not only the statistical but
also the harmonic structure of the observations.
Let if/t = elwt be any of the harmonics of the Fourier expansion of
the observations (1.67). In accordance with (1.68), the point estimate
(ft is

that is, each harmonic with the frequency u is multiplied by the


factor

which does not depend on time /. I he resulting factor ol the fluctu-


ation amplitude with frequency u;, which is equal to

will be called the frequency characteristic of the filtering scheme


(1.68). By analyzing functions (1.76) and (1.77) for the given values
of a,-, it becomes possible to study the way in which the frequency
composition of (1.69) changes as compared with that of (1.67).
The basis for constructing different classes of filters is presented
by the character of the random and deterministic variations of the
observations (or by different statistical and physical assumptions).
The principal requirement is, of course, to provide the most accurate
estimation (unbiased and with minimal variance). One of the classes
of filters that approximately satisfied these conditions can be derived
from the least squares polynomials, which were considered in Section
1.2. These filters have been used in applications so widely and so long
that they can be called classical digital filters. They are analyzed in
the next section.

1.7 Regressive Filters


The least squares point estimates, which were examined in Section
1.3, have a minimal variance; it is convenient to use them as a basis
for constructing different filtering schemes. Such least squares esti-
mates are linear functions of the observations, and the coefficients
1.7 Regressive Filters 29

(weights) of these functions can be calculated. The analysis of point


estimates for polynomial approximations, which was made in Section
1.3, shows that the most accurate estimates are in the center of the
observational interval. Therefore, it makes sense to construct the
smoothing procedure for the central point of the interval.
Putting t - 0 and p = 0 in (1.22) and (1.45) yields

where

and c™2v are the elements (1.35)-(1.37) of the inverse matrix of the
system (1.27), which was obtained for the least squares polynomial
of degree m.
Filters (1.79) will be called regressive filters. For polynomials of
2p and 2p+l (p — 0,1,...) degrees, these filters have identical weights
because such polynomials have identical corresponding coefficients flo
(see Section 1.2). For this reason, parameters a"1'7 can be determined
only for odd m:

In Section 1.2, the formulas (1.35)-(1.37) for elements c% (for m < 5)


were obtained. Using these formulas, the expressions for a™' r are
noted in the following way:
For m = 0 and 1

and

For m = 2 and 3
30 / Digital Filters

and a in (1.68) is

For m = 4 and 5

and a in (1.68) is

Expressions (1.84) and (1.86) can be used as recipes for software


development. For instance, different smoothing numerical filters,
which are described by numerous authors (Anderson, 1971; Berezin
and Zhidkov, 1965; Box and Jenkins, 1976; Jenkins Watts, 1968; and
Lanczos, 1956), are the particular examples of formula (1.84). Those
filters can easily be obtained from expression (1.84) by substituting
the appropriate numerical value for r.
Figure 1.6 shows the shape of the smoothing windows of the re-
gressive filters as functions of m and r. Taking into consideration
(1.73) and (1.78), the variance of point estimates (1.69) can be pre-
sented as

The graphs of the dependence of c™0 on value k — 2r for different m


1.7 Regressive Filters 31

Figure 1.6: Dependence of the smoothing window shapes on the polynomial


degree (left) and on the filter width (right).

are given in Figure 1.3.


The covariance function of the point estimates, obtained by ap-
plying the regressive filter, can be determined by substituting pa-
rameters a™'r from (1.80) for a, in (1.72).
Such substituting yields:
32 1 Digital Filters

For example, for m = 1

Figures 1.7 and 1.8 show the graphs of the covariance functions
(1.88) (to within factor <r 2 ) of the point estimates (1.69) (the results
of smoothing by regressive filters) for different r and m = 1,3, and
5.
The correlation function corresponding to (1.88) is

For example, setting m — 1 in (1.90) yields

The dependence of the shape of the correlation function on the pa-


rameters m and r is illustrated in Figures 1.9 and 1.10.
The natural property of the regressive filters is that for any m
and r, the identity

is true. Really, substituting 1 for each / in (1.78) gives

Let us study the change of the observation Fourier transform am-


plitudes imposed by the smoothing procedure. For this purpose we
1.7 Regressive Filters 33

Figure 1.7: Dependence of the covariance function (1.88) of the estimates on


filter width (r) and polynomial degree (m).
34 I Digital Filters

Figure 1.8: Dependence of the covariance function of the estimates on degree


in when the filter widtli is fixed (2r + \ — 11).

will deduce the analytical expressions for the frequency characteris-


tics (1.77) of the regressive filters with TO = 1,3, and 5; then we will
obtain a general scheme for any m.
Substituting the coefficients determined by (1.82),(1.84), and (1.86)
for OLJ in (1.76) yields for m — 0 and 1:

where
1.7 Regressive Filters 35

Figure 1.9: Dependence of the correlation function (1.90) of the estimates oil
filter width (r) and polynomial degree 0
36 1 Digital Filters

Figure 1.10: Dependence of the correlation function of the estimates on degree


m, when the width is fixed (2r + 1 = 11).

and the frequency characteristic is

This is, of course, a well-known formula (see, for example, Hamming,


1973).
To infer the other formulas, one must notice that the derivatives
of the even orders of L(w) are
1.7 Regressive Filters 37

Figure 1.11: Frequency characteristics of the polynomial filters for different


degrees and fixed filter width (r = 5).

So

and

The graphics of the frequency characteristics (1.96), (1.98), and


(1.99) for different r and m are presented in Figures 1.11 and 1.12.
These illustrations show that smoothing by the regressive filter sup-
presses the power of the high-frequency harmonics. Increasing m or
decreasing r leads to an increase in the width of the corresponding
spectral window. These illustrations make it possible to choose dif
ferent filters depending on the required width of the low-frequency
part of the spectrum.
38 1 Digital Filters

Figure 1.12: Frequency characteristics of the polynomial filters for different


degrees (m) and filter widths (2r + 1).

The amplitudes of the Fourier transform of estimates (1.69) can


be obtained from the corresponding amplitudes of the Fourier trans-
form of observations (1.67) with the aid of multiplication by the
corresponding frequency characteristic.
To suppress the large power of the high frequencies (or retained
power leaked out through the side lobes of the frequency characteris-
tics), the set of observations can be smoothed several times. In this
case, the amplitudes are multiplied (the same number of times) by the
corresponding frequency characteristics. In such multiple smoothing,
the filtering schemes must be chosen so that the points of maxima
and minima of the side lobes of the frequency characteristics of the
two consecutive smoothing procedures will coincide.
However, it is important to notice that multiple smoothing can
lead to the misrepresentation of the real frequency composition of
the set of observations. This can occur because

and
l.8 Harmonic Filters 39

So in the vicinity of the u> axis origin, amplitudes do not change


markedly no matter what values of m and r are chosen. Formula
(1.100) shows that the point estimates, obtained by multiple smooth-
ing, converge to a constant because <pt —> 0, when u> ^ 0.
The general expression (for arbitrary m) for the frequency char-
acteristics of the regressive filters is obtained by substituting a"1^
(1.80) for aj in (1.77) and taking into account (1.97). This yields

where c^ are the elements of the inverse matrix of the normal equa-
tions.

1.8 Harmonic Filters


In the previous sections the regressive filters, obtained on the basis of
the least squares polynomial approximation, were analyzed in detail.
Now the analogous class of the Fourier set filters will be considered.
First, let us rewrite (1.60) in the following way. Substituting
expressions (1.59) for the parameters in (1.60) and rearranging yields

where
40 / Digital Filters

This means that the point estimates (1.60) represent the smoothing
values, obtained by the filter with the coefficients coinciding with the
covariance function (1.64) of the point estimates (to within factor
a').
Expression (1.103) determines the class of the filters, which de-
pends on the parameter (degree) v and number of observations k.
These filters will be referred to as the harmonic filters. The change
of the harmonic structure of the time series as the result of the har-
monic filter application is the same as in the case of the truncated
Fourier approximation: The terms with frequencies u p for numbers
p greater than v are omitted (filtered out). The covariance of point
estimates Yt and Yj is

where r — t — j.

1.9 Applications of Digital Filters


Designing, constructing, and developing filters, and then applying
them to solve a variety of practical problems, is just as significant
as identifying the function, which is used for approximating the ob-
servations. But applications of digital filters have some features and
difficulties caused by the fact that identification of the order and
the width depends on both statistical and deterministic (geometric)
structures of observations. The process of smoothing by a digital fil-
ter is an empirical procedure with solid theoretical ground. General
rigorous statistical inferences and hypothesis testings for evaluation
of the correspondence between filter parameters and data is difficult
to develop in this case. Identification of the regressive filter param-
eters is reduced to a compromise between the smallest possible m
(order) and the greatest possible r (width) that provide approxi-
mate unbiasedness and sufficiently small variance to the estimates.
Let us consider some questions of such identification.
1. For the most practical applications, the order of the filter
m < 5 is sufficient, and the problem is to choose m (1, 3, or 5) and
find an appropriately large width r, thus ensuring a small variance.
1.9 Applications of Digital Filters 41

The dependence of the estimate accuracy on m and r is determined


by (1.87); its graphical representation is given in Figure 1.3. As
a rule, for the final identification of the filter parameters, multiple
consecutive smoothings are needed by varying m and r. For each set
of m and r, a value

is computed, which is considered an approximate estimate of the


variance of the observations. If this estimate does not change (be-
comes stable) much for several values of r, the largest r is accepted
as an appropriate filter parameter. An approximate variance of the
estimate Yi can be computed with the aid of the following formula:

where aj are the parameters of the fitted filter. For the regressive
filters.

2. In the statistical literature, several digital filters are recom-


mended for a wide range of applications. But even with only one
filter, it is possible to construct an infinite number of different filters
by the superposition of its application. For example, if the time se-
ries is smoothed two times, and if each time the smoothing value is
the average of the two adjacent terms, it yields

This filter is the Tukey filter, whose correlation and spectral charac-
teristics are well-known (Jenkins and Watts, 1968).
3. Superposition of smoothing procedures can also be used by
applying different filters. As an example of constructing another
filter, let us apply any filter {otj}r-=_r to point estimates (1.102):
42 1 Digital Filters

The variance cf2 of random variable Yt is

where T = i — j and MT is determined by (1.104). By disregarding


the correlations of values (1.102), one can obtain an approximate
expression for the variance of estimates Yt:

Assuming a} = Cjtm, (1.108) yields

If the Tukey filter (1.105) is used in (1.106), then (1.108) gives

4. In order to compare different smoothing schemes, let us intro-


duce a notion for the equivalence of different filters. Two different
filters can be considered equivalent if their point estimates have equal
variances. As an example, we will find conditions for the equivalence
between the regressive filters and scheme (1.106), when «j is the
Tukey window. Equating expressions (1.87) and (1.110) yields (for
sufficiently large k)

For the commonly used regressive schemes, the relationships are


1.9 Applications of Digital Filters 43

Figure 1.13: The equivalence of the filters.

The graphic representation of (1.112) is given in Figure 1.13,


which can be used for choosing the width of the regressive filter, cor-
responding to a certain ratio v/k of the Fourier set in the scheme
(1.106). For example, if v/k in (1.106) is equal to 0.1, then in or-
der to obtain the estimator with the same variance by applying the
regressive filter, it is necessary that 2r + 1 w 13, if m = 1; or that
2r + 1 ~ 31, if m = 3; and so on.
Conversely, smoothing by a regressive filter with width 2r+l = 21
is equivalent to the ratio value in (1.106) of v/k ~ 0.05 , if m = 1,
or to v/k w 0.12, if m = 3. Obviously, such an approach does not
take into account any errors caused by smoothing that would lead to
a bias.
5. It is impossible to list all of the smoothing filters that are
used in applications. Identification of the filter parameters or con-
struction of a new one is dictated by the statistical character of the
observations as well as by the physical features of the problem under
consideration.
It is important to notice that the unsubstantiated or careless use
of the smoothing technique can distort the real information of the ob-
servations and result in incorrect conclusions. This is especially clear
AA 1 Digital Filters

in the illustrations of the frequency characteristics (Figs, 1.11and


1.12), which show the transformation of the frequency composition
of data by the process of smoothing (which is not always desirable).
These figures show that smoothing eliminates or reduces the high-
frequency part of the Fourier expansion of the observations. There-
fore, its application is justified only when the high frequencies are
not of any interest in the studies being performed.

1.10 differentiating Filters


Now let us consider the procedures for the moving average differenti-
ation of an equally spaced set of observations. The inference is based
on the least squares polynomial approximation.

1.10.1 The First-Order Derivative


We proceed from (1.22). The estimation of the mean of the first-
order derivative in the central point of an approximational interval
yields

The value of /3i can be accepted as the derivative estimate by the


moving average procedure for the different points of the observational
interval. As in the case of smoothing, it is impossible to apply this
procedure to the derivative estimation for the first and last r points
of the interval. To overcome this difficulty, one can estimate all of
the coefficients /3Z in order to differentiate the obtained polynomial
and calculate the derivatives for these points. Another approach to
finding a derivative for these points is the periodic expansion of the
observations.
The general formula (based on the least squares polynomials) for
estimating the first-order derivative is

where {7J n ' r K=_ 7 . is the set of the coefficients, which depend on the
polynomial degree m and number of points r,
1.10 differentiating Filters 45

Formula (1.115) is obtained from (1.45) under the assumption that

It is obvious that

and that the number of terms of the sum (1.114) is always even (for
the odd k + 1 = 2r + 1, the coefficient for f0 is 0).
Because the estimates of the coefficient /3i for degrees 2p — 1
and 2p (p = 1,2,...) polynomials are identical, their differentiating
schemes are identical, too.
Practical applied formulas for the parameters 7jm.r (m is from 1
to 5) are easily derived from expressions (1.35), (1.36), and (1.37).
For m = 1 and 2

For m = 3 and 4

For m = 5 and 6

where values cf- are determined by (1.37). Formulas (1.117)-(1.119)


can be used as recipes for software development.
Finding the variance a^ of the derivative point estimate obtained
by these moving average differentiating filters is not difficult because

and expressions for the c"\ (m is from 1 to 5) have already been


found in Section 1.2.
Illustration of the dependencies of values \/c^\ (normalized stan-
dard deviations of the point estimates of the derivatives) on m and
r are given in Figure 1.14.
46 1 Digital Filters

Figure 1.14: Dependence of the normalized standard deviation of the derivative


estimate on the filter width.

It is possible to investigate (1.115) in detail in order to determine,


for example, the covariance functions of the point estimates; but we
will forego this investigation so as not to complicate our discussion.
Let us carry out the harmonic analysis of transformation (1.114),
presenting it in the form

used in the applications. Let <p(t) = elwt be any of the harmonics of


the Fourier expansion of observations (1.67). According to (1.121),
1.10 differentiating Filters 47

The precise equality is obtained by immediately differentiating the


¥>(*)

The ratio

does not depend on t; it shows the effect of the moving average


differentiating procedure (1.121) on each harmonic of the derivative.
Replacing coefficients j"1'1" in (1.124) with formulas (1.117) to
(1.119), based on the least squares polynomials, yields analytical
expressions that show the transformation of each harmonic with a
frequency u> by the differentiating regressive filter.
If m = 1 and 2

where

To obtain other formulas, one must notice that the 2v-order


derivative of the function p(w) is

If m = 3, then from (1.124) and (1.118),


48 1 Digital Filters

Figure 1.15: Function J m,r (w) for different m and the fixed filter width (r
5).

If m = 5, formulas (1.124) and (1.119) yield

Illustrations of functions (1.125), (1.128), and (1.129) for different m


and r are provided in Figures 1.15 and 1.16.
Notice that formula (1.124) can be presented through the ele-
ments c of the inverse matrix of the normal equations for the least
squares polynomials obtained above. By replacing 7J"'r in (1.124)
with (1.115) and by taking into account (1.127), one gets
1.10 differentiating Filters 49

Figure 1.16: Function Jm,r(u) for different r and m.

1.10.2 The Second-Order Derivative


Let us obtain a system of linear filters for estimation of the mean of
the second-order derivative. From (1.22) we have

where

A variance of the derivative estimate (1.131) is

For m = 3, (1.131) and (1.36) yield


50 1 Digital Filters

For m = 5, (1.131) and (1.37) give

Formulas (1.133) and (1.134) can be used to develop software in


problems connected with estimation of the second-order derivative
of equally spaced observations.

1.10.3 Any Order Derivative


Estimation of the mean of any order derivative can be based on least
squares polynomials as well. For example, for the order p derivative,
computed at the central point of the observational interval, one has
[from (1.22)]

where flp is determined by expression (1.44).


Therefore, the order p derivative estimator is

where ajmrp is determined by (1.46).


The variance of of estimate (1.1.36) is equal to

Numerical differentiation schemes were derived for the grid point


equal to one. If the grid step is equal to any number h, estimate
(1.135) must be divided by hp to get a derivative in the appropriate
units of measurement.
Let us carry out a harmonic analysis of transformation (1.136).
If </?(t) = clujt is any of the harmonics of the Fourier expansion of
observations Yt, then (1.136) gives
1.11 Two-Dimensional Filters 51

Because

the ratio

reveals the influence of transformation (1.136) on the terms of the


Fourier expansion of the observations. For example, (1.124) is a
particular case of expression (1.139) for the first-order derivative.
The statistical analysis of estimates (1.136) can be carried out
analogously to the approach that was developed in Section 1.8 for
smoothing schemes, but such consideration is very formal and will be
omitted. Notice that for the application of the various differentiating
schemes, which can be obtained from (1.136) for the fixed numeri-
cal values of p, m, and r, such an analysis can help to reveal the
transformation of a statistical structure of the observations and to
evaluate its significance for the physical problem under consideration.
The discussion (in Section 1.10) of some aspects of the application
of smoothing filters is also valid for differentiating filters.
An accurate identification of the differentiating filter parameters
is very important because the derivative estimate can have a very
large error, exceeding a corresponding observation value by several
times. As a rule, all the quantities used for the derivative estimation
should have a double precision type.

1.11 Two-Dimensional Filters


Consider the two-dimensional smoothing and differentiating schemes
that are analogous to the one-dimensional filters described in the
previous sections.

1.11.1 Smoothing Filters


The smoothing (estimating the mean) of the two-dimensional field
is necessary for solving many applied statistical problems connected
with processing observations that are dependent on two variables.
In particular, the two-dimensional filter is necessary for estimating a
spectrum of a two-dimensional homogeneous random field, for con-
ducting the spectral analysis of the nonstationary processes, and for
many other problems.
52 1 Digital Filters

Let us derive a formula for the point estimate YIV of the mean of
an element Yiv of the random field

under the assumption that its elements are independent, have equal
variances er2, and are given on a system of two-dimensional equally
spaced grid points, with the grid interval equal to one along each
axis.
The point estimate (smoothing value) of the mean of Y\v is pre-
sented as the weighted sum of the observations

in the vicinity of YIV. The optimum weights can be obtained with the
aid of the least squares two-dimensional polynomials, which approxi-
mate observations (1.141), analogous to the one-dimensional schemes
introduced in Section 1.2. Choosing a symmetric two-dimensional
coordinate system [analogous to (1-20)] in the vicinity of YIV, one
can ensure that half of the elements of the corresponding normal
matrix (that is, the sums with odd powers) will be equal to zeroes
[see (1.29)]. In other words, the least squares two-dimensional poly-
nomials of degree q of two variables, T and t, are built on the grid
(r = —s, — ( s — 1 ) , . . . , s — l , s; t = —r, — (r — 1 ) , . . . , r —1, r). Omitting
the formal (and cumbersome) algebraic transformations, we note the
resulting formulas for q = 1 and q = 3.
The smoothing scheme is

where weights

if q = 1.
If q — the weights are

where
1.11 Two-Dimensional Filters 53

From (1.143) and (1.144), it follows that the filter weights are sym-
metric relative to the coordinate system origin. This relationship is
also shown in Figure 1.17 for the two-dimensional window (1.144) for
the fixed values of the widths s and r.
It is easy to show that for any s and r, the identity

is true.
The variance a% of point estimate Y[v is

where for q — 1

and for q = 3, CQ is determined by (1.145).


If the widths of filter (1.144) are identical (s = r) along both
axes, then

For example, if s = r = 3 in (1.153), one can evaluate the two-


dimensional third-order polynomial filter
54 1 Digital Filters

\ /

The two-dimensional filters can be superimposed (multiple smooth-


ing) in exactly the same way as the one-dimensional filter. Such a
possibility is especially important for the first-order (q = 1) filters
because it enables us to obtain two-dimensional filters with positive
weights. [If q = 3 and greater, the filter weight matrices have some
negative values; see above example (1.154) and Fig. 1.17.] Multiple
smoothing by the first-order two-dimensional filters can be used, for
example, for solving an important class of problems: the spectrum
estimation of the two-dimensional random field. Filtering with even
a few negative weights is sometimes not appropriate for this purpose
because they can give unacceptable negative estimates of the spectral
density.
As an example of building a filter with positive weights, con-
sider the procedure of two-tuples smoothing by the simplest two-
dimensional first order (q = 1) filter of size (2r + .1) X (2s + 1) = 2 X 2.
The resulting matrix of weights is

This filter, the two-dimensional generalization of the Tukey filter,


will be used for estimating the two-dimensional spectra of the homo-
geneous random fields.

1.11.2 differentiating Filters


Now, consider the numerical differentiation of a random field (1.140).
The estimate Y/v of the mean of the derivative dY/dr, where T is the
variation along axis /, is given by
1.11 Two-Dimensional Filters 55

Figure 1.17: Weights of the third-order (q = 3) two-dimensional smoothing


filter for s — 3, r = 2 (top) and s = 15, r = 3 (bottom).
56 1 Digital Filters

where for q = 1 the weights are

For q = 3 the weights are

where

Figure 1.18 shows the shape of the window of the third-order (q = 3)


differentiating filter (1.158) when s and r are fixed. This filter is a
two-dimensional odd function with zero values for r = 0.
The variance a\ of estimate (1.156) is

And, finally, for estimating the mean of the second-order derivative


$ 2 F/$r 2 , if q — 3, one can obtain

where
1.11 Two-Dimensional Filters 57

Figure 1.18: Weights of the third-order (q = 3) two-dimensional differentiating


filter for s = 3, r = 1 (top) and s = 15, r — 3 (bottom).
58 1 Digital Filters

As an example, Figure 1.19 presents the windows of this filter for


the fixed widths s and r.
This filter is an even two-dimensional function with equal weights
corresponding to the fixed values of T.
The variance <r^ of estimate (1.163) is

The estimators considered in this section were obtained with the aid
of cumbersome algebraic arrangements, which have been omitted be-
cause of their purely formal character. The correlation and harmonic
analysis of the smoothing and differentiating filters of random fields
also demands complicated algebraic transformations, which can be
done analogously to those provided for the one-dimensional schemes
considered in Sections 1.6, 1.7, and 1.10.

1.12 Multidimensional Filters


The development and analysis of different filters is one of the inter-
esting aspects of computational statistics with a wide area of appli-
cations, The scope of problems to be solved has not been exhausted
by the analyses that have been carried out in the previous sections.
For instance, we did not consider many specific filters (for example,
a Kalman filter) that are widely used in many applications, or ques-
tions connected with studying the bias of estimates, or many other
problems of filtration. It is necessary to develop a theory of filters
with positive weights for estimating spectral density. Some of the
questions of filtering of correlated observations will be considered
briefly in Chapter 2. Further studies of different problems in na-
ture, science, and engineering will inevitably lead to the creation of
filters with varying dimensionality for smoothing and differentiating
corresponding random fields.
The general scheme of smoothing a ^-dimensional random field
^ii,i2, -, </i' which is given on the equally spaced system of grid
points, can be noted as follows:
1.12 Multidimensional Filters 59

Figure 1.19: Weights of the third-order (q = 3) two-dimensional differentiating


(for the second derivative) filter for s = 3, r = 2 (top) and a — 15, r = 3
(bottom).
60 / Digital Filters

One approach to determining the weights in (1.168) is to build the


least squares ^-dimensional polynomial and consequently obtain the
expressions for the coefficients (/3o or any other /?,-, when the deriva-
tives are estimated) in the form of a linear combination of observa-
tions.
2
Averaging and Simple
Models

Simple linear procedures for processing correlated observations are


considered and interpreted in this chapter. Primarily, they present
different schemes for averaging data. These procedures are important
because climatology has historically dealt with spatial and temporal
averaging of statistically dependent meteorological observations. The
accuracy of such averaging is determined by the volume of data arid
by its correlation structure. The examples presented in this chap-
ter illustrate the level of accuracy that can be achieved within the
framework of some assumptions about such correlation structure.

2.1 Correlated Observations


Let us consider the principal relationships of the least squares method
for the statistically dependent observations (see Rao, 1973). The
basic assumptions are as follows.
We are given a vector of observations

with a covariance matrix

and a system of conditions


62 2 Averaging and Simple Models

where X{j (i = 0,1, . . . , & ; j = 0,1,..., m) are fixed values; /3j (j —


0 , 1 , . . . , m) are unknown parameters.
Estimators of these parameters can be found by minimizing (over
the unknown parameters (3j) the quadratic form

where

The normal equation system in this case is

where

Let f3 be the solution of the system (2.5). Then C * is the covariance


matrix of vector /3.
If the covariance matrix of the observations (2.1) can be presented
as M = <T 2 B, where B is the correlation matrix and a is the variance,
then the estimator of a is

For covariance matrix Mp of point estimates Y = X/3, one can ob-


tain the following formula:

The diagonal elements of this matrix

are the variances of the point estimates.


As was shown in Section 1.1, for the independent equally accurate
observations the normalized variances of the least squares parameter
estimates are determined by the type of the approximation function
and by the number of points. In the case of statistically dependent
data, such variances are additionally dependent on the covariances;
therefore, each type of covariance matrix requires special considera-
tion.
2.1 Correlated Observations 63

Example 1
Let us consider the graphs of the normalized variances (2.8) of the
point estimates for the polynomial approximation of the correlated
observations, given on an equally spaced symmetrical grid (identical
to those considered in Sections 1.2 and 1.3). Let

and

For a = 0.1 and k = 30, the graph of the normalized variances (2.8)
of point estimates for TO — 0, !,...,? are given in Figure 2.1. The
location of the curves on this graph is higher than the location of
the corresponding curves for independent observations in figure 1.2
because correlated observations contain less information about an
expected value than independent observations; and, consequently,
the variances of the estimates are greater. Figure 2.1 shows the
variance curve shapes for different m and for the given level of the
correlation a. If the number of observations k increases, the point
estimate variance curves, while retaining their shape for each TO, shift
down to the horizontal axis.
If t = 0, then /(O) = /3o, and the variance, which corresponds
to the central point of the interval, is equal to the variance of ai. .
Pa
Figure 2.2 presents the dependence of <r~ on the level of correlation
A>
a of the observations. The interval and the number of points are
the same as in Figure 2.1. The polynomial degree m varies from 0
to 7. It is apparent that the point estimate variances of statistically
dependent observations are greater than the corresponding variances
of independent observations for identical TO and k (compare Figure
2.2 with Figure 1.3). If k is fixed, the minimum values for each curve
are equal to crH. computed for the independent observations.
A>
Example 2
Let us draw the variances of the point estimates for an example
with the third-degree polynomial approximation of the statistically
independent unequally accurate observations given on grid points t —
—15, — 1 4 , . . . , 15 (k — 30). Suppose that the observation variances
are 1 for all points except point t = —5. Consider two cases: In the
first, the variance in point t = — 5 is 100; in the second, this vari-
64 2 Averaging and Simple Models

Figure 2.1: Normalized variances of the point estimates for the correlated
observations for different polynomial degrees (m).
2.2 The Mean and the Linear Trend 65

Figure 2.2: Normalized variances of the /?o estimate for different polynomial
degree (m) and correlation level (a) of the observations.

ance is 0.01. (The variances of the observations are located along the
main diagonal of the covariance matrix M; all other elements of ma-
trix M are zero.) Figure 2.3 shows the curves of the point estimate
variances for these examples, as well as the curve of the variances
corresponding to the independent equally accurate observations with
the same number of points k = 30 and the same polynomial degree
m = 3. We can see that the low accuracy of at least one of the ob-
servations (f?.5 = 100) increases the variances of the point estimates
in the vicinity of the bad point (compared with the equally accu-
rate observations). The high accuracy of at least one observation
(<7?.5 = 0.01) leads to an analogous decrease of the point estimate
variances, which is illustrated by curve (3) in Figure 2.3.
66 2 Averaging and Simple Models

Figure 2.3: Normalized variances of the point estimates of the unequally


accurate observations. (1) Variance of the observation of point t = —5 is 100. (2)
Equal accuracy observations. (3) Variance of the observation of point t = — 5 is
0.01.

2.2 The Mean and the Linear Trend


The most widespread application of the least squares method consid-
ered in Section 2.1 is averaging correlated data. The matrix expres-
sions in this case are simplified, and the formulas can be presented in
scalar form. A knowledge of the statistical structure of data provides
the foundation for designing and developing the optimal averaging
procedure. The accuracy of averages will be demonstrated by simple
examples.
Let us introduce the following notation:

If one estimates the mean, matrix X is a vector of size k + 1 with all


elements equal to one. Vector f3 is scalar fJQ. Therefore, (2.5) yields
2.2 The Mean and the Linear Trend 67

The estimator for the mean is

where

VarianceCT?-of the estimator Y is obtained by taking into account


the fact that C"1 is the covariance matrix of vector (3:

In the particular case of independent unequally accurate observations


(Mij = 0, if i / j), (2.11) and (2.13) yield

and

Let us determine the optimal estimator for the mean of the ran
dom variables with the covariance matrix
68 2 Averaging and Simple Models

where B is the correlation matrix.

Example 1
First consider the observations with exponential correlations

This idealized structure is given as an illustrative model to obtain a


comparative representation of the possible accuracy of the optimal
and nonoptirnal mean estimators in the case of a large sample.
It is easy to verify that

where 7 = 6 a is the correlation coefficient for the lag i — j = 1.


Adding all the elements of this matrix and substituting the results
in (2.13) yields

From (2.11) and (2.18) we get the mean estimator

According to (2.19), the accuracy of the mean estimator depends


upon the number of observations (k) and on the level of correlations
(a).
The normalized standard deviation
2.2 The Mean and the Linear Trend 69

Figure 2.4: Normalized standard deviations for the optimal estimator of the
mean of the observations with correlations (2.17).

of the estimator (2.20) for different correlations 7 and number of


points (k) is presented in Figure 2.4.
If k increases, the accuracy also increases without limit, in spite
of the fact that the observations are statistically dependent.

Example 2
As a second example, let us study the accuracy of the optimal
mean estimator when the elements of matrix B in (2.16) are deter
mined in the following way:

It is possible to show that the elements of the inverse matrix


70 2 Averaging and Simple Models

are equal to

Substituting the /<,j in (2.13) for these values yields

For sufficiently large k.

and the variance of the sample mean actually does not depend on
the number of observations. Therefore, if the correlations of the
observations are constant, then, beginning with some value of k,
increasing the number of observations in order to increase accuracy
of the estimation is senseless.
The optimal estimator of the mean for this example is the same
as for the independent observations

The normalized standard deviation

of the estimator (2.26) is illustrated in Figure 2.5. From (2.24) it fol-


lows that, even if p is small (approximately 0.1 to 0.2), the accuracy
of the sample mean is very far from the accuracy which, it seems,
could be obtained if one has several dozen (or several thousand) ob-
servations.
In Table 2.1 the mean and the linear trend parameter estimators
and corresponding variances are given for two situations: indepen-
dent and statistically dependent observations with the correlations
of the above considered type (2.17).
Example 3
Now let us consider a typical climatological example.
We are given 31 daily temperature observations (in °C)
Table 2.1: The mean and the linear trend parameter estimators.

Par. Independent observations Observations with correlation function e - air| (e- Q = 7)

Mean E(Yi) = const.

Continued)
Table 2.1 (Cont.)
Linear trend E(Yi)
2.2 The Mean and the Linear Trend 73

Figure 2.5: Normalized standard deviations for the optimal estimator of the
mean of the observations with correlations (2.21).

7.4 6.7 3.7 4.7 3.4 1.4 2.6 2.5 3.2 5.8
7.2 6.2 8.9 7.2 5.5 4.4 4.9 4.7 4.4 3.5
4.2 5.0 5.2 5.5 3.6 3.4 6.5 6.4 8.8 6.1 4.3

for August 1972 in Hveravellir, Iceland (data are taken from Tong,
1990).
Let us estimate the mean, the standard deviation a, and the stan-
dard deviation a~ of the mean estimate of temperature for August
1972 in Hveravellir. Assuming that the sample presents independent
and equally accurate observations of daily temperatures, we can ap-
ply corresponding formulas from Table 2.1.
The estimates are
74 2 Averaging and Simple Models

Now we want to estimate the climatic mean of the August tem-


perature in Hveravellir. Assuming that the above data presents a
sample of a stationary process with autocorrelation function 7~ T ,
where T = i — j, and i (and j) is the day number, we can apply the
corresponding formulas for correlated observations from Table 2.1.
For instance, setting 7 = 0.7, we get

The following table gives analogously derived estimates for differ-


ent values of autocorrelation r (the sample autocorrelation coefficient
of this data for r = 1 day is 0.606).

7 0.00 0.10 0.20 0.40 0.50 0.60 0.70 0.80 0.90 0.95

Y 5.07 5.08 5.09 5.11 5.12 5.14 5.18 5.23 5.36 5.50

ff
Y
0.32 0.34 0.36 0.43 0.50 0.61 0.80 1.17 2.15 3.70

The numerical results show that statistically dependent observa-


tions carry less information about the mean value than independent
observations and their standard deviation crp is greater than for in-
dependent observations.

Example 4
Consider estimates of the linear trend parameter ft\ obtained
through the seasonal and annual mean surface air temperature time
series of the Northern Hemisphere for the period from 1891 to 1975
(data from the archives of the World Data Center, Obninsk, Russia).
The observations were spatially averaged over various 15° latitude
bands (see Table 2.2). It was assumed that the correlation of the
time series terms is exponential (2.17) with the first autocorrelation
7 = 0.3 (see Chapter 6). Each time series was approximated by the
straight line /?0 + i/2i. As the results in Table 2.2 show there are only
six statistically significant (t-statistic is greater than 2) estimates for
the seasonal time series and two for the annual data. Temperature
rate is more noticeable for the winter and summer seasons and for
the northern latitudes.
These time series were also used to find the least squares straight
lines in the assumption of independence. The estimate values in
2.3 Nonoptimal Estimation 75

Table 2.2: Estimates of /?i (C°/year) and corresponding t-statistics


for the zonal temperature time series of different seasons (p\ = 0.3).

Winter Spring Summer Fall Annual


Lat. band fli estimates

90-75° 0.026 0.001 0.011 -0.002 0.009


75-60° 0.015 0.004 0.006 0.001 0.007
60-45° 0.004 0.005 0.002 0.002 0.004
45-30° 0.004 0.004 0.003 0.002 0.003
30-15° 0.003 -0.001 -0.003 0.001 0.000

t-statistics

90-75° 2.5 0.2 3.2 -0.3 1.8


75-60° 2.5 0.9 2.5 0.3 2.2
60-45° 0.8 1.8 1.4 0.2 1.8
45-30° 1.3 2.8 1.9 1.4 3.0
30-15° 2.4 -0.3 -1.6 0.5 0.2

this case were approximately the same as in Table 2.2 but the stan-
dard deviations were greater that changed the statistical significance
markedly: overall, 16 values of /—statistics were greater than 2.
More general results for the monthly mean surface air tempera-
tures of different geographical and political regions of the world are
given in Section 4.8 (Table 4.2) and Chapters 6 and 7. Many of these
estimates are statistically significant.
But no statistically significant estimates were found among lin-
ear trend parameter values of climatic times series of some Russian
stations which were carefully maintained and controlled by several
generations of Russian climatologists.

2.3 Nonoptirnal Estimation


When the covariance matrix is unknown, the calculations are often
performed by applying the same method used for independent obser-
76 2 Averaging and Simple Models

vations (see Section 2.1). Such procedure is not statistically optimal


(in other words, it does not provide the minimal variance); but, in
many situations, if the volume of data is large, it is close to the
optimal estimation.
Let us derive the formulas for the covariance matrices of the pa-
rameters and point estimates for the nonoptimal estimation. Assume
that observations (2.1) with an unknown covariance matrix (2.2) are
used for estimation of vector 3 under conditions (2.3). As an esti-
mator of 3, one takes the solution of the system

which was obtained in Section_l.l.


The covariance matrix of 3 (as a linear function of observations
Y) is

Covariance matrix My- of point estimates Y — X/3 is

The diagonal elements of this matrix

determine the variances of the nonoptimal point estimates.


This scheme is often applied to estimating the mean of the cor-
related observations with unknown matrix M. The estimator of the
mean

in this case has a variance

We will use this scheme for averaging observations with the par-
ticular type covariances M,-j, which depend only on \j — i\ (i.e.,
M;; = MI,-_,-I). Then matrix M is presented as
2.3 Nonoptimal Estimation 77

Such a matrix is called the Toeplitz matrix.


The variance of the mean of the observations with a covariance
matrix of (2.33) type is derived from (2.32):

Example 1
Let us compare the variances of the sample means obtained by
the various methods considered in Sections 1.1, 2.1, and 2.3. If the
mean of the random variables with the unknown correlation matrix
(2.17) is estimated by (2.31), then (2.34) yields

For sufficiently large k, (2.35) gives

The comparison of the variances (2.19) and (2.36) illustrates the


possibility of obtaining a nonoptimal mean estimator with a variance
that (for a large number of observations) is only slightly distinguished
from the variance of the optimal estimator. Moreover, it is possible to
show that for this example, the numerical values of variances (2.19)
and (2.35) do not differ markedly even for a small amount of data.
These examples also reveal the importance of the independence (for
estimation of the expected value) of the observations, because in this
case, the variances of the estimates are minimal. The distinction
of the variance of the nonoptimal scheme from the optimal one are
affected by the covariance values and the number of observations.
Formula (2.31) is used for estimating the mean of the time series
(statistically dependent observations with an unknown covariance
matrix) of the stationary random process.
As for the example (2.21), the estimator (2.31) (and its variance)
is the same as that for optimal and nonoptimal approaches. To some
extent, the example (2.21) is the idealized illustration of the practical
78 2 Averaging and Simple Models

meteorological situation when the spatial or temporal span of the


collected data is relatively small; indeed, the correlations between
the observations can be assumed to be approximately constant. For
instance, rain data obtained by averaging radar measurements within
a small elemental data box is consistent with the above assumptions.

Example 2
Let us draw the graphs of the variances of the nonoptimal point
estimates for the polynomial approximation of the observations with
exponential correlations:

and

Figure 2.6 shows the shape of the variance curves (2.30) of the nonop-
timal point estimates, when the polynomial degrees m and the level
of correlation a are fixed but the number k varies. Figure 2.7 presents
the graph of the normalized variances (2.80) for the fixed k and a
(A: = 30, a = 0.1) and for the different polynomial degree m. Figures
2.6 and 2.7 show that the variances of the nonoptimal point esti-
mates (for the points close to the interval ends) can be greater than
the variances of the original observations.
A comparison of Figures 1.2, 2.1, and 2.7 offers a representation
of the shape and location of the variance curves of the independent
and correlated observations when different approaches are used.

Example 3
The examples in Figure 2.8 are provided so that the accuracy of
the considered methodologies may be compared more clearly. This
figure presents the graphs of the normalized variances of the point
estimates, when k = 14, a = 0.3, and for polynomial degrees m = 1
and 3. Curve 1 was obtained for independent observations (1.15);
curve 2 corresponds to the optimal scheme (2.8); and curve 3 is
the result of application of formula (2.30). The graphs show that
for these types of correlations the variances (2.30) of the nonoptimal
point estimates only slightly exceed corresponding minimal variances
(2.8).
2-4 Spatial Averaging 79

Figure 2.6: Normalized variances (2.30) of the nonoptimal point estimates


when the polynomial degree m and a are fixed (m = 3, a = 0.1) and the number
of points k varies.

2.4 Spatial Averaging


Observations of a three-dimensional field (depending on two spatial
and one temporal arguments) can be considered as multiple time
series. Spatial averages of this field for each moment of time present
a new univariate time series with a statistical structure different from
the statistical structure of the original time series. Averaging is a
linear operation; therefore, having the information about spatial-
temporal correlations, one can evaluate the temporal correlations of
the time series of averages.
Consider k + 1 time series yo(t), y^t),..., yk(t) at different points
on the earth's surface. Let

be the spatial-temporal covariance of the observations yi(t) and yj(t-}-


r) at points i and j ; T is the time interval between these two points.
For each moment of time t, we will define an area mean

where
80 2 Averaging and Simple Models

Figure 2.7:Normalized
: variances (2.30) of the nonoptimal point estimates:
when the number of points k and a are fixed (k = 30; a = 0.1) and the polynomial
degree m varies.
2-4 Spatial Averaging 81

Figure 2.8: Normalized variances of the optimal polynomial point estimates for
independent (1) and correlated (2) observations, and corresponding variances for
the nonoptimal estimates (3); m — I (left), m = 3 (right).
82 2 Averaging and Simple Models

are the weights of averaging.


Estimates Y(t) (t = 0,1,2,...) present a new process, the autoco-
variance function (K(r)) of which is determined by spatial-temporal
covariances (2.37) and by weights to,-:

The corresponding autocorrelation function PC.P(T) is

If

we have

Let us assume that the time series yi(t) is the sum of the signal Si(i)
and the white noise N i ( t ) ,

and that noise components Ni(t) are independent in space and have
identical variance ajy. Let us also suppose that the statistical de-
pendence of yi(t) is completely determined by the signal component
5,-(0, i.e.,
2.4 Spatial Averaging 83

where a^s is the variance of process 5,-(<); pij is the spatial correlation
of values Si(t) and Sj(t); pij = 1 if i = j; and p(r) is the auto-
correlation function which is identical for all processes Si(t). Such
spatial-temporal statistical dependence is called separable (Chris-
takos, 1992). It is not something very particular because it is ex-
pressed in general form in (2.44)-(2.45) without assumptions about
the types of the spatial and temporal correlations, pij and p ( r ) .
Spatial averaging of yi(t) leads to the equality

Taking the variance of both sides of (2.46) yields

Furthermore, we replace a\ as follows:

where n + I (n < k) is the number of statistically independent vari-


ables, the mean of which has the same variance, <r|, as the mean
of the given system of correlated values S-i(t). (The definition of the
system of statistically equivalent independent variables can be found,
for example, in Bayley and Hammersley, 1946, or in Polyak, 1975.)
We have

The estimation of the autocorrelation function of separate time series


yi(t) of the signal-plus-white-noise-type process actually leads to the
estimation of the function

where
84 2 Averaging and Simple Models

is the signal ratio (Parzen, 1966; see also Section 4.8). PH(T] have
lower absolute values than p(r] because a < I. The distinction
between p ( r ) and pH(r) becomes more noticeable with increasing
ajy. Of course, if there is no additional white noise, PH(T) coincides
with p(r).
The estimation of the autocorrelation function p(r) of the means
Y(t) leads to the estimation of the function (T > 0)

where

The autocorrelation function pcp(i~) of Y(i) is closer to p(r) than


PH(T)- For positive correlations, we have

because

In any case, the result reveals that what is really estimated for the
time series of the signal-plus-white-noise type is the autocorrelation
function of the signal multiplied by a (or a). The reason for the
relatively large autocorrelations of the means Y(t) is the spatial-
temporal dependence in the signal. If Si(t) were spatially indepen-
dent (pij = 0), averaging would not improve the estimate of /o(r),
because in this case n = k and a — a.
The higher the spatial statistical dependence of the values £,-(*),
the smaller is n + 1, the closer a is to 1, and the closer the estimate
PCP(T) is to p ( r } . Non-zero spatial-temporal correlations (2.45) of the
original time series are grounds for improving the estimates of the
autocorrelation function of a signal (with the aid of the time series of
means). Notice that a decrease of the value of a"^ with an increase
of k + 1 occurs faster than a decrease in the value of a^.
The principal result of this section, equation (2.53), can be ex-
2.4 Spatial Averaging 85

pressed as follows:

where n is the number of statistically independent time series, equiv-


alent to the given system of time series yi(i] in the sense that the
variances of their spatial means are equal. The ratio (n + !)/(& -f 1)
characterizes the spatial statistical dependence of the yi(t). The
closer (n + l)/(k + 1) is to zero, the higher this correlation is. If
(n + l ) / ( k +!) = !, the yi(t) are statistically independent in space,
and their averaging does not lead to an increase in the signal ratio of
the spatial means. If (n + !)/(& +!)<!, then a > a and the signal
ratio of means is larger than that of the yi(t). Thus, spatial aver-
aging of time series with the spatial-temporal statistical structure
(2.44)-(2.45) gives new time series with the same signal but with a
smaller noise component.
Figure 2.9 illustrates the dependence (2.56) of a on the parameter
(n+l)/(k+l)and the signal ratio a. The graph shows the signal ratio
associated with the spatial averaging. Even with very low a (about
0.1), the series of means enables us to reliably identify a signal against
a background noise with the appropriate values of ratio (n+l)/(A;+l),
that is, with the fairly high spatial-temporal statistical dependence
of the observations.
Let us consider another example of the spatial-temporal statisti-
cal dependence of a signal. Let

that is, spatial-temporal covariances of the signal are equal to zero.


Then from (2.42) we have
86 2 Averaging and Simple Models

Figure 2.9: Dependence of signal ratio a of the spatial means on the signal
ratio a of the original time series and the ratio (n + !)/(& + !)•

In this case, the signal ratio

of averages is smaller than the corresponding ratio of the original


time series.
These results emphasize that the non-zero spatial-temporal cor-
relations are the basis for detecting a signal by spatial averaging of
the signal-plus-white-noise-type time series.
2.5 Smoothing of Correlated Observations 87

2.5 Smoothing of Correlated Observations


The principal requirements for design and development of a digital
filter are conditioned by the geometry of the observed function as
well as by the statistical structure of the observations. Obviously,
the application of the filters considered in Chapter 1 for smoothing
the correlated observations is not theoretically correct because those
filters were obtained under the assumption of independence and equal
accuracy. For correlated observations, the considered schemes do not
supply the minimum of the point-estimator variance. If observations
are statistically dependent, the change in the correlation structure
of the time series, as a result of filtration, has a more complicated
character. In this case, the covariance matrix of smoothing value
vector Y (1.69) is FMFT, where M = {Af,-j}^_0 is the covari-
ance matrix of observations Y. But for certain types of statistical
structures of observations (and when the width of the filter is large),
there are grounds to suppose that the variances of the point esti-
mates, obtained by applying the Chapter 1 filtering schemes, differ
insufficiently from the minimal possible values [as it was for the esti-
mates of the mean (see Section 2.3)]. This fact justified the practical
application of the digital filters considered in Chapter i in situations
when observations are not independent or do not have equal accu-
racy. The losses of accuracy as a result of the application of the
considered filters for smoothing correlated observations will depend
on the type of covariance matrix M. The simplicity of the filters
developed in Chapter 1 and the lack of information about matrix M
do not ordinarily allow an option for the experimenter. But the rea-
soning about the estimator accuracy may not make any sense when
the correlated data are smoothed by the digital filter derived for the
independent observations.
To construct a minimal variance filter, it is necessary to know
the covariances of the observations. For example, if the covariances
of the observations are kr = o- 2 e~ a l T l, then instead of averaging with
equal weights, as was done in the case of regressive filters with m — 1,
(2.20) yields

where 7 = 6 "is the correlation coefficient. To construct the regres-


sive-type filter when the statistical structure of the data is known, one
88 2 Averaging and Simple Models

must solve the system of corresponding normal equations obtained


by using the covariance matrix of the observations.
The statistical structure of the point estimates, obtained by the
moving average procedure, is determined by the covariances of the
original data. The formal approach to developing a procedure for
evaluation of the statistical structure of the point estimates for the
observations with stationary correlations follows.
Consider observations

of the stationary random processes Y ( t ) with covariance function


M(r).
The covariances RT of the point estimates

(where aj are the weights of the numerical filter, aj = «-j) are

The Yt is a stationary random process with covariance function R(T),


which is determined for points r = 0 , 1 , . . . , N — 2r as

Correlation function KT of the estimates (2.62) is

When smoothing is simple averaging

we have
2.5 Smoothing of Correlated Observations 89

and

The expression

determines the variance of the point estimates.


If estimates (2.66) are computed not for all the points but only
those with step equal to 2r, then the autocorrelation function of the
new time series (consisting of the consecutive means of the observa-
tions on the nonoverlapping subintervals with 2r + 1 terms) is

For example, the autocorrelation function of the monthly (or annual)


mean time series (as linear transformation) of the meteorological ob-
servations for the individual stations is determined by the autocor-
relation function of daily time series. By assuming that the latter
are stationary, we can apply the above methodology to calculate the
correlation function of the series of the means (annual or monthly).
Let us assume that MT is the autocorrelation function of the time
series of daily temperature anomalies measured by a point gauge. Av-
eraging these anomalies over the consecutive nonoverlapping subin-
tervals with, for example, 2r +1 = 31 points on each, one obtains the
90 2 Averaging and Simple Models

monthly mean time series (the unit of v is one month). If 2r+l = 365,
the result will be the time series of annual means (the unit of v is one
year). It is known (Lepekhina and Fedorchenko, 1972) that the ap-
proximate autocorrelation function of daily temperature anomalies
(for middle latitude stations' data) is MT = pr\ where p ~ 0.8 for
r = 1 day. By substituting this value in (2.70) for MT and assuming
that 2r + 1 = 31 (or 2r + 1 = 365), we obtain the autocorrelation
function of the monthly (or annual) mean time series of temperature
anomalies.
To perform a general study of the effect of the correlation of the
daily data on the correlation structure of the monthly and annual
mean time series, a broader analysis can be carried out by varying
p from, for example, 0.35 to 0.95 with the temporal step of 0.05.
Substituting Mr = pT, [p = p ( l ) = 0.95] for Mr in (2.70), we obtain
(for v = 1,2, and 3) the first three autocorrelations of the monthly
means: 0.40, 0.08, and 0.02; as well as the first three autocorrela-
tions of the annual means: 0.03, 0.00, 0.00. In these examples, the
first autocorrelation of the time series of the means significantly ex-
ceeds any other autocorrelation. If p <0.95, this result becomes even
more evident in the sense that the values of all other autocorrelations
(for lags equal two and greater) have three or more zeroes after the
decimal point. Therefore, it makes sense to consider only the depen-
dence of the first autocorrelation of the time series of the means on
the value of /?(!) of the original observations. This dependence is
presented in Figure 2.10.
Assuming, for example, that p ( l ) = 0.8 for daily data yields the
value of p(l) ~ 0.07 for the monthly data and p ( l ) ~ 0.006 for the
annual data. Therefore, statistical dependence of the monthly mean
temperature time series, conditioned by the synoptic fluctuations,
is significantly smaller than the autocorrelation actually observed.
(Such autocorrelation of the historical monthly mean temperature
records is equal to approximately 0.1 to 0.3 (see Section 6.4). Analo-
gously, the first autocorrelation (estimated in this way) for the annual
means is 15 to 20 times smaller than the autocorrelations actually
observed for the historical records.
The possible cause of the discrepancies obtained is that auto-
correlations in the historical records were conditioned by not only a
synoptic, but also a low-frequency fluctuation. If these fluctuations
are not the result of measurement errors, then it is a natural tem-
perature trend, which can be interpreted as a confirmation of the
nonstationary character of climate change.
2.6 Filters of Finite Differences 91

Figure 2.10: Dependence of the first autocorrelations of monthly (1) and an-
nual (2) mean time series on the autocorrelation p(l) of the time series of daily
fluctuations.

2.6 Filters of Finite Differences


The filters considered above enable us to smooth data or to remove
high-frequency components of the fluctuations. But in many prob-
lems one needs to analyze the high-frequency components and filter
out the long period variations (the low-frequency part or the fluctu-
ations). Such removal can be done with the aid of a wide range of
linear filters, the simplest class of which, the finite differences filters,
is considered in this section. Such filters are used, for example, in
the theory of random processes with stationary increments and in its
application to fitting the nonstationary stochastic models (Box and
Jenkins, 1976) to the time series with trend.
The procedure for computing finite differences, as for any linear
transformation, changes the statistical and harmonic structure of
92 2 Averaging and Simple Models

data. For different types of the correlation structure of time series,


such changes will be different. Here we will study such transforma-
tions for the simplest types of autocorrelations of observations.
The order v finite differences Vvyt of the time series ?/o, ? / i , . . . , yk
are determined in the following way:

Let us compare the harmonic structure of the time series yt and


V"j/(. If etlut is any of the harmonics of the Fourier transform of
the time series y t , then the frequency characteristic / 2 (w) of the
transformation of (2.73) is

Figure 2.11 illustrates this dependence. It shows that difference fil-


ters suppress low-frequency power and that, the greater the finite
differences order v, the wider the low-frequency interval of such sup-
pression.
The correlation structure of the time series of finite differences
Vvyt depends on the statistical and harmonic structure of the original
time series yt.
If yt is a stationary time series with an autocovariance function
M(r), then the time series of the first differences Vt/f is also station-
ary with the autocovariance function,

From this, we get the expression for the variance

and for the autocorrelation function


2.6 Filters of Finite Differences 93

Figure 2.11: Frequency characteristics of the different order v finite differences.

of the time series of VT/J. By using analogous reasoning for the dif-
ferences Vvyt of order v, we get the recursion relations

These formulas make it possible to study the correlation structure of


the finite differences time series.

Example 1
Let yt be the sequence of independent random variables with
94 2 Averaging and Simple Models

Table 2.3: Variances and autocorrelation functions KV(T) of the


white noise process finite differences.

KV(T) for T

Time series -I 1 2 3 4

Vt 1 0 0 0 0
Vj/ ( 2 -1/2 0 0 0
V2yt 6 -2/3 1/6 0 0
V32/( 20 -3/4 3/10 -1/20 0
VV 70 -4/5 2/5 -4/35 1/70

Using (2.78)-(2.80) in sequence, we find the variances d^ as well as


the autocorrelation functions Kv(r) of the time series of finite differ-
ences (for instance, up to the fourth order). Functions KV(T} are
given in Figure 2.12 and in Table 2.3. The results show that applica-
tion of the finite difference filters to the white-noise time series forms
a random process with a negative autocorrelation for the first lag
as the autocorrelations approach zero, subsequently changing their
sign.
The variance of the finite difference time series is increased (with
increasing ?;) in accordance with the formula

which can be obtained by taking variances of the right and left parts
of (2.73).

Example 2
Let yt be the normalized Markov random process with autocor-
relation function
2.6 Filters of Finite Differences 95

Figure 2.12: Autocorrelation functions of the different order v finite differences


of the white noise.

In the case of the first differences, we get [from (2.75) to (2.77)]

If v = 2, we have
96 2 Averaging and Simple Models

Table 2.4: Variances and autocorrelation functions KI(T) of the first


finite differences of the Markov process.

KI(T) foir

p «l 1 2 3 4

0.25 1.5 -0.37 -0.01 0.00 0.00


0.50 1.0 -0.25 -0.12 -0.06 -0.03
0.75 0.50 -0.12 -0.09 -0.07 -0.05

Given different values of p (p = 0.75,0.50,0.25) and using formulas


(2.83) and (2.84), one can calculate the autocorrelation functions of
the first and second differences of the Markov process. The residts
of the calculations are presented in Tables 2.4 and 2.5 and in Figure
2.13.
Figure 2.13 illustrates the dependence of the statistical structure of
the first-order finite differences of the Markov process on the p value.

Let us consider a more general case for a time series, which is the
sum of the deterministic and random components:

where P(t) is the polynomial of degree ?;, and where 6t is the sta-
tionary random process. The finite differences are presented by the
formula

Assuming that E(#t) = 0, we have

where /3V is the coefficient of /" of the polynomial P(t).


Each term of the finite difference time series (2.86) consists of
two summands: the constant /?„ (mean value) and the random Vv6t
component. In other words, the procedure for computing the finite
differences filters out the polynomial trend.
2.6 Filters of Finite Differences 97

Table 2.5: Variances and autocorrelation functions KI(T} of the sec-


ond finite differences of the Markov process.

KI(T) for T

p A 1 2 3 4

0.25 4.125 0.31 0.08 0.02 0.01


0.50 2.500 0.05 0.02 0.01 0.00
0.75 1.125 0.005 0.003 0.002 0.001

Stochastic modeling of the finite differences of the observations


(as well as its spectral and correlation analysis) brings into focus
some questions. What properties of the results are imposed by the
procedure of finding the finite differences, and what properties cor
respond to the nature of the original data? As a rule, it is not known
beforehand whether or not the time series of the finite differences is
the sample of a stationary process. The formal procedure of filtering
out a trend by finding the finite differences of the adjacent terms
could be nonoptimal for the consequent fitting of a stochastic model.
It should be noted that in some problems the estimate of the
mean of the finite difference time series Vvyt could be important.
For example, accepting

as the estimate of a mean value can lead to very large errors. For
instance, when v — 1

In other words, the estimate of the mean of the first differences de-
pends only on the first and last terms, and ignores all other observa-
tions.
98 2 Averaging and Simple Models

Figure 2.13: Autocorrelation functions (1-3) of the time series and the corre-
sponding autocorrelation functions Ifi(r) (!' — 3 ' ) of their first finite differences.
Curve... 1 2 3
M(r)... 0.75r 0.5T 0.25T

Generally, the mean of Vvyt determines the asymptotic behavior


of the nonstationary model forecast (see Section 4.9); consequently,
the problem of choosing its appropriate estimate cannot be under-
estimated. If v = 1, the slope of the straight line determines the
asymptote to which the forecast approaches when there is an increase
in the lead time or when there is a decrease in the autocorrelations.
The optimal scheme of estimating /?„ by the least squares method
depends on the unknown autocorrelation structure of random values
6t. Therefore, taking /?„ as an estimate of the mean of the cor-
responding finite differences does not reduce the difficulties. As a
preliminary estimate of such mean, it is possible to take the estimate
of /?„, when the assumption about independence of 6t is true. The
optimal estimation of the mean can be accomplished by the least
squares method along with other parameters of the model.
2.7 Regression and Instrumental Variable 99

2.7 Regression and Instrumental Variable


The linear regression and the instrumental variable methods (Kendall
and Stuart, 1963) are the next step (after averaging and the linear
trend) toward more complicated statistical modeling of data. In
this section we will analyze the normalized standard deviations of
the point estimates of two simple examples of these methods; this
analysis is preliminary to the autoregressive-moving average models
studied in Chapter 4. Autoregressive and moving average schemes
(see Chapter 4) are the results of application of linear regression and
instrumental variable methods to the analysis of time series.
Our purpose here is to numerically interpret and compare the
corresponding accuracy characteristics of the point estimates of these
methods.

2.7.1 Linear Regression


Let us begin with linear regression, considering three random vari-
ables x , y , and z with the unit variances and correlation matrix

The linear regression equation y on x and z is

where a and 6 are parameters that must be estimated and e is the


random error, independent of x and z. The least squares method
provides the estimators for a and 6:

The error variance el is


100 2 Averaging and Simple Models

The domain of the permissible correlations of the linear regression


model is determined by the equality

By fixing the value of one of these three correlations pxy, pxz, or


pyz (for example pxz), it is possible to compute the field of standard
deviations se by (2.94) and to outline the domains (2.95) on the plane
of two other correlations. If pxz is fixed (for example, pxz = 0.7),
these domains have an elliptical form, which is presented in Figure
2.14.
Isolines of ee in this figure characterize the accuracy of the lin-
ear regression model in different points of the domain of permissible
correlations. More general results, which demonstrate the depen-
dence of ee on all three correlations, are presented by the sequence
of illustrations in Figure 2.15a. As in the case of the polynomial
approximations, the standard deviations ee can be considered for
evaluating the linear regression accuracy when one has numerical
values of correlations pxy, pxz, and pyz.
If absolute values of the correlations are small (for example, less
than 0.3), Figure 2.15a shows that the linear regression model gives
estimates of y that are close to the expected value (because the value
of £e is close to one). Setting pxy = pyz — p in (2.94), we have

Isolines of this standard deviation ee (see Figure 2.16) illustrate that


it is possible to obtain estimates with ee < 0.8, when p is sufficiently
large (\p\ > 0.4). For small p, the dependence of £K on the correlations
of x and z is very low.

2.7.2 Instrumental Variable Method


In the case of the instrumental variable method (see Kendall and
Stuart, 1963) for the variables ar,y, and z with the same correlation
matrix (2.90), the following model must be built:

where m is the expected value of random variable x; e and 6 are inde-


pendent random errors; and a and b are the parameters, which must
2.7 Regression and Instrumental Variable 101

Figure 2.14: Domain of the permissible correlations and isolines of the normal-
ized standard deviations ee of the two-variate linear regression model (pxz = 0.7).

be estimated. The third variable z is considered an instrumental


variable.
The estimator of parameter a is given by the formula

Assuming that 6 = /3ei, where /3 is a unknown parameter, and e\ is


independent of e random variable, one obtains 6 + e = f3e\ + e; that
is, y is presented as a sum of ax and of a linear combination of two
independent random variables:

(Notice that this scheme, applied to the time series modeling,


produces the first-order autoregressive-moving average model
ARMA( 1,1), see Chapter 4).
102 2 Averaging and Simple Models
2.7 Regression and Instrumental Variable 103

Figure 2.15: Domains of the permissible correlations and isolines of the


normalized standard deviations ee of the two-variate linear regression (a) and
of the method of instrumental variable (b) for different values of correlations
Pxy, Pxz, Pyz-
104 2 Averaging and Simple Models

Figure 2.16: Domains of the permissible autocorrelations and isolines of the


normalized standard deviation £« of the two-variate linear regression (left) and of
the instrumental variable method (right) if pxy = pyz — p.

According to the instrumental variable theory (Kendall and Stu-


art, 1963), the following formulas for the variances of random vari-
ables e and S are correct:

The domain of the permissible correlations of the model (2.97) is


determined by the inequalities

The right set of inequalities in (2.101) occurs because, in the inverse


case, instead of needing to get estimates (2.97), one must get the
mean values of x and y.
So we have

It follows that the equivalent system of inequalities is


2.7 Regression and Instrumental Variable 105

These inequalities determine the domain of permissible correlations


pxy, pxz, and pyz for the model (2.97). By fixing the value of any of
these correlations, it is possible to outline the boundary of this do-
main on the plane of two other correlations. For example, assuming
that pxz = 0.7, inequalities (2.103) give

This part of the domain of permissible correlations [together with


the isolines of the standard deviations ee computed by (2.100)] is
illustrated in Figure 2.17. Isolines characterize the accuracy of the
point estimates y in the domain of the permissible correlations.
A comparison between Figures 2.14 and 2.17 reveals that the
domain of the permissible correlations for the linear regression is sig-
nificantly larger than the corresponding domain of the model (2.97).
Actually, the model (2.97) is determined in the domain, the total
size of which does not exceed the unit quadrant of the first quarter
of the coordinate system.
In spite of the clear distinction in the form of the domains of
the permissible correlations for the two considered models and the
character of the variation of corresponding isolines of ee, it is possi-
ble to make the following remarks: For the coinciding parts of the
domains and small values of the correlations (if pxy and pyz are less
than 0.4), values ee of both models are approximately the same, that
is, in this part of the domain, the accuracies of these approximations
are virtually indistinguishable. The instrumental variable model is
slightly more accurate on the periphery of its domain of permissible
correlations (when pxy and pyz are greater than 0.7).
Figure 2.15b presents the variations of the sizes of the domain of
the permissible correlations for the model (2.97) and isolines £e as a
function of all three correlations pxy, pyz, and pxz.
A comparison of the accuracies of these two models, presented in
Figure 2.15, confirms the inferences obtained by the analysis of Fig-
ures 2.14 and 2.17. Moreover, if the values of correlations pxy, pxz,
and pyz are known, one can use Figure 2.15 to establish the possibility
of building corresponding models and to determine their normalized
standard deviations ee.
If point (pxy, Pxz, Pyz) is outside the domain of the postulated
model, then the corresponding formally computed point estimates
will have a negative variance, as happened in an example in Kendall
and Stuart (1963).
106 2 Averaging and Simple Models

Figure 2.17: Domain of the permissible correlations and isolines of the nor-
malized standard deviations ee of the instrumental variable method (pxz = 0.7).

Let us assume that pxy = pyz = p in (2.94), (2.95), (2.100),


and (2.103). Then, the domains of the permissible correlations of
both models are the two-dimensional domains presented in Figure
2.16. This figure shows especially clearly the distinction between the
accuracies of the methods considered in the sizes of the domain of
the permissible correlations as well as in the shapes of isolines se.
As we conclude this comparative analysis, let us notice that lin-
ear regression is used more often in applications than the scheme of
the instrumental variable because its domain of permissible correla-
tions is greater. The slight advantage in accuracy (on the periphery
of the domain of permissible correlations) of the instrumental vari-
able method proves important for specific problems. The solutions
of these problems must be accompanied by testing to be sure that
the correlations are inside the domain of permissible correlations.
Figures 2.14 through 2.17 can be used for this purpose.
Let us also notice the analogy of the above consideration with the
analysis of the variances of the point estimates, which was carried
2.8 Nonlinear Processes 107

out in Chapter 1 lor the polynomial approximation, and with the


comparison (which will be given in Chapter 4) of the theoretical
accuracy of different ARM A models.
In general, theoretical analysis of normalized variances of the
point estimates can be done for any linear statistical method prior
to actual application. Such powerful methodology helps to get repre-
sentation about the entire class of corresponding models, to provide
an optimum way of obtaining the necessary estimates, or to draw a
conclusion without computations.
Moreover, it is possible to state the approximate requirements
for the accuracy of the model considered in terms of the normalized
variance values. For instance, the model can be considered reason-
able only if £e < 0.9. Such condition leads to an additional limitation
on the correlations (pxy, pxzi Pyz) even if they are inside the corre-
sponding domain of their permissible values: Their absolute values
must be greater than 0.3-0.4. If the correlations are smaller than
these values, building a linear regression equation (as well as the
instrumental variable equation) does not make sense.

2.8 Nonlinear Processes


The statistical methods based on the operations with the covariance
structure of the random variables makes it possible to evaluate a
level of corresponding linear statistical dependence. The relation-
ships of the elements within the climatic system are nonlinear. This
can be seen, for instance, from the hydrodynamic equations of the
atmospheric circulation (because the latter are nonlinear). The Co-
variance analysis of the nonlinear objects can be inaccurate (or only
a very rough approximation), not always reflecting the real character
of the dependence between variables.
A systematic presentation of the nonlinear theory of random pro-
cesses is given, for example, by Raibman (1983). The principal
concept of this theory (corresponding to the concept of the covari-
ance function) is the dispersion function, which is determined by the
second-order conditional moments. With the aid of the concepts,
defined by Raibman (1983), it is possible to carry out more in-depth
investigations, as compared with the analysis of the covariance struc-
ture.
In the simplest case of the univariate process z ( t ) , the core of the
nonlinear theory is the definition of the autodispersion function
108 2 Averaging and Simple Models

where ~E{z(t)\z(T)} is the conditional mathematical expectation of


random process z ( t ) given Z(T).
The standardized autodispersion function of z ( t ) is

where a2(t) is the variance of process z(i).


If the values of z ( t ) and Z(T) are connected in a deterministic
way, z(i) — /[Z(T)], then the autodispersion function 0(t,r) equals
the variance of the random process z(t,J:

and the standardized autodispersion function r)(t,r) equals one. The


values of z ( t ) and z(r) are called nondispersive for some values of t
and T, if 0(<,r) = 0. By definition, the autodispersion function is
non-negative and satisfies the following inequalities:

The standardized autodispersion function is greater than or equal


to the absolute value of the corresponding autocorrelation function
p(t, T), that is,

If z ( t ) and Z(T} are connected by the linear dependence, then

The indication of the existence of the nonlinear dependence between


z ( t ) and z(r) can be based upon the strict inequality in (2.110).
In order to understand the possibility of using the standardized
autodispersion function r/(f,r) in practice as a measure of the non-
linear statistical dependence of different values of a random process,
the following fact, proved by Raibman (1983), can be important: For
the Gaussian processes, the equality

is true. It follows that


2.8 Nonlinear Processes 109

Therefore, for the Gaussian processes, the measurement and analysis


of the nonlinear statistical dependence lead to the estimation of the
nonstationary autocorrelation function p(t,r).
The sample distributions of various climatic time series have been
estimated in many papers. In most cases, the normality of these
distributions is beyond question. Theoretically such normality is also
ensured by the corresponding Central Limit Theorem because annual
and monthly means are the results of averaging many separate daily
observations.
Talking about stationarity, one must notice that climatic time se-
ries do not always satisfy this requirement (for example, it is possible
to mention the time series of the meteorological elements with a sta-
tistically significant linear trend; see Chapters 6 and 7). Therefore,
the problem of considering the nonlinearity does not seem absolutely
hopeless, especially if we take into account the constantly increasing
volume of the climatic observations. In any case, the methods of
the nonlinear statistics, it seems, offer a future direction for studies,
especially in the case of analyzing the point gauge observations.
3

Random Processes and


Fields

In this chapter, the nonparametric methods of estimating the spectra


and correlation functions of stationary processes and homogeneous
fields are considered. It is assumed that the principal concepts and
definitions of the corresponding theory are known (see Anderson,
1971; Box and Jenkins, 1976; Jenkins and Watts, 1968; Kendall and
Stuart, 1967; Loeve, 1960; Parzen, 1966; Yaglom, 1986); theref
only questions connected with the construction of numerical algo-
rithms are studied.
The basic results ranged from univariate process to multidimen-
sional field are presented in Tables 3.1 and 3.2. These formulas make
it possible to compare and trace the formal character of developing
estimation procedures when the dimensionality is increasing. The
schemes in these tables, as well as the formulas in the previous chap-
ters, can be used for software development without any rearrange-
ment. In part, this approach presents the application of the methods
of Chapters 1 and 2 in evaluating random function characteristics.
Of course, the final identification of the algorithm parameters (for
example, the spectral window widths) can be made only through
trial and error and by taking into account the character of the prob-
lem under study, that is, the physical properties of the processes and
fields observed. The last section of this chapter presents results of the
application of these methods to the analysis of some climatological
fields.

110
3.1 Stationary Process 111

3.1 Stationary Process


Here the basic results of the univariate spectral analysis are briefly
discussed in order to develop algorithms for a multidimensional case
by analogous reasoning. The complete description of the estimation
procedures of the spectral and correlation analysis for univariate sta-
tionary process can be found, for example, in Jenkins and Watts,
1968.
It is known (Yaglom, 1986) that the autocovariaiice function
M(r) and the spectral density S(u>) of stationary random process
Y(t) are interconnected by the Fourier transforms

where M(r) is an even function of r, and S(ujJ is a non-negative even


function of o>.
The variance of stationary process Y(i) is cr2 = M(0).
Together with M(r) and .S'(w), we will consider the autocorrela-
tion function

and the normalized (standardized) spectral density

which are interconnected by equations analogous to (3.1) and (3.2).


The main problem of applied spectral and autocorrelation anal-
ysis of a random stationary process Y(t') consists of estimating the
mean, the spectrum S(u), and autocovariance function M(r). For
the ergodic stationary process, such a problem can be solved with
the aid of a single sample (time series) of Yt = Y(ti) given on the
finite grid to,t\,.. .,^._i; Ai = /,-+i — i{ = constant, if the number
of observations is sufficiently large.
The estimation of the mean was discussed in detail in Chapter
2. Here, let us suppose that the mean is known and that Yi are the
anomalies, that is,

A numerical methodology for estimating spectral density consists of


the application of different filtering procedures to the smoothing of
the periodogram (see Jenkins and Watts, 1968).
.12 3 Random Processes and Fields

The periodogram of observations FQ, YI, ..., Yk-\ is determined


by the following formula:

where

A graph of a periodogram (amplitudes of the Fourier expansion of the


observations) as a function of frequency consists of separate, random-
height lines, each of which presents the part of the variance condi-
tioned by the fluctuation with the corresponding frequency.
Formula (3.6) presents the estimator of autocovariance function
M(r). Furthermore, we assume that M(r) converges to zero if r
goes to infinity. We also assume that k is sufficiently large, so that
for

the values of MT are close to zero. Then, from (3.4), we can obtain

It has been shown (Jenkins and Watts, 1968) that for sufficiently
large &, the covariance C^ of statistics IIp and IIq is

For white noise process, this equality is precise.


3.1 Stationary Process 113

If 2r + 1 is the width of the filter and values Hp (p — 1,2,...) are


approximately independent and have approximately equal accuracy
on each - ^ r ^ " ^ length frequency subinterval, then the smoothing
procedure, described in Chapter 1, can be applied for estimating the
spectral density. Smoothing is performed with the aid of the spectral
(filter) window {aj}j__ r by the formula

Estimates Hp are statistically dependent. The study of their covari-


ance and correlation functions can be carried out by the approach
developed in Chapter 1.
It is easy to show that

where

and wr is the correlation window corresponding to the chosen spec-


tral window otj. Any specific weights of a spectral window OLJ in
(3.10) have a corresponding correlation window.
To build an optimal digital filter, one must know the correlation
structure of the values of periodograrn. Such a structure is known
only approximately (3.9), and in practice we apply different empir-
ical approaches to find the corresponding type of window and its
parameters.
Therefore, the core of spectral estimation methodology is har-
monic analysis of the time series with consecutive smoothing of the
amplitudes of the Fourier expansion. The representation of the sam-
ple variance as a sum of the amplitudes of the harmonics with differ-
ent frequencies [Parseval equation, see (3.8) for r=0] makes it pos-
sible to compare their power. Because the number of observations
(k) is finite and the step is equal to Ai, we can analyze fluctuations
with periods ranging from 2Ai (where At is the time step) to kAt.
A periodograrn of normalized time series gives the variance propor-
tion for each frequency that forms a random variation of the time
114 3 Random Processes and Fields

series. Every point of such a representation characterizes the power


of the corresponding fluctuations. For different frequencies the value
of power is a random value. Periodogram is a random sample that
has the same relationship to the spectral density as a separate ob-
servation of a random, variable to its expected value. The principal
theoretical characteristic (spectral density) presents the mean spread
(over all possible samples) of the variance over the frequency axis.
It is interesting to notice that, long before the stationary random
processes theory was created, the need to smooth (or to average over
different subinlorvals of the frequency axis) periodogram values was
discovered by Albert Einstein (Einstein, 1986). Indeed, Einstein an-
ticipated the development of the modern methodology of the spectral
estimations.
The principal aspects of the smoothing (by a digital filter) were
discussed in Chapter 1, where the necessity of the trial and error
method for obtaining the approximately unbiased estimates with
small variance was shown. Smoothing in the frequency space has
some features that contrast with the application of the filters for in-
dependent observations in the time domain. First, different points of
a periodogram can have different variances, which can cause signifi-
cant errors (bias) in the spectral estimates. For example, as will be
shown, for the annual mean surface air temperature time series, the
power of fluctuations in the low-frequency part of the periodogram
(with the periods in several dozens of years) is larger than the cor-
responding power in the higher frequency part. The spectrum of the
air temperature time series has a low-frequency maximum (in the
vicinity of zero), and, according to (3.9), corresponding values of the
periodogram have greater variances than in other parts of the spec-
trum. In the process of smoothing, when the left part of the spectral
window is applied to points with large variances and the right part
of this window covers the points with small variances, it is inevitable
that the point, estimates will, have bias. This fact can be illustrated
by a graph with superimposed curves of the periodogram and cor-
responding spectral estimates (analogous to those given in Chapter
6). A general approach for improving the estimates is to increase the
number of observations and, as a result, the number of points of the
periodogram, which allows us to vary the width of the filter.
The selection of the spectral window is affected by many factors,
one of which is the necessity of obtaining non-negative spectral es-
timates. The values of the periodogram are always greater than or
equal to zero, so negative spectral estimates can be obtained only by
using an inappropriate filter, which has some negative weights. Many
3,1 Stationary Process 115

different spectral windows for univariate time series that guarantee


the positiveness of the spectral estimates have been developed (Jenk-
ins and Watts, 1968). As becomes clear from the filtering schemes
considered in Chapter 1, only regression filters with m = 1 have posi-
tive weights. Therefore, the superposition of these filters [as was done
by Tukey, (1.105)] is one of the approaches in constructing smoothing
procedures for spectrum estimation.
Along with the harmonic and statistical analysis of the spectral
estimation procedure, the purely computational aspects of the chosen
numerical scheme are also important. There were two principal ways
to develop a computational procedure: the Blackman-Tukey method
and the Cooley-Tukey method.
The Blackman-Tukey method. An algorithm suggested by Black-
man and Tukey (1958) is the estimation using (3.6) of the first v
values of MT; (T -= 0,1,.. . ,f; v « n) of the covariance function.
After this, the computation of Hp is performed by (3.11), where the
appropriate correlation window wr (WT -— 0 if T > v) is applied.
The Cooley-Tukey method. The sequence of computation by
this scheme supposes the following: immediately finding the peri-
odogram by the Fourier transform (3.5) of the time series; computing
and smoothing (by using a numerical filter) the periodogram (3.4);
and estimating the covariance function by another Fourier transform
(3.8). The widespread application of this methodology became pos-
sible after the publication of the paper by Cooley and Tukey (1965),
where an algorithm of the fast Fourier transform (FFT) was ob-
tained.
For large &, the second scheme is thousands of times more eco-
nomical than the first one; for this reason, it has been widely used in
the last decades. Additionally, with the aid of FFT, it is possible to
analyze two time series simultaneously and to significantly increase
the effectiveness of the second method.
However, the Biackman-Tukey approach (with some of the well-
developed spectral windows) is not only an effective numerical scheme,
it is also a rational estimation methodology, which (for some prob-
lems) can be better than the other approaches. Therefore, the com-
prehensive coexistence of both schemes proves inevitable when one
creates a package of statistical computer programs.
The three Fourier transformation method. Our practice showed
that the combination of the above methodologies (Fourier transform
of the time series —> periodogram —> Fourier transform of the peri-
odogram —> correlation function —>• truncating and weighting of th
correlation function —> its Fourier transform —*• spectrum) with thr
116 3 Random Processes and Fields

Fourier transformations utilizes the advantages of both approaches.


The efficiency of this scheme has become especially apparent in the
case of multidimensional spectral analysis when direct smoothing of
the points close to the boundary of the multidimensional frequency
domain presents real algorithmic and programming difficulties.
These approaches do not exhaust all of the smoothing schemes
which may be applied for estimating the spectrum of the stationary
random process. Selecting a spectral window or developing a new
one is dictated by many factors (the statistical properties of data, a
physical feature of the problem, the sample size, and so on).
Along with the covariance function and spectral density, it is
interesting to consider the cumulative spectral function. Letting r =
0 in (3.2), we have the equality

which summarizes the spectral values of all the frequencies.


When the spectral density is analyzed, it is often necessary to
estimate a part of the variance corresponding to a finite frequency
subinterval [0,u;]. The function

which gives that part of a2, is called the cumulative spectral function.
It is a convenient way to consider the normalized (standardized)
cumulative spectral function

I(fjj) is a positive, nondecreasing function, and

The estimator of the normalized cumulative function (3.16) is


3.2 Cross-Statistical Analysis 117

On the graph of the sample cumulative spectral function, the jumps


of discontinuity, equal to the periodogram value, can be seen.
For testing the hypothesis that the time series is the sample of
a white noise process, the well-known Kolmogorov-Smirnov criterion
(van der Waerden, 1960) of constructing the confidence interval for
the sample distribution function can be applied. Actually, the stan-
dardized cumulative spectral density of the discrete white noise is
equal to a constant

and the corresponding /(w) is

This function presents a linear segment from point (0,0) to (?r/A£, 1).
By clioosing the appropriate scale along the u> axis, the straight line
segment bisecting the coordinate angle can be obtained. If estimates
(3.19), plotted in such a coordinate system, are far from the theoret-
ical line of the white noise, then it is plausible that the observations
are not the sample of the white noise. Therefore, the white noise
model is a useful standard with which the real processes of nature
and of technique can be compared.
According to the Kolmogorov-Smirnov criterion, if k > 200, the
confidence intervals for the standardized spectral density of the white
noise are determined by the strips ±S/\/k/2 — 1, where values of 6,
equal to 1.63, 1.36, and 1.02, correspond to significance levels of 0.01,
0.05, 0.25. If k < 200, the confidence intervals can be determined
with the aid of a table, which can be found, for example, in Jenkins
and Watts (1968).

3.2 Cross-Statistical Analysis


In the case of the two time series, the mutual statistical character-
istics must be estimated together with the spectral densities and
autocovariance functions of each of the processes.
Let us assume that observations Yj = Y(ti) and Z, — Z(ti) of sta-
tionary processes Y(t) and Z(t) are given in grid t0, t-i,..., tk--\; A£ —
ti+\ - /,- = const.
118 3 Random Processes and Fields

Cross-covariance function MY (T) and cross-spectral density


YZ
S (u)) of these processes are interconnected by the Fourier trans-
forms

and

The cross-covariance function can be presented as the sum of its even


and odd parts

where

Then (3.22) gives

where

is called the cospectrum and

is called the quadrature spectrum of the processes Y(t} and Z ( t ) .


With the aid of A(u;) and «3>(u;), the phase spectrum is determined
as

9(w) shows the lag of the harmonic with frequency u; of the process
Z ( l ) with respect to the corresponding harmonic of Y(t).
The next characteristic of the bivariate random process is called
the coherency function; it is determined by the formula
3.2 Cross-Statistical Analysis 119

where SY(u) and Sz(u>) are the spectral densities of Y(t) and Z(i).
F(u;) is a measure (the squared correlation coefficient) of the statis
tical dependence of the amplitudes of the harmonics with frequency
<jj of the Fourier expansions of processes Y(t) and Z(t).
The cross-spectral density estimation is performed by applying
an appropriate digital filter (spectral window) to the smoothing of
the cross-periodogram, which is determined as follows:

where u;p = 2irp/k (p = 0,1,..., k — 1),


Ap is a complex conjugate of Ap,

The estimates of the cospectrum A(u>) and quadrature spectrum


<&(u;) are obtained by smoothing the sample cospectrum

(where ' and " designate real and imaginary parts respectively of cor-
responding complex numbers) and the sample quadrature spectrum
120 3 Random Processes and Fields

Figure 3.1: Location of the discrete points of the even periodic function, (a) k
is even, (6) k is odd.

If k is large, the formulas for the variances of (3.35) and (3.36) are
(see Jenkins Watts, 1968)

Quantities G'p and G'p are considered preliminary estimates, which


must be smoothed by applying the appropriate digital filters. It is
possible to show that smoothing values G'p and G'p, considered as the
estimators of the cospectrum and quadrature spectrum, are asymp-
3.2 Cross-Statistical Analysis 121

Figure 3.2: Location of the discrete points of the even periodic function, (a) k
is even, (6) k is odd.

totically unbiased.
When smoothing the values of G'p and G f p, which are close to
the ends of the interval, their even or odd character must be taken
into account and corresponding periodical extensions must be per-
formed. The illustrations in Figures 3.1 and 3.2 can be useful for
such extension in developing computer routines.
The estimators of phase spectrum 6(w) and coherency function
r(u;) are

and
122 3 Random Processes and Fields

where /fj , HJ,, G'p, and G"p are the estimates of spectral densities,
cospectrum, and quadrature spectrum |of processes Y(t) and Z(t)],
which were obtained by smoothing the corresponding periodograms.

3.3 Nonstationary Processes


The principal statistical characteristics of the nonstationary process
Y(t) are the mean C(t) and the covariance function M(i, #), which
are determined by formulas

and

M(t,e) = E\[Y(t) - C(t)][Y(6) - C(0)]}, t,0e (-00,00). (3.41)

M(t, 9) is the function of two variables, and for its estimation, an


ensemble of samples is required.
If / = 0,

is the variance of the process. In general, a nonstationary process


has no spectral density. Some classes of a nonstationary random
processes that do have spectral density arc considered in this section.
Different random processes can have nonstationary moments of
different order. For example, the time series of the monthly means of
meteorological elements have a mean and a variance with an obvious
nonstationary component, a seasonal cycle. Sometimes, the study of
such nonstationary processes can be performed by simple transfor-
mation (removing the seasonal cycle) of the samples, which makes
it possible to obtain the time series of normalized (standardized)
anomalies, which can be considered as approximately stationary.
Before we consider the numerical scheme for the M(i,r) esti-
mation, it is necessary to mention some of the approaches that are
used for analyzing some types of nonstationaritics having only one
sufficiently Song time series.
Let us denote the observations
3.3 Nonstationary Processes 123

given in the temporal moments

as a random vector

Let the covariance matrix of Y be

where M,j = M(*,-,0j)

1. Estimation of the mean that is variable in time. The least


squares numerical schemes described in Chapter 1 can be used to
approximate the itonstationary mean if it can be presented as a linear
function

where f)j are unknown parameters that have to be estimated; Xj(£),


(j = 0,1,...,TO)is the chosen system of functions. For example, the
simplest variation in time is the linear trend,

Let us consider a general case (3.47). If parameters /3j are estimated


by the least squares method, we must minimize (with respect to j3)
the quadratic form

where

In this case the estimates f3 are not optimal because matrix (3.46) is
unknown and is not used in the process of estimation.
Sometimes, further spectral and correlation analyses are applied
to the residuals,
124 3 Random Processes and Fields

where C(f) is the estimated mean.


Whatever method has been used for estimating a nonstation-
ary mean, the second moments of the original observations and the
residuals are different. The scale of this difference depends on the
statistical structure of the process as well as on the methodology that
is used to estimate the mean. This statement will be illustrated by
the formula for covariance matrix of the residuals (3.53) of the least
squares method. These residuals are the elements of vector

the covariance matrix M' of which is

It is clear that matrices M and M' arc different. The conditions,


when matrices M and M' can be asymptotically (k —> oo) close, are
given in Anderson (1971).
If k is riot large, the correlation structures of the processes, pre-
sented by residuals (Y — X/3) and by F(t), are significantly different
even when Y(t) is the white noise or the Markov process. In spite of
the fact that the derivation of the matrix expression (3.55) is simple
enough, the practical aspects of the evaluation of the closeness of
matrices M and M' are difficult.
Therefore, estimating the second moments of residuals Y — X/3
by any method, we obtain the estimates of the elements of matrix
M' (not the M). It is clear that scheme (3.47) should be used very
carefully and only in situations when the primary interest is the
estimation of the mean, not in the accurate analysis of the second
moments.
If parameters /3j are estimated by some other method than the
least squares method, the correlation, structure of vector (3.54) differs
from that of (3.55). For the linear estimation scheme, the deduction
of the expression for the covariance matrix of the residuals does not
present any formal difficulties.
2. Moving average harmonic analysis. The process with variable-
in-time autocorrelation function and spectral density has a more
complicated nonstationary structure. Sometimes, such variation in
time is slow, and on certain subintervals (of the observational inter-
val) the process can be considered as approximately stationary. It
3.3 Nonstationary Processes 125

is possible to get a qualitative picture of the variable-in-time statis-


tical structure of the process if the sample size is sufficiently large.
In this case one must carry out the standard spectral analysis of
data on separate (possibly overlapping) subintervals and compare
the results. Estimates of the correlation function and spectral den-
sity, which correspond to separate subintervals, are referred to as
the local estimates. Their consecutive analysis can show such in-
teresting features as the transfer of the power from one frequency
band to another; the process of the formation of the spectral max-
ima; the temporal localization of separate observed quasi-periodic
fluctuations; and the variation of the white noise level. In this case,
the estimates of the spectra can be suitably presented in the form of
a two-dimensional (frequency and subinterval number) function. In
a sense, this algorithm is the generalization of the moving average
procedure, because on each subinterval, the estimates of the mean,
as well as some other statistical characteristics, are obtained.
3. Harmonizable random processes. The most general class of
nonstationary processes that have spectral representations is referred
to as the harmonizable processes. Their practical spectral analysis
is possible if the corresponding covariance function has the Fourier
expansion

where

The function S(fl,u>) is called the double frequency spectral den-


sity of the nonstationary random processes Y(t). The estimation of
M(t,0) and 5(ft,u>) can be performed with the aid of the ensemble
of samples.
Consider n time series of the harmonizable random processes with
k observations in each. These observations are given for the same
moments of time, and they can be presented in the form of matrix

Let us suppose that the random process has a zero mean.


126 3 Random Processes and Fields

The estimator R^ of covariance function M(t, $) is

The estimator Hpq = H(ilp,u>q) of function 5(J2 p ,w g ) is denned in an


analogy with a definition of the double frequency spectrum (3.57):

Substituting (3.59) for Rtg in (3.60), we have

The last three formulas demonstrate the possibility of finding the


double frequency spectrum estimates by one of the following ways:
the Fourier transform of weighed estimates Rto, or the Fourier trans-
form of the separate samples with consequent computation by for-
mula (3.61).
To find the estimate of the double frequency spectrum, it is nec-
essary to smooth the field of values IIpq by an appropriate two-
dimensional digital filter. An estimator of the covariance function
in this case is

We considered a formal numerical approach for estimating the dou-


ble frequency spectrum of the harmonizable process. This approach
does not contain any new numerical ideas. But the physical interpre-
tation of the estimates obtained by this method is not simple, and
this might be why relatively few papers discuss such results. Notice,
furthermore, that the generalization of the considered numerical ap-
proach on the cross-analysis of the harmonizable processes has no
difficulty in principal. However, in an effort to avoid encumbering
the text and repeating very similar ideas, we do not present such a
generalization. A review of some of the algorithms of the numerical
analysis of nonstationary processes is provided by Anderson (1971).
3-4 Nonergodic Stationary Process 127

3.4 Nonergodic Stationary Process


The methodology of spectral and correlation analysis for the station-
ary random process (Section 3.1) can be applied when the condition
of ergodicity is satisfied. This condition theoretically makes it pos-
sible to find the estimates with the aid of the only sample. If this
condition is not satisfied, then the estimation of the spectral density
and correlation function requires an ensemble of time series. We will
discuss the estimation scheme, which can be applied in this case. Let

be n time series (with k observations in each, given in the same time


grid) of the nonergodic stationary random process with a zero mean.
The sample covariance matrix,

of these observations is formed by the estimates

The process is stationary, and its covariance function depends on only


one argument (lag). Therefore, the elements, situated on any matrix
(3.64) diagonal that is parallel to the main diagonal, present the
estimates of the same correlation coefficient. (An exact equality of
these estimates has not taken place because of the sample variability.)
The estimates IIT of covariance function M(r) can be found by
averaging the estimates on the diagonals, which are parallel to the
main diagonal of matrix (3.64):

Obtaining Rt,t+T by (3.65) and substituting it in (3.66), we have

where
128 3 Random Processes and Fields

Formula (3.67) is the estimator of the covariance function for an


ergodic stationary processes. The estimator (3.67) is the mean of
the estimates of the covariance functions, obtained for the separate
time series of the ensemble by applying scheme (3.6), which was
developed for the ergodic processes. The estimation of the spectrum
by (3.11) yields

where H-* is the estimate of the spectral density of sample j by the


scheme (3.10). Therefore, the estimate Hp (3.69) is the average of
the spectral estimates for the separate time series, and RT (3.67) is
the average of the corresponding sample covariance functions.
In the situation considered, all of the samples of the ensemble had
the same size. In practice, however, this is not often the case. For
example, the time series of the meteorological elements of different
stations can have different sizes. In order not to lose the informa-
tion by reducing the sizes of time series to the common overlapping
subinterval, it is possible to use the following estimation procedure.
1. Compute the periodogram for each time series.
2. Divide the frequency interval (from zero to the Nyquist fre-
quency) to a certain number of noiioverlapping equal-width
bands.
3. Average the values of each separate periodogram inside such
bands.
4. Average the obtained estimates for each frequency band over
the whole ensemble.
5. Find the estimates of a spectrum by smoothing the set of the
results using a digital filter.
6. Estimate the covariance function by the Fourier transform of
the spectrum estimates.
This estimation scheme can be easily generalized for the case of
the cross-correlation analysis of the two stationary nonergodic ran-
dom processes, each of which is presented by the ensemble of samples.
The estimates of the cross-statistical characteristics in this case are
the ensemble means of the estimates, which are found for the separate
sample pairs under assumption of their ergodicity.
3.5 Time Series with Missing Data 129

3.5 Time Series with Missing Data


One of the features of the geophysical time series is the possible
existence of random gaps. These gaps are often caused by the nature
of the phenomena being studied. For example, observations of direct
solar radiation cannot be obtained when clouds cover the sun. As an
added example, the discontinuity of the precipitation or cloud data
is part of their physical nature.
There are numerous methodological procedures in this field, some
of which can be found in Parzen's references (1984).
Consider a general approach to the problem of a discrete har-
monic analysis of observations

given on irregular grid points t0,ti,.. .,tk-i', ti+i > £,•; <;+i - ti ^
constant (so the time steps are not equal).
The estimates of the coefficients of the approximating trigono-
metrical polynomial

(where 1v -f 1 < k; fii,ft2) • ••,£}</ are an irregular system of grid


points of some frequency interval [0, !}„]) must be found.
In this case, the harmonic analysis is connected with the identifi-
cation of the proper system of points {Qp}p_0 of a frequency interval.
This situation is analogous to the identification of the algebraic poly-
nomial degree, when least squares polynomial approximation of a set
of observed data is performed.
The proper analysis of the physical contents of the problem can
be helpful in such identification. It is difficult to make general rec-
ommendations; nonetheless, the following qualitative reasoning can
assist in some practical situations.
For the reliable estimation of coefficients Ap and Bp correspond-
ing to any frequency

it is necessary that the time interval [0, tk-i] contain a sufficient


number of nonoveriapping subintervals of Tp length with at least two
observations on each subinterval.
130 3 Random Processes and Fields

If A£ m j ri is the minimal distance between observations, then in-


terval [0,Jlj/] C [0,7r/Ai m i n ]. In searching for fi^, one can use the
value

(where Atcp = (tk-i ~ fo)/&) as an approximate high possible value


of frequency ($}„) in (3.71).
The coefficients (Ap and Bp) of the expansion (3.71) can be es-
timated by the least squares method. The amplitudes obtained for
frequencies fl p correspond to the uonequal frequency steps. The
smoothing of such estimates can be done by approximation using an
appropriate function.
This approach is very cumbersome. On irregular grid points, the
trigonometrical functions have no orthogonal properties, and estima-
tion of parameters Ap and Bp requires that one construct and solve
the system of normal equations. For large t/, this problem requires a
huge number of computations.
The analysis of the equally spaced time series with a small number
of missing data can be carried out with the methods described, for
example, by Parzen (1984).
The following qualitative reasoning characterizes the possible in-
accuracies in the spectral estimates arising as a result of missing data
in the time series (given on the regular grid points, if A'i = i;+j —ti is
a constant). According to the Nyquist theorem, the minimum period
of the Fourier expansion is

and it contains two observations. Therefore, gaps affect, the estimates


of the spectrum corresponding to the periods, for which condition
(3.73) does not hold all the time. Thus, one can expect that a small
number of gaps will not significantly distort the results for the low
and middle frequencies. The closer the estimates are to the Nyquist
frequency, the lower their accuracy is. With an increasing amount
of missing data, the improperly estimated part of the spectrum is
shifted to the domain of the middle frequencies.
When the interest is only in the low-frequency part of the spec-
trum, it is possible to intentionally reduce the frequency interval by
temporally averaging the terms of the original time series. For ex-
ample, instead of observations with step Af, one can take the values
with step AA/ (where A is a positive integer); these new values arc the
averages of at least A consecutive observations of the original time
scries. In this case, the Nyquist frequency is reduced A times (from
3.6 Two-Dimensional Fields 131

7T/A/ to 7r/AA£), but the newly created time series contains fewer
missing terms than the original one. The frequency characteristic

of the above averaging does not accurately reflect the real modifica-
tion of the amplitudes because the real averaging of the original data
was performed only for the existing observations.
The main idea of the statistical formulation of the problem with
missing data (Parzen, 1984) is in introducing a random function ?/(£),
which equals zero if the observation is absent and which equals one
in all other points. Thus, the sample can be presented in the form
of the amplitude-modulated time series

where X(t) is the sample with no missing data. Under some condi-
tions (first, if the number of observations is sufficiently large), the
estimates of the covariance function and the spectral density, found
with the aid of Y(t), will be close to the corresponding characteristics
of X(t).
Notice here that an additional deterministic consideration of this
problem is possible. The effects imposed by function 'f](t) are condi-
tioned by the convolution theorem. For each observed sample, the
function rj(t) is known, and one can perform its harmonic analysis.
Therefore, the values of the Fourier expansion of ri(t] reveal the char-
acter of the distortion of the amplitudes of the original time series.
If such distortions are randomly and evenly distributed along the
frequency interval, one can choose the appropriate spectral window,
taking into account the information obtained about the frequency
characteristics of rj(t). If these distortions lead to significant misrep-
resentation of the information on the separate frequency subintervals,
obtaining reliable results can be impossible. In any case, detailed
analysis is labor-intensive.

3.6 Two-Dimensional Fields


The formal generalization of the theoretical concepts and methods
of the stationary random processes to homogeneous random fields
does riot present any difficulties. At the same time, however, the
estimation methodology (together with the natural complication of
132 3 Random Processes and Fields

the numerical algorithms) of the two-dimensional analysis contains


some peculiarities. The purely formal succession to the analogy with
the estimation theory for stationary processes can lead to loss of a
part of the information contained in the samples.

1. We will consider the algorithms for the cross-statistical analy-


sis of two-dimensional homogeneous random fields presented by two
samples. For the single sample, the required characteristics will be
obtained as a particular case of the methodology carried out. For
the two homogeneous random fields

with zero means, the cross-covariance function is

This function can be presented by Fourier integral

where

is called the cross-spectral density. If Y(0, i) is equal to X(0,i),


then (3.76) determines the covariance function, arid expansion (3.77)
determines the spectral density of the homogeneous random field
Y(9,t).

2. Consider two finite discrete samples

of random fields Y(0, i) and X(0, i).


3.6 Two-Dimensional Fields 133

The two-dimensional discrete Fourier expansions of these samples


can be written as

where

and

where

Let the step along the 9 axis be equal to A#, and let the step along
the t axis be Ai. Then in the frequency domain, the step along the
O axis is

and the step along the u> axis is

The expression

determines the cross-periodogram of Y(d, t) and X(8, t).


With the aid of the following definitions for the sample cross-
covariance function,
134 3 Random Processes and Fields

(3.86) can be presented (Polyak, 1979) as

Now, let us assume that the cross-covariance function

and JV and k are so large that for

M(V,T) is close to zero. Then


3.6 Two-Dimensional Fields 135

The inverse transform is

We can see that the cross-periodogram and the cross-covariance func-


tion are interconnected by the two-dimensional Fourier transforms.
By analogy with the stationary random processes, the two-dimen-
sional cross-spectral density is estimated by smoothing the corre-
sponding cross-periodogram with the aid of the two-dimensional dig-
ital filter (spectral window):

Substituting (3.89) for Gp+jtq+i in (3.91) we have

where

Filter aji is called the two-dimensional spectral window and function


wvr is a corresponding two-dimensional correlation window.
To get estimates for points Gpq close to the boundaries, field Gpq
must be periodically expanded in accordance with the symmetry of
the Fourier coefficients (see Figures 3.1 and 3.2) and the fact that

The accuracy of the estimates can be found under the assumption


that, in each rectangular area equal to the size of the filter, the values
of Gpq are approximately independent and have equal accuracy with
variance a^. In this case, variance crJL of estimate Gpq is
136 3 Random Processes and Fields

3. To obtain spectral and correlation estimators for one random field,


we can assume that Ygt = Xgt. Then (3.87) yields

The statistic MVT is called the sample covariance function of a two-


dimensional random field. The corresponding periodogram is

and the estimation of the spectral density is performed by smoothing


IIPq with the aid of a two-dimensional digital filter.
The main feature of the two-dimensional numerical scheme is
that the cross-covariance function is determined by the four different
types expressions (3.87) (one for each quadrant), instead of the two
types, as in the one-dimensional case. This fact must be considered
when developing methodology that estimates the covariance function
as the first step and the spectral density as its Fourier transform.
The key to the practical application of the algorithms considered
lies in the construction of the appropriate two-dimensional filter. The
simplest approach is the superposition of the simplest filters (1.143).
The high-order filter (1.144) cannot be applied because it has nega-
tive weights, which can lead to negative spectral estimates.
Solving specific physical problems connected with an estimation
of the two-dimensional spectral density makes it necessary to cre-
ate different classes of two-dimensional digital filters with positive
weights.
Let us conclude this section with some methodological remarks.
1. When representing the graphs of the estimates of correlation and
spectral fields, one must place the origin of the coordinate system in
3.6 Two-Dimensional Fields 137

the center of the field. The correlation functions must be placed in


the lags' domain (v, r), where

and

The periodograms and spectra must be placed in the frequency do-


main (flp,u;g), where

This presentation emphasizes the symmetry of the estimates in the


case of the only field. The values of the correlation functions with
the negative lags correspond to the values of the periodograms and
spectra with the negative frequencies.
2. Increasing the dimensionality creates some new problems that
must be solved. For example, different two-dimensional meteorolog-
ical fields can have different deterministic trends (latitudinal, sea-
sonal cycle, and so on) that must be removed beforehand in order
that the sample analyzed will satisfy the condition of homogeneity.
Therefore, one must begin with estimation of the trend and sepa-
ration of the deterministic and random components, or filtering out
a two-dimensional trend of the mean and the variance. Although
the estimation algorithms can be different, their analysis must be
carried out simultaneously with the estimation of the spectral and
correlation functions. Different filtration methodologies can lead to
different final results. After filtration, the random field must have a
zero mean and a variance that is constant for all points of the field.
Further calculations must be done in the standard way.
All of the two-dimensional periodograms and spectra given in this
book were obtained with the aid of the normalized observations for
which the sample mean equals zero and the sample variance equals
one.
The methodology considered presents the standard numerical
technique for estimating the correlation functions and the spectra
138 3 Random Processes and Fields

of the two-dimensional fields. The convenient way for software de-


velopment is to use the fast Fourier transform of the normalized ob-
servations, adding the zero pads to make the sizes of the field equal
to the powers of two.
3. Reasoning about univariate computational methodologies given in
Section 3.1 remains valid for the two-dimensional case. The combina-
tion of the Blackman-Tukey and Cooley-Tukey approaches (Fourier
transform of the field --> periodogram —> Fourier transform of the pe-
riodogram —> correlation function —> truncating and weighting of the
correlation function —» its Fourier transform —> spectrum) with three
Fourier transformations is the most effective algorithm for software
development. Using this algorithm, one can avoid direct smooth-
ing, which presents algorithmic and programming difficulties for the
points close to the boundary of the two-dimensional frequency do-
main.
Currently in climatology, the univariate spectral and correlation
analysis of time series are widely used. Many results, devoted to
the study of the character of the temporal fluctuations of different
meteorological elements with periods ranging from, seconds to mil-
lenniums, have been published. Applying multidimensional spectral
analysis will make it possible, as we will see shortly, to obtain more
general and interesting results.

3.7 Multidimensional Fields


The estimators for the spectral and correlation characteristics of mul-
tidimensional fields can be obtained by analogy with the one- and
two-dimensional cases.
Let us consider multidimensional random field

where /x is dimensionality; kj is the size (number of points) along the


tj axis.
All the formulas for the spectral and correlation analysis of this
field (together with the analogous formulas for the one- and two-
dimensional cases) are given in Tables 3.1 and 3.2.
To smooth the multidimensional periodogram, one must create a
corresponding multidimensional filter (see Sections 1.12). As a first
step in developing such a, filter, it is possible to use the simple moving
average procedures analogous to (1.143).
Table 3.1: Discrete schemes of the Fourier analysis of time series and fields.

Univariate time series Two -dimensional field

Sample

Inverse Fourier transform

Fourier transform

Periodogram

Continued)
Table 3.1 9Cont.)

Parseval's identity

Fourier expansion

of periodogram

Approximate Fourier expansion of


periodogram

Fourier expansion of covariance func-


tion

Estimator of spectral density


Multidimensional field

Sample

Inverse Fourier transform

J_*
t—i
Fourier transform

Periodogram

Parseval's identity

(Continued)
Table 3.1 (Cont.)

Fourier expansion of peri-


odogram

Approximate Fourier ex-


pansion of periodogram

Fourier representation of
covariance function

Estimator of spectral den-


sity
Table 3.2: Discrete schemes of the cross-Fourier analysis of time series and fields.
Univariate time series Two-dimensional fields
Sample

Inverse

Fourier transforms

Fourier

transforms

Cross-periodogram

Fourier representation of
cross-periodogram

(Continued)
Table 3.2 (Cont.)

Approximate Fourier
representation of cross-
periodogram

Parseval's identity

Fourier representation of
l cross-covanance function

Estimator of cross-spectra!
density

Coherency function

Phase difference
Multidimensional fields
Sample

Inverse Fourier

transforms

Fourier transforms

Cross-periodogram

Fourier representation of
cross-periodogram

(Continued)
Table 3.2 (Cont.)
Approximate Fourier
representation of cross-
periodogram

Parseval's identity

Fourier representation of
cross-covariance function

Estimator of cross-spectral
density

Coherency function

Phase difference
3.8 Examples of Ciimatological Fields 147

Because there are standard routines for multidimensional Fourier


transformations, it is not difficult to develop software for spectral
and correlation analysis of multidimensional fields. The method-
ological problem (even for a three-dimensional case) one can meet
in the pictorial representation of the results. The principle con-
ceptual problems are in an appropriate transformation of different
non-homogeneous fields into homogeneous ones.

3.8 Examples of Climatological Fields


Climatological studies can consider many different types of observed
and simulated fields. The analysis of each field requires a special
approach. First, a special scheme to filter out two-dimensional trends
(of the mean and the variance) and to transform the original field
into the form with the zero mean and unit variance is required. But
even after conducting such a procedure, a meteorological field would
remain nonhomogeneous because, as a rule, any of its correlations are
spatially (latitude and/or altitude) and temporally (climate change)
dependent. The practical way to overcome this inconvenience is by
reducing the field size and increasing its resolution.
Let us enumerate some types of fields found in climatology and
give examples of their spectral and correlation estimates. We denote
a random field as Y(£,ri}.

3.8.1 The 500 mb Surface Geopotential Field.


Variables £ and n are the rectangular coordinates x and y, obtained
by projecting the geographical latitude-longitude coordinates onto
the plane of the standard isobaric surface. Examples of such fields
are the hemisphere fields of air temperature, atmospheric pressure,
and so on in the fixed moment of time. It is clear that such fields can
have a latitude trend. They are also distorted because the earth's
spherical shape restricts the possibility of accurate study of the part
of the spatial spectrum with high wave numbers. Such fields occur
frequently in climatology, and their spectral and correlation analysis
has significant interest in many areas.
Let us consider observations (in the fixed moment of time) of
the 500 mb level geopotential given as a 32x32 ( x , y ) rectangular
field with the origin in the North Pole and the spatial steps equal to
about 570 km along each of the axes. Such a field does not satisfy the
homogeneous conditions because of its latitude trend (and, may be,
148 3 Random Processes and Fields

Table 3.3: The 500 mb geopotential field spatial correlation function


estimates.

Lags (103 km) Two-dimensional Fortus Udin


analysis

0.0 1.00 1.00 1.00


0.6 0.83 0.78 0.90
1.2 0.60 0.47 0.67
1.7 0.33 0.17 0.47
2.3 0.10 -0.08 0.32
2.9 -0.02 -0.20 0.21
3.5 -0.03 — —

latitude dependent other statistical characteristics). To remove this


trend, the field was approximated by the truncated two-dimensional
Fourier set in which only the zero and the first two-dimensional har-
monics were retained. The field of the deviations of the original
observations from the point estimates of this truncated Fourier set
was then analyzed.
The spectral estimates (Figure 3.3, top) of this field of devi-
ations was found by smoothing the periodogram by the simplest
(1.143) two-dimensional scheme with the widths s = r = 2. The
spectrum has a widespread maximum, centered in the origin and
slightly stretched along one of the diagonals. The corresponding
two-dimensional correlation function (Figure 3.3, bottom) shows that
positive statistical dependence can be traced up to 2000 km, and the
isolines are only slightly different from the circles. Therefore, this
field is close to the isotropic one that confirms the approximate va-
lidity of the assumption about the isotropic character of such 500 mb
surface geopotential fields, which has been accepted in some publi-
cations (Gandin, 1965; Gandin and Kagan, 1976). The isotropic
analog (of the correlation function in Figure 3.3), obtained by aver-
aging two-dimensional estimates equally distant from the origin, is
given in Table 3.3.
For comparison, two other isotropic correlation functions (from
Gandin and Kagan, 1976) are also presented in this table. These two
correlation functions were obtained by M. I. Fortus and M. I. Udin
3.8 Examples of Climatological Fields 149

Figure 3.3: Two-dimensional spectral density (top) and correlation function


(bottom) of the 500 mb geopotential surface.
150 3 Random Processes and Fields

from a long-term time series of observations. Comparison shows that


the values of our estimates are situated between the results computed
by Fortus and Udin. Therefore, in this case the two-dimensional ap-
proach was very effective because the volume of long-term observa-
tions, used by Fortus and Udin, is greater by many times than the
number of observations of the above considered field.

3.8.2 Latitude-Temporal Fields


Another type of two-dimensional structure widely used in climatol-
ogy is the latitude-temporal field, in which £ is latitude and ?/ is time.
For instance, such fields can be obtained by spatially averaging data
in different latitude bands (see, for example, Oort, 1983). These
fields can have spatial (latitudinal) as well as temporal (seasonal cy-
cle) trends that must be filtered out. As a rule, such field must be
split along the latitude direction into (at least) three subfields corre-
sponding to the tropical, rnidlatitude, and polar regions for each of
which the violation of the homogeneity would not be very significant
(see Chapter 7).
Let us consider three monthly mean observed fields of thickness
(500 1000 mb), geopotential, and zonal component of geostrophic
wind of the 500 mb surface for the midlatitudes of the Northern
Hemisphere with the observational interval 1948 through 1978. The
source of data is the archives of the World Data Center, Obninsk,
Russia. These fields were obtained by data averaging in different 5°-
width latitude bands, centered in the points 30°, 35°, ..., 80°. More
elaborate analysis of these data has been provided by Polyak, 1979.
Here we will not discuss questions connected with the data accu-
racy conditioned by the temporal and spatial variation of the volume
of observations. Such discussion can be found in Oort, 1978.
The filtering out of the trend of this data was done by estimating
the monthly means and standard deviations for the time series of
each latitude band and standardizing the corresponding deviations
from these means. Spectral estimates, obtained by smoothing cor-
responding two-dimensional periodograrns by the simplest scheme
(1.143), are given in Figures 3.4 through 3.6 (top).
The geopotential and wind spectra have a very small maxima
centered in the origin, while analogous spectral maximum of rela-
tive topography is slightly more noticeable. Therefore, the thickness
fluctuations are mainly realized by means of the low-wave-number
spatial and, to lesser degree, by low-frequency temporal waves. Cor-
responding correlation functions, presented in Figures 3.4 through
3.8 Examples of Climatological Fields 151

Figure 3.4: Central parts of two-dimensional spectrum (top) and correlation


function (bottom) of the thickness (500 mb-1000 mb) latitude-temporal field
152 3 Random Processes and Fields

Figure 3.5: Centra! parts of two-dimensional spectrum (top) and correlation


function (bottom) of the 500 mb surface geopotential latitude-temporal field.
3.8 Examples of Climatological Fields 153

Figure 3.6: Central parts of two-dimensional spectrum (top) and correlation


function (bottom) of the 500 mb surface geostrophic wind latitude-temporal field.
154 3 Random Processes and Fields

3.6 (bottom), demonstrate significant spatial statistical dependence


of data. Spatial correlation coefficients for the 5° latitude lag are 0.53
for the relative topography and geopotential and 0.46 for the wind.
Temporal statistical dependence is smaller: Correlations of the same
variables for the one month lag are 0.37, 0.21, and 0.20 respectively.

3.8.3 Monthly-Annual Fields


The next type of dimatological field is the monthly-annual field of
the monthly mean time series when £ is months (12 points) and ?/ is
years. Considering a monthly time series as a field enables us to split
the interannual and intraannual variability for comparative study.
This type of field has a temporal (seasonal cycle) trend that must be
removed before the spectral analysis is done.
Time series of monthly means for each of the latitude band of
any of the three fields considered above (thickness, geopotential,
and zonal component, of geostrophic wind of the 500 mb level of the
Northern Hemisphere) can be presented and analyzed as a separate
field. Because the total number of such time series for the variables
considered above is too large, we will study only one of them (corre-
sponding to the 55° latitude band) for each of the variables.
Spectra of these three fields are given in Figures 3.7 through 3.9
(top). An interesting feature of the thickness and geopotential spec-
tra is the well-known small high-frequency maxima corresponding to
the period of about two years. Additionally, the thickness spectrum
has another small, wide maximum centered in the origin.
As for the geostrophic wind, the power distribution over the fre-
quency domain is more or less even, and corresponding observations
are very close to the sample of the two-dimensional white noise. The
same conclusions can be drawn from consideration of the correlation
functions in Figure 3.9 (bottom), because estimates of interannual
(as well as intraannual) correlations are very close to zero.
Geopotential (Figure 3.8) exhibits small autocorrelation (0.2) at
one-month lag and even smaller negative correlations (-0.12) for the
one-year lag, which is the cause of the high-frequency maximum.
The two-dimensional correlation function of the relative topogra-
phy (Figure 3.7) reveals a small temporal correlation (about 0.3) for
the one-month lag.
As will be shown in Section 4.7, the most plausible conclusion
from the above study is that the analyzed information has no fore-
casting value.
3.8 Examples of Climatological Fields 155

Figure 3.7: Central parts of two-dimensional spectrum (top) and correlation


function (bottom) of the thickness (500 mb - 1000 mb) monthly mean time series.
156 3 Random Processes and Fields

Figure 3.8: Central parts of two-dimensional spectrum (top) and correlation


function (bottom) of the 500 mb surface monthly mean time series of geopoten-
tial.
3.8 Examples of Climatological Fields 157

Figure 3.9: Central parts of two-dimensional spectrum (top) and correlation


function (bottom) of the 500 mb surface monthly mean time series of geostrophic
wind.
158 3 Random Processes and Fields

3.8.4 Altitude-Temporal Monthly Temperature Field


Simulated by the Hamburg GCM
The next type of climatological field is the altitude-temporal field,
in which £ is altitude and TJ is time. This type of field presents the
vertical cross-section of the three dimensional observed or simulated
structure (latitude, altitude, time).
As Matrosova and Polyak (1990) have shown, such temperature
fields are highly nonhomogeneous, or their first- and second-moment
values vary with altitude. In particular, such fields of monthly data
have a spatial trend (conditioned by the variation of the mean as a
function of the altitude) as well as a seasonal cycle. For the example
provided below, such trends were removed. Nonetheless, the results
have a pure formal character because the correlations remained non-
homogeneous. To get a more accurate analysis, one must split this
field along the altitude into at least four (Matrosova and Polyak,
1990) vertical subfields of a smaller size, simultaneously increasing
their resolution. Such a procedure can be done with a specially de-
signed simulation experiment.
Consider a vertical cross-section of the temperature data, simu-
lated by one of the Hamburg GCMs (see Chapter 7). The she of the
field is 10 (altitudes, from 1000 mb to 100 mb with a vertical step
equal to 100 mb) X 372 (temporal points, 31 years of data with a
one-month step). For each altitude level, data were averaged within
a 10°-width latitude band (5° to 15°) of the Northern Hemisphere.
The seasonal cycle was removed.
The estimated spectrum (Figure 3.10, top) shows a concentration
of power along the frequency axis with a significant maximum near
the origin. This indicates that the major simulated variability is
taking place in time and that the data from the model has a trend.
The most powerful vertical fluctuations occur because of the long-
length waves (greater than 350-400 mb).
The temporal statistical dependence of fluctuations (Figure 3.10,
bottom) primarily indicates an intraannual character; the vertical
correlation of data can be traced up to Jags of 600 mb.
Of course, these examples do not exhaust all the field types, but
this formal, consideration demonstrates a diversity of climatological
fields and their two-dimensional spectral and correlation characteris-
tics. The estimates obtained were not accompanied with any physical
analysis (as well as with accuracy estimation) because such insight
requires special wording of climatological problems. Such problems
will be considered in Chapters 7.
3.8 Examples of Climatological Fields 159

Figure 3.10: Spectrum (top) and correlation function (bottom) of the altitude-
temporal simulated temperature field for 5° to 15° latitude band.
160 3 Random Processes and Fields

As has been mentioned, the principal obstacle to the two-dimen-


sional spectral and correlation analysis of meteorological fields is
in their nonhomogeneous statistical structure. One of the practi-
cal ways to overcome this obstacle is to reduce the field size while
simultaneously increasing its spatial resolution. There are no prin-
ciple difficulties in such procedures for the simulated data: a GCM
must be integrated with a corresponding spatial grid. As for the
observed, data, it will require a new means of measurements, and
especially the satellite observations.

3.9 Anisotropy of Climatologicai Fields


The spectral and correlation estimation schemes above were deduced
for homogeneous random fields. As a rule, climatological fields are
not homogeneous (and, of course, not isotropic). Processing their
samples by the above estimation procedures leads to some approx-
imations, although the accuracy varies for different elements and
types of sampling. Corresponding reasoning about possible scales
of distinction of the particular climate field under study from the
homogeneous one can be done by analyzing its specific physical and
statistical nature. In any case, as a rule, spectral and correlation
estimates of climatological fields carry new information that some-
times helps to get approximate solutions to the problem considered
or at least to make a step in a right direction.
However, for some meteorological variables and sampling types,
the homogeneity assumption is nearly satisfied. Moreover, spatial
correlation functions of the two-dimensional fields (of the above geopo-
tential type) for many elements (air temperature, atmospheric, geopo-
tcntial, and so on) were estimated under the assumption of isotropy
and used in the objective analysis procedures (Gandin and Kagan,
1976).
Furthermore, some types of homogeneous random fields can be
transformed to isotropic ones by rotating the coordinate system and
scaling the axes differently. For this purpose, one can use, for exam-
ple, modeling by the Laplace equation in accordance with the results
of the paper by Jones and Vecchia (1993).
Consider the following two-dimensional stochastic Laplace equa-
tion:

where a is positive and e ( x , y ) is two-dimensional white noise. The


solu-tion £(x, y) for this equation is an isotropic random field. The
3.9 Anisotropy of Climatological fields 161

correlation p(r) between two points of this field at distance r apart


is

where K1 is the modified Bessel function of the second kind, order


one. The corresponding standardized rational spectral density is

If the new coordinate axes are u and v, the rotation of the coordinates
by an appropriate angle a is the standard transformation

The scaling of the two coordinate axes differently is carried out by


an additional positive parameter A2. The scaled distance between
observations i and j is

Therefore, to transform an anisotropic field to an isotropic one, two


parameters, a and A, must be estimated.
Practical applicability of this methodology can be evaluated by
careful consideration of the isolines of the two-dimensional correla-
tion function estimates to reveal that the anisotropic character of the
field is not very complicated. As a result, it is possible to transform
the field into an isotropic one by rotating and scaling coordinates. If
the angle a, as well as the scaling parameter A2, is not constant and
differs for different points of the field, this transformation cannot, be
done. The methodology (Jones and Vecchia, 1993) of transforming
the anisotropic field into an isotropic field has many applications in
climatology.
4

Variability of ARMA
Processes

In this chapter, the numerical and pictorial interpretation of the de-


pendence of the standard deviation of the forecast error for the dif-
ferent types and orders of univariate autoregressive-moving average
(ARMA) processes on the lead time and on the autocorrelations (in
the domains of the permissible autocorrelations) are given. While
the convenience of fitting a stochastic model enables us to estimate
its accuracy for the only time series under consideration, the graphs
in this chapter demonstrate such accuracy for all possible models of
the first and second order. Such a study can help in evaluating the
appropriateness of the presupposed model, in earring out the model
identification procedure, in designing an experiment, and in opti-
mally organizing computations (or electing not to do so). A priori
knowledge of the theoretical values of a forecast's accuracy indicates
the reasonable limits of complicating the model and facilitates evalu-
ation of the consequences of certain preliminary decisions concerning
its application. The approach applied is similar to the methodology
developed in Chapters 1 and 2. Because the linear process theory has
been thoroughly discussed in the statistical literature (see, for exam-
ple, Box and Jenkins, 1976; Kashyap and Rao, 1976; and so on), its
principal concepts are presented in recipe form with the minimum of
details necessary for understanding the computational aspects of the
subject.

162
4-1 Fundamental ARM A Processes 163

4.1 Fundamental ARM A Processes


Consider a discrete stationary random process Zt with null expected
value [E(^) = 0] and autocovariance function

where cr2 is the variance and p(r] is the autocorrelation function of


zt.
Let at be a discrete white noise process with a zero mean and a
variance a\. Let us assume that processes zt and a< are normally dis-
tributed and that their cross-covariance function Mz a (r) = 0 if r >
0. Let us also assume that zt can be presented as

where <fi(i = 1,2, . .,p) and 0 j ( j = 1,2, . . . , < / ) are parameters that
satisfy the following conditions: The roots of the equations

and

are outside the unit circle (|a;| > 1). In this case, zt is referred to
as the autoregressive-moving average process of p, q order, and it is
denoted as ARMA(p, q). Parameters <pi are referred to as autoregres-
sive parameters, and parameters 9j are referred to as moving average
parameters. The total number of parameters of the ARMA(p, q) pro-
cess is equal to p + q + 1 (</?;,#j, and <r^). The condition (|,T| > 1)
determines the domain of the permissible values of parameters (pi
and Oj.
Multiplying (4,2) by zt_T and taking the expectation of the prod-
uct yields the following equations:

Letting r = 1 , 2 , . , .,p+q in (4.5) leads to a system oip+q equations,


which connect (in a single way) parameters </?; and Oj with the first
164 J, Variability of ARM A Processes

p + q autocovariances and the first q cross-covariances. This system


[together with conditions (4.3) and (4.4)] reveals that the domain of
the permissible values of the parameters of the ARMA(p, q} process
determines the domain of the permissible values of the autocorrela-
tions.
Formula (4.2) enables us to forecast process Zt one step ahead.
In such forecasting, the value of at is neglected and the forecast error
variance is equal to the variance a\. By using (4.2) as a recursive
expression for forecasting zt for l+l steps ahead, we obtain value zt+i.
The variance v2p,q(l) of the forecast error (zt+i — zt+i) is determined
by the formula (Box and Jenkins, 1976)

where

Together with Vp<q(l), the normalized (standardized) variance

will be considered.
Later, when misunderstanding cannot arise, subscripts p and q
will be omitted; that is, e(l} will be written instead of £p,q(l) and v(l)
instead of v p i 9 (/).
Our goal is to study the theoretical limits of the forecast accuracy
of the ARMA(p, q) processes in order to outline reasonable, practical
situations for fitting a mode! and forecasting. Consequently, the
main object of the numerical and pictorial interpretation will be the
normalized standard deviation e(l) of the forecast error for /+ 1. steps
ahead.
A fitting of the ARMA(p, q) process to a time series is provided
by identifying the values of p, q and estimating the corresponding
parameters <fi,0j. The basis for such estimation is the system of
p -j- q equations [obtained from (4.5) for T = 1, 2 , . . . ,p + </], where
the values of the auto- and cross-covariance functions are replaced
by their estimates.
4.2 AR Processes 165

4.2 AR Processes
If the stationary process zt and white noise at are statistically in-
dependent, then all parameters 6j are equal to zero. The process
(4.2) is then called the autoregressive process of order p [abbreviated
AR(p)]. It is presented by the formula

The number of the AR(/>) process parameters equals p + 1 (</?,- and


<Ta)-
From (4.5) and (4.1), the system of linear algebraic equations
referred to as the Yule-Walker equations can be written as

This system connects parameters (f{ with the first p autocorrelations


P(r).
From (4.6), we have

where

is the squared multiple correlation coefficient of zt and zt-1, Zt-2-, • • •,


z
t—p.
Therefore, all p + 1 parameters of the AR(p) process and the
values of its autocorrelation function are interconnected in a unique
way.
Let e2 = ££<,(!). Then (4.12) yields

in other words, the normalized variance of the forecast error for one
step ahead and the squared multiple correlation coefficients mutually
determine each other.
By denoting the multiple correlation coefficient of zt+i and
Zt_i, 2 ( _ 2 , . . -, zt-p as r 2 (l), we get the relationship

Formula (4.14) is a particular case of (4.15).


166 4 Variability of ARM A Processes

The AR(p) processes can be presented in matrix form in the


following way:
Let us introduce the notation

Then

where

is the matrix of the Yule-Walker system; the T is the sign of trans-


position.
The algorithm of forecasting for / steps ahead can be presented
in the following way:

As we will see in the next chapter, such a notation does not dif-
fer from the corresponding representation of the algorithm for the
multivariate case.
4.3 AR(1) AND AR(2) Processes 167

Figure 4.1: Dependence of the normalized standard deviation e(J) of the forecast
error of the AR(1) process on autocorrelation p and on lead time /.

4.3 AR(1) and AR(2) Processes


If p = 1, formulas (4.10) to (4.12) yield

The normalized variance £ 2 (/) of the forecast error for / steps ahead
is obtained from (4.6) and (4.7) by the simple rearrangement

The value of e(l) as a function of parameters p and / is presented in


Figure 4.1.
We can see that, if p < 0.3, the standard deviation of the forecast
error for any I is greater than 0.95 (or the standard deviation of the
forecast error constitutes more than 95% of the standard deviation
of process zt). The practical usefulness of such a forecast is doubtful.
If 0.3 < p < 0.5, the value of e ( l ) varies from 0.87 to 0.95; that is,
a forecast, even for one step ahead, has a very low accuracy; if / is
greater than 1, such a forecast will likely have no practical value. If
p is about 0.6, the forecast makes sense for not more than two steps
ahead, and in this case e(/) is varied from 0.8 to 0.93. If p is about
0.7, the forecast can be considered for two or three steps ahead. It is
168 ^ Variability of ARMA Processes

only if p is greater than 0.8 that the forecast for four of more steps
ahead is meaningful.
Because dependence of the variance on the autocorrelation (4.26)
is continuous, the above statement about the boundaries, which sep-
arate a reasonable forecast from a senseless one, is purely subjective.
However, from (4.26) and its numerical interpretation, it is clear that
for a first-order process each value of \p\ < I determines a theoretical
limitation on the forecast accuracy. In particular, small autocorre-
lations (p < 0.3) (or their estimates, even statistically significant)
mean that corresponding forecasts will be very close to the mean
value (zero in our case) and that such forecasting makes no sense,
even for one step ahead.
The above assertions become especially obvious when one con-
siders the dependence of the lead time / on autocorrelation p and s.
In this case, (4.26) yields (for /» 0)

This dependence, presented in Figure 4.2, demonstrates a useful


fact: that with p increasing from 0 to about 0.6, a possible lead time
/ increases very slowly, and a forecast for more than two steps ahead
(as has already been, shown) is senseless.
It is only for p > 0.7 (up to 1) that the possible lead time begins
to increase rapidly, converging to the infinite value (when p is equal
to 1 and the forecast is deterministic). The detailed consideration
of the simplest AR(1) process is justified by the fact that, as will be
shown shortly, the above numerical interpretation of the relationship
between autocorrelation p and the forecast accuracy (or lead time)
is approximately correct for the processes of any orders. Parameter
e(/) is conveniently used in climatology as a primary characteristic
of the forecast accuracy; therefore, it will be of constant interest to
us.
The second-order autoregressive process

has the Yule-Walker system


4.3 AR(1) AND AR(2) Processes 169

Figure 4.2: Dependence of the lead time of the AR(1) process on autocorrelation
p and on standard deviation e(l).

the matrix of which

has the inverse matrix

Using (4.12), the normalized variance £2 of the white noise a< is


presented as

The AR(2) forecast accuracy is determined by the first and second


autocorrelations [p(l) and p(2)}. By defining a two-dimensional grid
170 ^ Variability of ARM A Processes

on the plane [p(l), />(2)] with the domain of the permissible autocor-
relations and performing the computations by (4.12) for each point
of the grid, it is possible to obtain the fields of the normalized stan-
dard deviation i(l] for different /. The results of such computations
are symmetrical relative to the plane [/o(2), £(/)], and for this reason,
they are presented in Figures 4.3 and 4.4 only for p ( l ] > 0. These
figures show the shapes and corresponding isolines of the e(l) sur-
face as a function of the autocorrelations; that is, they illustrate the
character of the variations of the theoretical accuracy of the forecast
corresponding to different lead times. If 1 = 1, the part of domain
[/>(!), p(2)} where e(l) is close to 1 [i.e., e(l) > 0.95] is not large. With
increasing /, this part of the domain gradually increases, spreading
over almost the entire area of the permissible autocorrelations. The
closeness of £ to 1 means that variance «2(7) of the forecast error is
close to variance a2 of process zt and that the forecast is senseless in
the sense that it does not differ markedly from the mean.
The greater />(!) and /?(2) (i.e., the closer the point is to the
boundary of the domain), the higher the accuracy of the forecast.
If the statistical predictability of a time series is studied, one must
evaluate the possible lead time of a forecast for the fixed accuracy
values. In this case, together with the results presented in Figures
4.3 and 4.4, the lead time I field [as functions of />(!) and p(1} and
parameter s] can give a clearer picture of the predictability limits of
the autoregressive processes.
Let us consider the numerical interpretation of the fields of / for
the two fixed values of the parameter e (e = 0.95 and £ = 0.8). The
results (Figure 1.5) show that the lead time is less than one (step
ahead), if p ( l ) < 0.3 and p(2) < 0.3. Even for the minimal accuracy
(e — 0.95), one can speak about a forecast for only two steps ahead,
if />(!) > 0.5 and p(2) > 0.3, and so on.
Therefore, the requirements for the autocorrelation values are ap-
proximately the same as those for the AR(1) process, and the fore-
cast accuracy (and the lead time) of the second-order AR processes
is also approximately the same as for the AR(1). Of course, a class
of physical processes, which can be approximated by the AR(2) is
extended: One can obtain a more detailed and accurate description
of the approximated time series, including some subtle features of
spectral density. But the significant increase in forecast accuracy
and in the lead time cannot be expected in theoretical or practical
terms. Later, this assertion will be illustrated more clearly.
Information presented in Figures 4.3 to 4.5 simplifies the prelim-
inary analysis and identification of a model, as well as the accuracy
evaluation and the building of a forecast confidence interval.
4.4 Order of the AR Process 1.71

Figure 4.3: Normalized standard deviation e(/) of the forecast error of the
AR(2) process for I =1 to 5.
172 Variability of ARM A Processes

Figure 4.4: Contour lines of the structures in Figure 4.3.


figure 4.5: Dependence of the lead time / of the AR(2) process on the autocorrelations p(\) and p(2)
and on the normalized standard deviation e(() of the forecast error.
174 4Variability of ARMA Processes

The advantage of such an approach is that it permits one to


formulate the requirements to the statistical characteristics of the
time series, which guarantee the desired forecast accuracy (or lead
time). Such an analysis can also be helpful when working with a
time series with little data. In this case, it is possible to find the
confidence intervals of the autocorrelation estimates which determine
the corresponding domain in lag space which, in turn, determines
some region of the normalized standard deviations of the forecast
error.

4.4 Order of the AR Process


The numerical and pictorial interpretation, analogous to those in
Section 4.3, for the high-order (greater than 2) autoregressive pro-
cess is cumbersome, because the elements of the large-size tensors
must be computed and illustrated. Instead of obtaining a complete
theoretical picture of the dependence of the forecast accuracy of any
AR processes on possible autocorrelations, lead times, and orders,
we simplify the problem and consider an example with a special type
of autocorrelation function:

and p ( r ) —.> 0 if r —> oo. To determine parameters ipi (i = 1, 2,... ,p)


of the corresponding autoregressive model

it is necessary to find the solution of the Yule-Walker system

This solution
4-4 Order of the AR Process 175

makes it possible to present (4.34) as

It is clear that the maximum weight <p in (4.37) obtained for p equals
1 when the process is a linear function.
The identity of all the autocorrelations (4.33) leads to the identity
of all the parameters in (4.34) as well as to the identity (indepen-
dence of lead time) of the standard deviations of the forecast errors.
Substituting <f> and p in (4.12) and (4.13) for (4.36) and (4.33) yields

Formula (4.38) describes the behavior of the variance as a function


of parameters p and p. If p is equal to zero, then e2 = 1, and the
forecast is equal to the mean of the process. If p = 1, then £2 = 0,
and the forecast is deterministic because separate values of zt are
linearly dependent.
If p — 1, then

is the normalized variance of the forecast error of the AR(l) process.


If p increases, the e2 monotonically decreases and

This fact shows that the normalized variance of the forecast error is
limited from below (by the quantity 1 -p), and building up the order
cannot lead to the unlimited decrease of £2.
A more detailed study of the behavior of the normalized standard
deviation e can be done by computing values (4.38) for different p > 0
for some two-dimensional grid points of p and p.
The results of such computations (Figure 4.6) show that for small
values of p (p < 0.2), the normalized standard deviation e is greater
than 0.9 independent of the order. In other words, it is equal to 90%
or more of the standard deviation of the process.
Such a forecast is insufficiently deviated from the mean, and its
estimation will not make any practical sense. In the band 0.2 <
p < 0.3 the normalized standard deviation gradually decreases (with
increasing p) and (if p > 5) becomes very close to the value 1 — p
176 4 Variability of ARM A Processes

Figure 4.6: Normalized standard deviation e of the forecast error as a function


of autocorrelation p and order p.

(0.84-0.9); so the practical usefulness of increasing order (p > 1) is


doubtful.
As will be shown in Chapter 6, which deals with the monthly and
annual means of the historical records, the range (0.0-0.3) contains
the values of the autocorrelation estimates of climatological time se-
ries of separate stations. Therefore, the above inferences characterize
the possible accuracy of the forecasts of such a time series if they are
approximated by the AR process. The results show that the benefit
of forecasting by corresponding AR models is very low, if it has any
benefit at all.
If 0.3 < p < 0.5, the processes of the first through the fourth or-
ders can provide relatively acceptable forecast accuracy (e = 0.84,...,
0.71); however, building up the order further does not increase the
accuracy. (For the first-order moving average process, the interval
\P\ < 0.5 presents the entire domain of the permissible autocorre-
lations; see the following section.) In this domain, the model type
identification must distinguish between the autoregressive and the
moving average model.
The values p > 0.6 can ensure a sufficiently high accuracy (e >
0.64); therefore, fitting AR models can be of significant interest for
the time series with such correlation values.
4-4 Order of the AR Process 177

The results (Figure 4,6) show that increasing the forecast accu-
racy by increasing the process order occurs very slowly, especially if
p > 3. For this reason, if p > 4, inaccuracy in one or two units in
the order identification cannot result in a significant increase of the
forecast error.
From (4.39) and (4.40) one can derive the inequality

which determines the value band lor the normalized variance of the
forecast error of the autoregressive process independently of its order.
The above reasoning can be summed up with the following com-
ments,
1. Evaluating the forecast accuracy of the AR processes, one can
notice that the principal meaning presents not simply the non-
zero autocorrelation between the terms of the time series but,
primarily, its value. For instance, if 0.1 < \p\ < 0.3, the statis-
tical dependence of the terms of time series is obvious but the
corresponding forecast offers virtually no helpful information.
2. Large autocorrelations are the decisive factor for providing a
high-accuracy forecast. Identification of the model's order can
only slightly improve it. For this reason, sufficiently increasing
the order (> 2) of the fitted model can be based mainly on
very accurate physical reasoning. (Spectral analysis presents
significantly higher requirements for the accuracy of the model
order identification than forecasting does.)
3. The results obtained reveal the uncertainty of categorical state-
ments, which can sometimes be found in climatological papers,
about the predictability of various climate time series, when the
latter differs from the sample of white noise. Because e con-
tinuously depends on p, the autocorrelations and the forecast
accuracy mutually determine each other. Therefore, reasoning
about the predictability limits, as well as the practical ascer-
tainment of any actual (fixed) value of £ as an acceptable nor-
malized standard deviation of the forecast error, is subjective
(arbitrary).
The main identification problem is determining an appropriate
order p which guarantees that the All(p) process fits the statistical
and physical properties of the observations. Practically, such identi-
fication is carried out by a multi-step empirical-statistical procedure
178 4 Variability of ARM A Processes

(see Anderson, 1971; EikhofF, 1983; and Hannan, 1970); that is, by
the multiple fitting to given data of the AR(p) process by conse-
quently increasing order ;;. For each step the two following problems
(together with the parameter estimation) must be solved:
1. The evaluation of the statistical significance of the parame-
ter ipp estimate. If size of the sample is large, the simplest
procedure for the evaluation is application of the asymptotic
criterion, stating that the statistic

has an approximately normal distribution with a zero mean


and a variance equal to one.
2. Testing the hypothesis that the order is equal to p versus the
alternative hypothesis that the order is p + s. Such testing can
be carried out with the aid of the following statistic:

where s2p+v(1) and .5^(1) are the estimates of the variance a^


(4.12) for the p + v and p order models respectively. For large
k, quantity kQ2 has x'2 distribution with v degrees of freedom.
Finding the parameter estimates if>i by solving the system of Yule-
Walker equations (4.11), where the theoretical autocorrelations are
replaced by their estimates, is the first part of the process of fitting
an All model to the time series. After that, the estimates of t>(/)
are found by (4.7), and a confidence interval of the forecast for / + 1
steps ahead is constructed by the formula

where v = 2 or 3 respective to the 5% or the 1% level of significance.


4.5 MA(1) and MA(2) Processes 179

4.5 MA(1) and MA(2) Processes


The moving average process zt of order q [MA(q)] presents each value
Zt as a linear combination of the uncorrelated random quantities.
Such a process is denoted as

The process MA(q) is determined by q + 1 parameters 61,62,. • -,$75


and <T(J. It is clear that (4.45) is the particular case of (4.2), if all the
autoregressive parameters are zero.
The equation for the autocorrelation function of the MA(g) pro-
cess is derived from (4.5) as

I he equation tor trie variance aa is derived trorn (4.6) as

Because the values of the MA(q) process are uncorrelated for lags
greater than q, the corresponding forecasts do not make sense for
more than q steps ahead.
The variance of the forecast error for / steps ahead is determined
by the formula

The normalized variance of the forecast error for l steps ahead is


obtained from (4.48) as
180 4 Variability of ARM A Processes

Let us consider normalized standard deviation of the forecast


error for the MA(1) process.
Setting q — \. in (4.45), the MA(1) process is presented as

The autocorrelation function and parameter 0 are interconnected by


the relationship (4.46) (if q — 1)

From this formula, it follows that the estimator of 9 can be found by


solving the quadratic equation

The forecast of the MA(i) process makes sense only for one step
ahead. If q — 1, the normalized standard deviation (4.49) is

The domain of permissible autocorrelations for an MA(1) process is


determined by the inequality

Figure 4,7 enables us to compare the values of e(l) for MA(1) am


AR(i) processes.
This reveals that the MA( 1) process forecast has a slightly highei
theoretical accuracy compared with the A R ( l ) process, but the smal
size of its domain of permissible autocorrelations limits the possibility
of its practical application. If p increases, the distinction of the
forecast accuracy of the MA(1) and AR(1) processes increases. This
distinction is especially large in the subinterval (0.4-0.5), reaching ?
maximum at the endpoint of the domain (if p = 0.5).
Consider normalized standard deviation of the forecast error o:
the MA(2) process.
4-5 MA(1) and MA(2) Processes 181

Figure 4.7: Dependence of the normalized standard deviation e(l) of the fore-
cast error on the autocorrelation p(l) for the MA(1) and AR(1) processes.

The domain of the permissible autocorrelations of the MA(2)


process

is bounded by the lines

This domain is presented in Figures 4.8 and 4.9.


The variance of process (4.55) is

and the relationships between the autocorrelation function p ( r ) and


parameters #,- are obtained from (4.46) (setting q = 2) as

Our interest lies in analysis of the normalized variance of the forecast


error. From (4.49), we have
182 4 Variability of ARM A Processes

Figure 4.8: Dependence of the normalized standard deviation e(l) of the forecast
error of the MA(2) process on autocorrelations p(l) and p(2).

The forecast for / > 2 coincides with the forecast for 1 = 2. The
normalized standard deviations (4.59) are determined by the corre-
sponding parameters which, in turn, depend on autocorrelations.
Having computed parameters 0\ and 6-2 [with the aid of (4.58)]
for some grid points of the p(l) and p(2) from the domain (4.56), one
can use them to compute the values of e(l) and e(2) by (4.59) and to
draw the isolines of the corresponding fields. The numerical outcome
of such computations (Figure 4.8) demonstrates the sensitivity of
the forecast accuracy on the autocorrelation values and on the lead
time. The results show that, even for 1 = 1, forecasts with reasonable
accuracy [for example, with e(l) < 0.9], can be obtained only for the
narrow strip of the points along the boundary of the domain of the
permissible autocorrelations. If I — 2, such a strip is narrowed down
to an even smaller size, which corresponds to the high possible values
of the autocorrelations.
Together with the representation of the results in the form of
the dependence e(/) on autocorrelations /?(!) and /a(2), the study
of the dependence of the lead time l on p(1) and p(2) can be use-
ful. The corresponding numerical results are presented in Figure 4.9.
This figure shows that for very low requirements of forecast accuracy
[E(/) = 0.95], there is a narrow region in the domain of the autocor-
relations in which the forecast for two steps ahead can. have practical
value. If the requirement for accuracy is slightly high [s(l) ~ 0.8],
4.6 ARM A (1,1) Process 183

Figure 4.9: Dependence of the lead time / of the MA(2) process on autocorrela-
tions p(l) and p(2) and on the normalized standard deviation e(l) of the forecast
error.

the forecast for two steps ahead is senseless for almost all points of
the domain.
A comparison of the results in Figures 4.8 and 4.9 with the corre-
sponding results for the AR(2) process (see Figures 4.3-4.5) reveals
that the MA(2) process can provide slightly higher theoretical accu-
racy than the AR(2) can.
However, the domain of the permissible autocorrelations of the
AR(2) process is greater, making its practical use more convenient in
spite of the slight disadvantage in the accuracy value. On the whole,
the normalized variances of the forecast error of the MA(2) process
are smallest on the coinciding parts of the domains of the permissible
autocorrelations compared with other types of the above-mentioned
processes. But this fact does not mean sufficient distinction of the
requirements [obtained for the AR(1) and AR(2)] for the values of
autocorrelations p(l) and p(2), which provide a reasonable forecast.

4.6 ARM A (1,1) Process


The ARMA(1,1) process
184 4 Variability of ARMA Processes

is determined by parameters <f and 9, which satisfy the following


relationships:

The domain of the permissible autocorrelations is determined by the


following inequalities

and it is presented in Figure 4.10.


The variance of the ARMA( 1,1) process is

For the ARMA(1,1) process, the expression for the normalized vari-
ance of the forecast error for / + 1 steps ahead is obtained from (4.7)
as

The parameters (p and 0 [and, consequently, the normalized standard


deviation e(l)] are determined by p ( l ) and p(2). The computation of
e(/) can be done as described below.
Consider any mesh of two-dimensional grid points of p(1) and p(2)
in the domain of the permissible autocorrelations (4.62). Substitut-
ing the values in the grid for p(1) and p(2) in (4.61), the sequence of
the corresponding values of <p and 0 is computed. And, finally, e(l)
(for / = 1,2,3,4,5) is calculated by (4.64). The resulting fields of
the normalized standard deviation e(l) are symmetric [as in the case
of the AR(2) and MA(2) processes] relative to the plane [ j o(2),e(/)];
for this reason, they are given in Figure 4.10 only for p ( l ) > 0. The
isolines of these fields demonstrate the main features of the variation
of the forecast accuracy as a function of the autocorrelations and of
the lead time.
4.6 ARM A (1,1) Process 185

Figure 4.10: Dependence of the normalized standard deviation e(l) of the fore-
cast error of the ARMA(l,l) process on the autocorrelations p(l) and p(2) and
on the lead time /.

The domain of permissible autocorrelations of the ARMA( 1,1)


process is smaller than the corresponding domain of the AR(2) pro-
cess, and its practical application is not as simple as the AR pro-
cess. A detailed analysis reveals that, for the coinciding areas of
the domains of the permissible autocorrelations, the accuracy of the
ARMA( 1,1) process is slightly worse than that of the MA(2) process,
but it is slightly better than for the AR(2) process. The distinction
in the values of e(7) can be clearly seen, especially on the periphery
of the domain of the ARMA(1,1) process.
As we did in previous sections, let us consider the numerical fields
of the lead time I of the forecasts as the function of autocorrelations
/?(!) and p(2) for the two fixed values (0.95 and 0.80) of the normal-
ized standard deviation £(/). Figure 4.11 shows that it is possible
to consider the forecast of the ARMA(1,1) process for three or four
186 4 Variability of ARMA Processes

Figure 4.11: Dependence of the lead time / of the ARMA(1,1) process on the
autocorrelations p(l) and p(2) and on the normalized standard deviation e(l) of
the forecast error.

steps ahead only if the requirement for accuracy is minimal [for ex-
ample, e(l) = 0.95] and autocorrelations />(!) and p(2) are greater
than 0,5-0.8.
If £(7) = 0.8, such a forecast makes sense only for p ( l ) and p(2)
greater than 0.7-0.8, but the corresponding area of the domain of
the permissible autocorrelations is very small.

4.7 Comments
Carrying out the numerical analysis of the simplest ARM A(p, q) pro-
cesses enables us to consider the principal features of the forecast of
the linear process.
1. The variance of the forecast error for / steps ahead is

where a\ = <j 2 e 2 is the variance of the white noise at and e2 is the


normalized variance of the white noise.
The normalized variance of the forecast error for /-f-1 steps ahead
v a2
is
q(l) ~~ p the principal theoretical
q(l)/ ' characteristic of the
4-7 Comments 187

model accuracy. If e(l) is close to 1, the forecast error variability is


close to the natural variability of the process, and the forecast value
is close to the mean. The closer the e(l) is to zero, the more accurate
the forecast. For the standardized time series (a2 = 1), £2(l) is the
precise variance of the forecast error for / steps ahead.
2. The normalized variance £pq(l) of the forecast error of the
ARMA process is functionally connected with the first p + q auto-
correlations and with the lead time /; the values of these parame-
ters mutually determine each other. The numerical interpretation
of the dependence of e(l) (and l) on the autocorrelations was done
for several simple, frequently used processes that clarify the possible
accuracy and lead time of the forecasts.
3. The value of the normalized variance £ 2 (l) converged to one
with increasing lead time l or decreasing autocorrelations. From this,
the conditional character of the forecast behavior inevitably follows:
If £(l) —* 1, the forecast smoothly approaches the mean. In principle,
therefore, it is senseless to apply the stationary model for a forecast
trajectory, which moves away from the mean with increasing /. Time
series predictability is conveniently characterized by a fixed value of
£ (or I). This fixed value always has a subjective meaning because
the dependence between the autocorrelations, accuracy (e), and lead
time (V) is continuous.
4. The investigations carried out show that, when the autocor-
relations are small, the variation of the fitted process type or its
order value does not sufficiently affect the possible forecast accuracy
or the lead time value. Therefore, the accuracy characteristics are
stable, relative to the parameter values and the linear process types,
as are the fluctuations of the samples of stationary processes. The
numerical correspondence between the autocorrelation and the fore-
cast accuracy, established for the AR(1), is approximately the same
for other types of linear processes of any order. This correspon-
dence can be expressed as follows: When the largest autocorrelation
p of the process is less than 0.3, the forecast accuracy is very low
(£ > 0.95), and the forecast has no practical value. If 0.3 < p < 0.5,
the forecast does not make practical sense for more than one step
ahead; if p w 0.6, for more than two steps; and if p fa 0.7, for more
than three steps ahead. The practical consideration of the forecast
for four or more steps ahead makes sense only when p > 0.8. Let
us assume, for instance, that if e(l) > 0.95, the forecast accuracy
is unsatisfactory and its computation makes no sense. Then isoline
e(l) w 0.95 in the corresponding domains of the permissible autocor-
relations (see Figures 4.4, 4.8, and 4.10) outlines the regions with the
Figure 4.12: Increase of the domain [e(l] > 0.95] of the senseless forecasts with
increasing / for all types of the second order ARMA processes.
4.7 Comments 189

senseless [e(/) > 0.95] forecasts. Such isolines, presented in Figure


4.12, clearly demonstrate the growth of the areas within these regions
(with increasing l) for all types of second-order linear processes.
5. The results obtained enable us to evaluate the limits of poten-
tial predictability, corresponding to different autocorrelation values
of the ARMA processes, and to get representation about the possi-
bility of forecasting many physical processes of science and nature,
the autocorrelation structures of which have already been estimated.
One of the fields that benefits from this application is climatology.
The pictorial materials presented can help to simplify the model iden-
tification process. While a particular model, fitted to a given time
series, enables us to estimate its forecast and accuracy, Figures 4.1
through 4.11 allow us to obtain the theoretical accuracy and possible
lead time for all first and the second-order ARMA processes. The in-
tuitive reasoning (that small autocorrelations yield bad forecasts and
that, conversely, large autocorrelations yield good forecasts) finds its
clear quantitative expression: Having fixed autocorrelations, the cor-
responding numerical value of the normalized standard deviation e(l)
(or lead time) can be found from Figures 4.1 through 4.11.
6. The results presented can also be helpful when there is little
data and when the confidence intervals of the estimates are wide.
In this case, the numerical interpretation obtained helps to establish
the agreement between the confidence interval of the autocorrelations
and the corresponding region of the forecast error (or lead time); that
is, some additional information can be obtained.
7. A comparative analysis of the accuracy characteristics of the
second-order ARMA processes reveals that, in doubtful cases (when
it is possible to fit any of the three considered models), the highest
theoretical accuracy (and the largest lead time) is provided by the
MA(2), then by the ARMA(1,1), and, finally, by the AR(2). How-
ever, the distinctions are not significant, and, because the AR(2)
has the largest domain of the permissible autocorrelations, the cor-
responding models have the most practical application.
8. The ARMA processes make it possible to approximate a wide
class of the observed atmospheric processes and to study their phys-
ical and statistical properties. The results of the theoretical analysis
of the above five types of processes [AR(1), AR(2), MA(1), MA(2),
ARMA( 1,1)] present (as is shown in Section 6.5) virtually exhaustive
information about their applicability to modeling the climate time
series of the historical records (or to describing them with the aid of
the corresponding first- and second-order differential equations).
9. Fitting the stochastic models is one of the approaches for
estimating the spectral and correlation functions. This question
190 4 Variability of ARMA Processes

is considered, for example, by Anderson (1971), Box and Jenkins


(1976), and Kashyap and Rao (1976), among others. Estimating the
spectrum (and its physical analysis) is a more in-depth and subtle
procedure than estimating the forecast. This is evident because a
significant increase in the order of the ARMA process does not con-
tribute significantly to improving the forecast accuracy (or increasing
the lead time). However, the same increase in the process order can
drastically change the spectrum shape, which, in turn, changes the
understanding of the physical nature of a process.
10. The identification process is the key element in fitting the
stochastic model to the observed time series. Some identification
procedures, recommended in special publications (Kashyap and Rao,
1976; Eikhoff, 1983), are based on statistical criteria that evaluate the
appropriateness of the presupposed model to the given time series.
However, the final evaluation must be based not only on statistical
properties but also on physical concepts. For instance, the identi-
fication of a model of climatological time series actually states the
character (tendency) of climate change. But theoretical assumptions
and conditions, which linear processes must satisfy, are mathematical
abstractions. No natural processes strictly obey these abstractions,
and the identification is never absolutely precise. The conventional-
ity and subjectivity of the model identification procedure result from
the fact that there cannot be absolute adequacy between the process
of nature and the stochastic (for example, ARMA) process. Such in-
adequacy predetermines the nonuniqueness of climate modeling and
the necessity of reasonable compromise, which must reconcile the
difference between the reality and its approximate image contained
in the fitted model. The above study also makes it possible to ob-
tain an understanding of the extent of the distinction in the forecast
accuracy, when the identification is not accurate.

4.8 Signal-Pius-White-Noise Type Processes


As will be shown shortly and in Chapter 6, a climatic time series can
be modeled by signal-plus-white-noise type processes. Such processes
allow an accurate theoretical interpretation in the framework of linear
parametrical processes (Box and Jenkins, 1976; Parzen, 1966). In
this case, the white noise is referred to as additive white noise.
The addition of a white noise to the general ARMA process does
not change the order of the autoregressive part and affects only the
moving average component. More accurately (Box and Jenkins,
1976), if the white noise is added to the ARMA(p, q) process, the
4-8 Signal-Plus- White-Noise Type Processes 191

result is the ARMA(p, Q) process, where Q is the maximum of p and


q. For example, if p < q, then Q = q and the order of the process
with additive white noise is the same as the order of the original one;
only the values of the parameters 0 are changed.
Consider an example (see Kashyap and Rao, 1976) of a first-order
ARM A process with additional white noise component. This exam-
ple has application as an appropriate model for historical records of
climate time series of temperature, pressure, precipitation, stream
flow, and so on.
Let

where

is the first-order autoregressive process (signal); Nt and at are the


white noise processes; and a^ cr|, <7jy, a% are the variances of pro-
cesses Zt, St, Nt, and at respectively.
Let zt be the ARMA(1,1) process

where parameter 0 is the solution of the equation

satisfying the condition \0\ < f. The variance d^ of the white noise
V,\sis

One of the informative approaches (with a clear physical interpreta-


tion) to the analysis of the signal-plus-white-noise process has been
presented by Parzen (1966). Pareen's results match the well-known
empirical procedure of "extrapolation to zero" of the autocorrelation
function estimates used for many years in meteorology to estimate
the noise level in observed data.
Let us briefly consider this simple approach. Let random process
Zt be the sum of the signal St and the white noise Nt (4.66), where
St is the AR(p) process. The signal ratio
192 J, Variability of ARM A Processes

is a measure of the signal level (naturally, 1 — a is the measure of


the noise level). The ratio a shows the proportion of the variance
of process zt generated by signal St. Notice here that the ratio sig-
nal/noise, used in radio engineering, is connected with the ratio a
by the relationship

It is easy to show [for autocorrelation functions PZ(T) and ps(r) of


processes zt and St] that

Having estimated pz('r) and a, one can use (4.73) to compute


autocorrelation function />s(r) of the signal. For estimating ratio a,
let us notice that St is the All(p) process, and its parameters </?,• must
satisfy the Yule-Walker system

The autocorrelations PS(T) are unknown because the signal St is not


observed; only the process Zt is observed. However, from (4.73) it
follows that

where f>z(j} can be estimated using the observations of zt- Adding to


the Yule-Walker system (4.74) the equation for r = p-\-1 and replac-
ing PS(T) by (4.75), a system of p+ I algebraic equations (relative to
j» + 1 unknown parameters y?i, < p 2 , . . . , (fp, and a) is obtained. These
equations are nonlinear, but the analytical expressions for a can be
found for small p.
If signal St is the AR(1) (4.67), parameters ip and a are estimated
from the equations

Taking into account (4.75) yields

which gives
4-8 Signal-Plus-White-Noise Type Processes 193

Based on this formula, the least squares estimator for the ratio a can
be found using the first m + 1 estimates rT of autocorrelations pz(r\
Tills estimator is

where the appropriate value of m depends on the particular time


series considered. With the aid of 2, it is possible to estimate auto-
correlation function ps(i~) by (4.75).
The absolute value of autocorrelation PS(T) is greater than the
corresponding value of PZ(T} (because a < 1), which shows the dis-
tinction in the potential predictability of signal St as compared with
observed processes zt.
If the signal is the AR(2) process, parameter a can be found by
solving the following equation

where from the two values

it is necessary to choose only the one that is in the interval (0,1).


The problem of finding a is complicated for the process with the
signal that is presented by the autoregressive process of order greater
than two.
Estimates of a will be systematically provided in this section and
in Chapters 6 with climatological applications. In all these cases,
we assume that St is the AR(1) process and that parameter a is
estimated by (4.78).

4.8.1 Examples: Signal Ratio of Climate Time Series


Monitoring the modern climate requires evaluation of the potential
statistical predictability of the historical records of different meteo-
rological observations.
First we consider the longest temperature record in central Eng-
land (the 315 years from 1659 to 1973), published by Manley (1974).
Because the first 40 years of observations of this time series are not
accurate enough, we use only the most accurate part of this record-—
the period from 1700 to 1973—for the evaluation of ratio a.
194 4 Variability of ARM A Processes

Another source of temperature data, for which the ratio a is esti-


mated, is the data base collected in the World Data Center, Obninsk,
Russia (Gruza, 1987). The Northern Hemisphere temperature data
from these archives for the period from 1891 to 1985 are used to
obtain the zonal and annual mean time series cited in Table 4.1.
Table 4.1 provides estimates of the signal ratio a for some cli-
mate time series. The results reveal that for the temperature data
the smallest estimates of the first autocorrelations correspond to the
time series of the stations (point gauges) [pz(l) = 0.18 for the an-
nual and pz(l) = 0.3 for the monthly data of the central England
temperature]. According to the results of Section 4.7, it does not
make sense to use the ARMA model to forecast the time series with
such small values of autocorrelations independently of the order and
type of the fitted model. The value (0.37) of the first autocorrelation
of the monthly temperature time series, obtained by averaging data
within the ten-degree width latitude band (50°-60°) which includes
the entire territory of England, can provide a low-accuracy forecast
for one step ahead, according to the results of Section 4.7. Auto-
correlation for the annual data, averaged over the entire Northern
Hemisphere, is about 0.8. In fact, it makes sense to forecast such
data for up to three steps ahead.
For the first three cases in 'Fable 4.1, the model of the signal [the
AR(1) process] plus white noise appears to match perfectly with the
data. The lowest level (a = 0.25) of the climate signal corresponds
to the annual mean temperature time series of the separate station,
and the highest level (a = 0.60) to the monthly data averaged within
the 50°-60° latitude band.
For the temperature time series of data averaged over the entire
Northern Hemisphere, the « estimate [obtained by (4.78)] is greater
than one, which means that this simple model [AR(1) process plus
white noise] is not an appropriate approximation. The procedure of
averaging all hemisphere observations that have sufficiently different
level of autocorrelations in different regions of the Earth drastically
changed the statistical structure of the resulting time series, com-
pared with the original ones. Averaging of temperature observations
with very diverse temporal statistical properties (for example, almost
white noise in the polar regions and high autocorrelated time series
near the equator) has led to the creation of a new time series with a
new statistical nature. However, such averaging is widely applied in
climatology without questioning its statistical legitimacy.
Spatial averaging reduces the noise component and increases the
ratio a of the climate signal; therefore, it is one of the ways of detect-
ing a climate signal when averaged time series have approximately
identical statistical structures.
4-8 Signal-Plus-White-Noise Type Processes 195

Table 4.1: Estimates of signal ratio a for different time series.

Time Series Mi) M2) a

The central England monthly mean


temperature time series (1.891 1985) 0.30 0.16 0.56
The zonal (50°-60° latitude band of the
Northern Hemisphere) monthly mean
temperature time series (1891-1985) 0.37 0.23 0.60
The central England annual mean tem-
perature time series (1700-1973) 0.18 0.13 0.25
The Northern Hemisphere annual mean
temperature time series (1891-1985) 0.80 0.57 1.12
The Brazos River (Richmond gauge)
monthly mean streamflow time
series (1900-1984) 0.48 0.28 0.82
The Brazos River (Richmond gauge)
annual mean streamflow time
series (1900-1984) 0.08 0.06 0.11
Annual solar radiation budget in the
USSR cities:
Karadag (1934-1975) 0.20 0.10 0.40
Tashkent (1937-1975) 0.23 0.19 0.30
Tbilisi (1937 1975) 0.23 0.16 0.35
Irkytsk (1938-1975) 0.25 0.19 0.34
Vladivostok (19391975) 0.19 0.15 0.23
196 4 Variability of ARM A Processes

The next cases in Table 4.1 concern the normalized anomalies


of the monthly mean streamflow time series (Richmond gauge, the
mouth of the Brazos River), matching the above signal-plus-white-
noise model with a = 0.82. Analogous results obtained for the other
18 gages of the Brazos River basin gave the a estimates, which vary
from 0.29 to 0.82. Corresponding annual streamflow time series are
very close to the white noise samples.
The results in Table 4.1 for the annual solar radiation budget
(the most accurate time series of several cities of the former USSR
from the Pivovarova archives, Main Geophysical Observatory, St. Pe-
tersburg) also very closely follow this simple signal-plus-white-noise
model with the a estimates in the relatively narrow interval of values
from 0.23 to 0.40 (see also Polyak et al., 1979).
The fact that the variance proportion of the gauge annual time
series corresponding to the signal is very small explains the difficul-
ties connected with description and prediction. However, the results
obtained provide evidence that such a signal does exist and can be
considered as a long-period variation with a very small amplitude
observed only on a short-term interval of instrumental observations.
Climate time series models with an additive white noise com-
ponent clearly reveal the existence of the non predictable aspect of
climate variability.
Consider more general results for the surface monthly mean tem-
peratures obtained through the temperature time series mentioned
above for the observational period from 1891 to 1985. The observa-
tions were spatially averaged over various geographical and political
regions of the Northern Hemisphere, as cited in the first column of
Table 4.2.
Before estimating the simple statistical characteristics, the sea-
sonal cycle in the means and variances of the data was removed, and
the time series of the normalized values were analyzed. It is assumed
that the climatic signal can be presented by a first-order autoregres-
sive model; so, after estimating p z ( 1 ) and /->z(2), the signal ratio a
was computed by formula (4.78). After analyzing the estimates given
in Table 4.2, the following comments can be made:
1. The first autocorrelations of the spatially averaged time series
are greater than the corresponding autocorrelations of the point
gauges (see, for example, Polyak, 1975, 1979), which reveals the
importance of spatial averaging procedure to detect the climate
signal. The signal, ratio for the time series of averages is also, as
a rule, higher than that for the time series of the point gauges
(see Polyak, 1975, 1979), for which the signal is responsible for
not more than 30% of their variability.
4.8 Signal-Plus-White-Noise Type Processes 197

Table 4.2: Signal ratio and linear trend of the surface air tempera-
ture monthly mean time series averaged over different regions of the
Earth. Observational interval 1891 to 1985.
01
Regions of averaging P.(l) P. (2) or 1o-4oc t
per month
5-degree-width NH latitude
bands with the center in:
87.5° 0.28 0.11 0.71 9 3.5
82.5° 0.31 0.13 0.73 7 3.1
77.5° 0.37 0.16 0.85 4 1.8
72.5° 0.38 0.20 0.70 7 5.1
67.5° 0.35 0.22 0.57 5 3.1
62.5° 0.31 0.20 0.46 4 2.9
57.5° 0.27 0.18 0.43 1 1.0
52.5° 0.33 0.16 0.65 2 2.4
47.5° 0.34 0.16 0.73 3 3.8
42.5° 0.34 0.17 0.66 3 3.7
37.5° 0.37 0.25 0.56 0 0.3
32.5° 0.38 0.30 0.47 1 1.3
27.5° 0.40 0.30 0.62 3 6.3
22.5° 0.51 0.39 0.67 3 9.1
17.5° 0.61 0.46 0.80 3 8.2
12.5° 0.64 0.53 0.30 3 7.1
87.5°-67.5° NH 0.38 0.20 0.72 5 3.1
67.5°-37.5° NH 0.40 0.24 0.64 2 3.0
37.5°-27.5° NH 0.38 00.29 0.49 2 3.8
27.5°-12.5° NH 0.63 0.50 0.79 3 8.1
82.5°-12.5° NH 0.53 0.43 0.67 3 6.3
Eurasia 0.32 0.13 0.80 2 1.9
Europe 0.30 0.15 0.61 2 1.2
Asia 0.20 0.10 0.70 2 1.8
North America 0.22 0.10 0.35 4 3.1
North Africa 0.36 0.18 0.71 1 1.3
Atlantic Ocean
67.5°- 2.5° NH 0.67 0.59 0.76 4 7.8
67.5°~37.5° NH 0.65 0.58 0.71 6 7.9
37.5°-12.5° NH 0.53 0.41 0.68 2 5.3
Pacific Ocean
57.5°- 2.5° NH 0.63 0.55 0.71 7 12.1
57.5°-37.5° NH 0.49 0.40 0.59 7 8.9
3r.5°-12.5° NH 0.63 0.54 0.74 7 12.7
USA 0.11 0.06 0.18 4 2.4
Canada 0.19 0.09 0.38 4 2.1
Russia 0.30 0.13 0.68 3 2.0
198 4 Variability of ARM A Processes

2. The greater the region of averaging (i.e., the greater the num-
ber of point gauges), the greater the corresponding autocorre-
lations.

3. The northward decreasing latitudinal trend of the first autocor-


relations has a noticeable maximum in the equatorial region.
For the high and middle latitudes (greater than 30°) of the
Northern Hemisphere, the first autocorrelation of the temper-
ature anomalies is less than 0.4. Therefore, according to the
results of Section 4.7, the forecast of the corresponding time
series makes sense for no more than one step ahead, and its ac-
curacy will be very low. Indeed, the forecasting of the monthly
or annual mean temperature time series by ARMA models for
one to two steps ahead makes sense only for the tropical re-
gions, where the first autocorrelations are greater than 0.5-0.6.
The forecast accuracy of the temperature time series is higher
for the equatorial regions than for the middle and northern
latitudes.

4. Although the signal ratio value a in Table 4.2 varies within wide
limits (from 18% to 85% of the variance), the appropriateness
of the signal-plus-white-noise model in all the cases considered
is unquestionable.

5. Finally, let us note the importance of the methodological study


of the forecast accuracy of the different types of the ARMA
models (see Sections 4.1 through 4.7). This study gave us the
possibility of obtaining a complete representation of the sta-
tistical predictability of different climate time series without
fitting any model.

Together with the signal ratio a, Table 4.2 gives the estimates of
the linear trend parameter fl\ and its ^-statistics values. Almost all
estimates of j3\ are positive, and their range of variation is from 0
to about (10~ 3 )°C per month. In most cases, these estimates are
statistically significant at any standard level of significance.

4.9 Process With Stationary Increments


Consider the nonstationary random process y t , the finite differences
4-9 Process With Stationary Increments 199

of which are the stationary ARMA(p, q) process. Under some condi-


tions, the process zt can be presented as

Replacing the finite differences by their definition (2.71-2.73) we ob-


tain a difference presentation of process yt as

The variance of the forecast error of yt for l steps ahead (see Box and
Jenkins, 1976) is determined by the formula

where

Therefore, in this case, the fitting of the linear model can be


performed for the finite differences (4.80).
The simplest nonstationary process is the first-order autoregres-
sive process

of the first finite differences. It can be denoted as

or

where <j>\ = 1 + p, and ^2 = —p- Substituting these parameter


values in (4.83) and (4.84) enables us to compute t> 2 (/) for any l.
Following the denotation described in Section 4.1, let us consider the
normalized standard deviation
200 4 Variability of ARM A Processes

Figure 4.13: Dependence of the normalized standard deviation of the forecast


error on the lead time I and on the first autocorrelation p of the AR(1) process
of the first finite differences Vy*.

Figure 4.13 shows the dependence of e(l) on / for different values of


the autocorrelation p of the first differences Vjfe.
The normalized standard deviation (4.88) is less than 1 only for
1 = 1. If / > 1 and the lead time increases, the value of standard
deviation s increases infinitely (whereas the forecast value approaches
a straight line), which means that the accuracy of the nonstationary
forecast (for / > 1) is lower than the accuracy of the observations.
An infinite growth of the standard deviation e of the forecast error
takes place for any ARMA process of the finite differences that is
distinct from the stationary process, for which e(l) < 1.
Evaluating the goodness of fit of a nonstationary model and its
forecast accuracy for several steps ahead, one must be guided by the
physics of the observed process rather than by statistical reasoning.
The normalized standard deviation e(l) of the forecast error loses its
comparative (with 1) value, and the forecast of the process with sta-
tionary finite differences asymptotically approaches (with increasing
lead time /) one of the branches of the degree v polynomial (straight
line if v = 1).
Therefore, the closeness of the expected value of nonstationary
process to the degree v polynomial is the decisive factor that deter-
mines the forecast accuracy. If the absolute values of the autocorre-
lations of the finite differences are small (for example, less than
4-10 Modeling the Surface Air Temperature 201

0.3), the forecast for several steps ahead will not be distinct from the
corresponding value of the polynomial.

4.10 Modeling the Five-Year Mean Surface


Air Temperature
For many years, the time series of the annual mean surface air tem-
perature, obtained by averaging gauge observations over the entire
Northern Hemisphere has been studied as an indicator of climate
change. Despite the doubtful statistical legitimacy of such averaging
(over the observations with diverse spatial and temporal statistical
structure), it is interesting to analyze the resulting time series, to
build simple stochastic models, and to estimate the forecast accu-
racy. The objective of this study is to illustrate the dependence of
the forecast scenarios on the assumption about stationarity or non-
stationarity of the time series.
The time series of temperature anomalies under study was ob-
tained with the aid of the Northern Hemisphere temperature data
base of the World Data Center at Obninsk, Russia. This time series
has a small linear trend (approximately 0.005°C per year), but it is
statistically significant at the 95% level. To make the result more
descriptive, the temporal scale of averaging was increased, and mod-
els of the five-year means of the Northern Hemisphere surface air
temperature time series were built.

4.10.1 Stationary Model


The estimates of the first and second autocorrelations of this time
series are 0.8 and 0.57, respectively; the standard deviation is about
0.2 C°; the standard deviation of the forecast error by AR(1.) is 0.11;
the predictability limit of this model for e(l) = 0.8 is equal to about
three steps (see Figure 4.2); and the model is zt — 0.8^<_i + <i(.
The results, presented in Figure 4.14, show that, if the lead time
increases, the forecast converges to zero that was a priori clear for
the stationary model. By looking at the picture (Figure 4.14) one cart
see some similarity in the forecast behavior with the cooling period
observations from 1950 through 1970.
Of course, such a stationary model cannot be used for the mod-
eling of any trend. In order to include any deterministic component
in the description of climate, one must fit a nonstationary model.
202 4 Variability of ARM A Processes

Figure 4.14: The Northern Hemisphere five-year mean surface air temperature
anomalies (1); forecasts by the AR(1) model: for one step (five years) ahead (2),
for 2 to 6 steps ahead (3); 70% confidence intervals (4); standard deviation of the
forecast error (5).

4.10.2 Nonstationary Model


The principal motivation for fitting a nonstationary model to the
above temperature time series of the Northern Hemisphere is that its
least squares straight line /?o + fl\t approximation has a statistically
significant estimate of the parameter j3\ (fi\ — 0.023C0 per five years;
the Student's statistic t of this estimate is about 2; /30 = 0 because
we process the anomalies and the origin of the coordinate system was
chosen in the center of the observational interval).
The ARM A models were fitted to the first finite differences of the
above time series, and the value 0.023 of the /3\ estimate was taken
as an estimate of the mean (/*) of these differences.
The best was the ARMA(1,1) model,

where the autoregressive parameter (y?) estimate is 0.056. The mov-


ing average parameter (8) estimate is 0.13.
The obtained estimates provide the expression for the model,
4-10 Modeling the Surface Air Temperature 203

Figure 4.15: The Northern Hemisphere five-year mean surface air temperature
anomalies (1); forecasts by the ARMA(1,1) model of the first finite differences: for
one step (five years) ahead (2); for 2 to 4 steps ahead (3); and linear approximation
(4).

the standard deviation of the forecast error of which is 0.1°C. The


forecasts by this model and their 70% confidence interval are given
in Figure 4.15.
The first autocorrelation estimate of the first finite differences
time series is small (0.18), and for this reason the forecast, even for
one step ahead, practically leads to the extrapolation of the straight
line 0.023i. In other words, the forecast shows that the mean tem-
perature increases approximately 0.5°C per 100 years and gives zero
value for the fluctuation with respect to this mean.
Comparing stationary (Figure 4.14) and nonstationary (Figure
4.15) models of the same time series, one can see that the results of
forecasting are very different, although their standard deviations of
forecast error are approximately the same (0.11 and 0.10).
204 4 Variability of ARMA Processes

It is interesting (methodologically) to attempt to increase the dif-


ference order and to build a model of the second finite differences.
In this case, the differences V 2 yt are proved to be statistically inde-
pendent, and the forecast of the corresponding model coincides with
the extrapolation of the corresponding second-degree polynomial

The branches of this parabola are directed downward, unlike the


forecasts of the previous model and the modern representation of
global warming.
The simple stochastic models clearly reveal the conditional char-
acter of the forecasts, depending on the fitted model type and on
the assumption of stationarity or nonstationarity. Fitting stationary
and nonstationary models to the same climate time series makes it
possible to obtain different climatic scenarios.
It is also important to notice that, since the period of instrumen-
tal observations is small, it is unrealistic to formulate an identification
problem for the stochastic modeling of this climate temperature time
series. In other words, it is unrealistic to determine the statistical
law of climate variation with the aid of a time series with only 21
observations. For this reason, fitting low-order models is a necessary
step in the right direction. .Indeed, such investigations help to gar-
ner the experience that is necessary for building multivariate climate
models.

4.11 Nonstationary and Nonlinear Models


As was shown, fitting a stochastic model to finite differences under-
mines the polynomial approximation of the trend and the modeling
of the residuals.
The following two-step procedure provides a more direct empiri-
cal approach to the modeling of a long time series with the nonsta-
tionary expected value (trend). For the first step, a time series is
approximated by the appropriate function, which is linear relative to
the estimated parameters. For the second step, the ARMA model is
fitted to the time series of the residuals.
in contrast to the stationary case, the variance of the forecast
error of the nonstationary linear model increases infinitely with in-
creasing lead time. For the polynomial trend, this variance is propor-
tional to the value of ( k / 2 f I)'2", where l is the lead time and v is the
polynomial degree or the finite differences order (see Section 1.3). If
4-10 Nonstationary and Nonlinear Models 205

/ > 1, the variance of the forecast error of the nonstationary model


is greater than the variance of the observations, and, of course, it is
greater than the corresponding variance of the stationary model.
If the statistical structure of the time series slowly varies in time
and the observational interval is sufficiently long, one can use another
empirical approach analogous to the smoothing moving average pro-
cedure. The interval of the observations can be divided into subin-
tervals, possibly overlapping, for each of which the linear stationary
model can be fitted, After that, one can study the temporal variation
of the parameters' estimates and, if suitable, can approximate them
by means of an appropriate function. In this simple method, mon-
itoring a random process can be replaced by monitoring the model
parameters.
Nonlinear modeling is, potentially, the most promising direc-
tion of climatology development. Fundamental concepts of nonlinear
modeling are presented by Tong (1990). According to Tong, the
general case of the nonlinear model can be denoted by

where et is the white noise.


The model, which has been intensively studied, is the threshold
autoregressive model, the simplest form of which is

where a and b are the parameters.


Such nonlinearity can be associated with discontinuities of the
observed process; for example, the climate time series of precipitation
in the form of water and snow, solar radiation with and without
cloudiness, and so on. Such discontinuities must be reflected in the
structure of the fitted model. The intake of the threshold in the
model is caused by the physics of the process under study and, in
particular, by the possibilities of catastrophes (Tong, 1990).
Another type of nonlinear model is the class of bilinear models,
the simplest example of which is

where a, 6, and c are the parameters; et is independent of Xt white


noise with a zero mean and a finite variance; and j is an integer.
206 4 Variability of ARM A Processes

The distinction of (4.92) from the ARMA model lies in the existence
of the bilinear term et-jXt-\. The important feature of the bilin-
ear model is that the time series approximated by this model can
have time-variable statistical structure, which is the case for some
hydrometeorological processes.
Tong refers to (4.91) and (4.92) as nonlinear models of the first
generation. A more general type of nonlinear model, the threshold
model of the second order, is

Fitting the nonlinear model makes it possible to decrease the stan-


dard deviation of the forecast error compared with the best linear
models. For forecasting more than one step ahead, the situation is
completely different; that, is, the standard deviation of the forecast
error becomes greater than the corresponding standard deviation of
the linear model.
The necessity of moving toward considering nonlinear and non-
stationary models has become especially clear in the process of iden-
tifying and agreeing on the structures of the stochastic and physical
(hydrodynamic) models because, as a rule, the latter are nonlinear
and nonstationary.
To conclude, the key points of the nonlinear modeling are as
follows.
1. If the lead time increases, the nonstationary model forecast
approaches the function (the polynomial), which approximates
the time series trend.
2. If / > 1, the variance of the forecast error of the nonstation-
ary and nonlinear models is greater than the corresponding
variance of the observations and, naturally, is greater than the
analogous variance of the forecast error of the stationary model.
3. Nonstationary and nonlinear models can describe a real cli-
mate change on the significant time intervals, whereas station-
ary models can characterize only the climate stability and in-
variability.
4. The fitting of stationary models on several subintervals of the
entire observational interval can represent temporal variation
of the model parameters.
4-10 Nonstationary and Nonlinear Models 207

5. Meteorological processes in the real-time scale are not station-


ary in principle. But in many cases, to meet the assumption
of stationarity, the corresponding observations are transformed
so that the resulting quantities do satisfy this condition. The
standard transformations are: filtering out the annual cycle or
the latitude trend, large-scale spatial and temporal averaging,
and computing the principle components. Averaging addition-
ally lowers the noise component and reduces the volume of
information that must be analyzed.
5
Multivariate AR
Processes

This chapter is devoted to different computational aspects of multi-


variate modeling. An algorithm for fitting such models with sequen-
tial increasing of the order is given. Several examples of approxima-
tion of multiple climatological time series by first-order AR models
are considered. It is emphasized that such modeling can be used not
only for forecasting, but also for the analysis of the parameter matrix
as a matrix of the interactions and feedbacks of the observed pro-
cesses. Some problems in identifying stochastic climate models are
also discussed. It is shown that formulation of climatology problems
within the strict frameworks of fundamental theory will facilitate
natural progress along with the development of these methods.

5.1 Fundamental Multivariate AR Processes


Let us consider q discrete random processes x\f<xiti • • - i x q t - Vector

is referred to as q-variate random process. We assume that this


process is stationary and that E(XH) — 0; i ~ 1,2,.. . ,g.

208
5.1 Fundamental Multivariate AR Processes 209

The auto- and cross-covariance functions of processes xn can be


presented in the form

where T means transpose; v,-j(r) = ~E(xnXjt+T) is the cross- (or auto-


if i = j ) covariance function of processes x,-/ and Xjt. The sequence
of the matrices V(r) (T = 0,1,2,...) is referred to as a (/-variate
covariance function or as the matrix covariance function of random
process X t .
The g-variate covariance function satisfies the equation

caused by the property of the stationarity

If T ^ 0, the matrix V(r) is not symmetric; however, the (/-variate


covariance function V(r) can be considered (and estimated) only for
T > 0 because of (5.4) [UJJ(T) is the continuation of w 8 -j(r) to the
domain of negative T].
With the aid of the g-variate covariance function, one can de-
fine the (q-variate correlation function (in the same way as for the
univariate case), the elements of which are the correlations

Let

be the q-variate process, which is defined in the following way:


1. Each rijj is the univariate white noise.
2. Processes n-tl can be statistically dependent only at the same
moment of time; for the nonzero lags, the covariances are ze-
roes .
210 5 Multivariate All Processes

The matrix covariance function S(r) of process Nj is

where O g x ? is the matrix of g X g size, all the elements of which


are zeroes. This definition generalizes the notion of the white noise
process for the multivariate case. The process N( with the matrix
covariance function (5.7) is referred to as the g-variate white noise.
If a sequence of square matrices oti

(size q X 17; aijk are the numbers) is such that the value of process Xf
in any moment of time is determined by the linear combination of
its p previous values and the value of g-variate white noise N<; that

then X( is referred to as the q-variate autoregressive process of order


p. Because the sequence of the matrices on consists of p x g2 elements,
and the white noise N< is determined by the matrix S consisting of
r/2 elements, the g-variate autoregressive process of order p is entirely
determined by (p + 1) X q2 parameters. Their estimation presents
the principal problem of fitting the multivariate Alt model to the
multiple time series of the observed system.
The fundamental relationship, which is the basis for such an es-
timation, can be obtained by multipling the left and right parts of
equation (5.9) by X(T|_1_.,r and by taking the expectation:

Then if r > 0 we have


5.1 Fundamental Multivariate AR Processes 211

that is, the matrix covariance function satisfies the matrix difference
equation (5.11).
Setting T = 1,2,.. .,p in (5.11) gives the multivariate analog of
the Yule-Walker equations (see Section 4.1)

This is a system of linear equations with p X q2 unknown scalar


parameters. By replacing V(r) by their estimates and solving the
system relative to the matrices of cxi, it is possible to obtain the
parameter estimates.
Let us introduce the following notations:

where UT and ex are block matrices of size q X qp; Wp is the block


matrix of size qp X qp.
Using these notations, the Yule-Walker system is

or

Equation (5.9) can be written in a more concise form by setting

and
212 5 Mullivariate AR Processes

The matrix Wp is the covariance matrix of vector XQ.


These notations simplify the algorithmic structure and make it
possible to use it in a recursive way when designing and developing
computer routines.
The equation for the matrix X is obtained from (5.10), if T — 0:

Therefore,

This equation is a multivariate analog of (4.12) for the variance of


the white noise.
The elements on the main diagonal of matrix X are the variances
of the components of vector Nt+1 . The estimation of X is necessary
for constructing confidence intervals of the forecasts. If the model is
fitted to the multiple time series, the estimated matrix S is referred
to as the covariance matrix of the forecast error for one step ahead.

5.2 Multivariate AR(1) Process


The multivariate autorcgressive process of the first order can be con-
sidered as the generalization of the concept of the scalar Markov
random process.
The first-order multivariate Ail process is

where

is the parameter matrix, which must be estimated when a model is


fitted to data.
The difference equation for the matrix covariance function is
5.2 Multivariate AR(1) Process 213

From (5.24) it follows that

If r = 1, we get the Yule-Walker system

which is the basis for the estimation of a:

The estimation of the elements of matrix S is based on the formula

The expression for the variances {S}n can be used to determine


the domain of the permissible correlations of the multivariate AR(1)
process. Actually, the inequalities

must be satisfied because the process is stationary.


The forecast for / steps ahead is

where Nt+1 is the (q-variate white noise. The covariance matrix E; of


Nt+1, is

or

where

and EI = S.
If the forecasts are estimated step by step,

then it is suitable io compute matrices X, by means of the recursive


equations

and
214 5 Multivariate AR Processes

Equations (5,31) and (5.35) are the multivariate analogs of the cor-
responding univariate relationships (4.12) and (4.20) for the Markov
random process.
Estimating the elements of the matrix covariance function V(r)
for zero and the first lags is the first step in fitting the model.
To demonstrate the dependence of the forecast accuracy of the
multivariate A 11(1) process on the correlations and the number of
variates 17, consider the following illustrative example.

Example
Suppose that the elements of the matrix covariance function V(0)
and V(l) are determined by the formulas

Let us approximate the process with this covariance function using


the first-order multivariate autoregressive process. Replacing the
values of Vij in (5.27) and (5.28) with (5.37) and (5.38), we have

and

The first equation of (5.39) determines the domain of the permissible


values of the autocorrelations because
5.2 Multivariate AR(1) Process 215

In spite of the particular character of this example, the analytic solu-


tion enables us to give some numerical interpretation of the results.
It helps to understand the nature of the mutual stipulation of the
correlations p and r for the zero and first lags, the number of variates
q of the process, and the accuracy characteristics (e and 7) of the
corresponding forecast.
Let us study the behavior of standard deviation e (5.39) for dif-
ferent r,p, and q. If r — 0, then e=l, and the forecast coincides with
the mean value of the process. If the value of q is increased, then e
is monotonically decreased in the following way:

Therefore, the accuracy of the forecast is limited by the value 1 — r2/p.


In addition, beginning with some value of q, increasing the number
of the mutually analyzed time series cannot lead to an unlimited
reduction of e. In general, the inequality

determines the range of the normalized variance of the forecast error


independently of the number of variates of the considered process.
This inequality outlines the possible accuracy of the forecast, of the
process [with V(0) and V(l), given by (5.37) and (5.38)] and the
expediency of fitting the corresponding model.
For a detailed numerical description of the behavior of the nor-
malized standard deviation £ as a function of the correlations (p and
?•) and the number of variates (q), let us carry out the calculations by
means of (5.39) for some fixed grid of /?, ?•, and q values. The results
are presented in Figures 5.1 and 5.2.
An analysis of Figure 5.1 enables us to make the following com-
ments. First, the smaller the absolute value of the correlation p for
zero lag, the smaller the domain (/>, r) of the permissible correlations
(and the higher the forecast accuracy). If q is increased, the largest
growth of accuracy (decreasing of g) occurs for small p (p < 0.05).
Therefore, for this example, the statistical independence of the time
series Xn at zero lag (and the maximum possible temporal correlation
r) is ideal. For large /5, the increase of the accuracy (with increas-
ing q) occurs more slowly than for small p. Actually, if p is greater
than 0.8, the standard deviation of the forecast error does not differ
markedly from the corresponding standard deviation of the forecast
of the univariate processes xn (5.1) with the same value of r. This
Figure 5.1: Normalized variance e2 (as a function of g, r, and p) of the forecast
error of the multivariate AR(1) process determined by (5.37) and (5.38).

216
5.2 Multivariate AR(1) Process 217

Table 5.1: Correspondence between number of variates (q) and cor-


relation (p) for zero lag.

[7"Tb.O-0.2 I 0.2-0.5 I 0.5-0.7 I P/MLO


\q^\ 8-10 I 6-8 I "1T6
qs I 8-10' 3-6 \| 2-3

fact is easy to understand; for, if p is close to 1, a statistical depen-


dence of the variates is close to a linear (deterministic) dependence,
and the cross-statistical analysis cannot provide a significant amount
of new information, The growth of q leads to the decrease of e, which
asymptotically approaches the fixed (limit) value. Beginning with a
certain value of <?s, some kind of saturation is taking place, and fur-
ther increasing the number of variates does not affect the forecast
accuracy. Approximate values of such qs for corresponding p are
presented in Table 5.1.
Second, an evaluation of the possibility of fitting the multivari-
ate model and forecasting reveals that, exactly as for the univariate
case, the principal meaning is not simply the existence of the tem-
poral statistical dependence (r / 0), but the quantitative level of
this dependence. For small r (r < 0.2), the standard deviation e
of the forecast error with any q and p is greater than 0.8; that is,
it is equal to 80% (and more) of the standard deviation of the pro-
cesses. The corresponding forecast does not significantly differ from
the mean value of the process, and its estimation is of no practical
value, Therefore, the interval (0.0, 0.2) of the variation of r is outside
the bounds of practical consideration.
If 0.2 < ?* < 0.3 (the range of the values of the first autocor-
relations for the monthly and annual mean time series of air tem-
perature, radiation, streamflow, and so on) and q is increased, the
standard deviation £ is gradually decreased to the values of 0.7-0.8,
only if p < 0.2. Therefore, the possible accuracy of the multivariate
models of the climate time series of point gauges is limited by these
values (0.7-0.8); that is, the above calculations exhaustively charac-
terize the possible accuracy of the statistical description of the local
climate.
If p > 0.3, fitting the multivariate model with r < 0.3 is of no
practical value because the standard deviation of the forecast error
exceeds 90% of the standard deviation of the process (independent
of the q value).
218 5 Multivariate AR Processes

Analyzing Figure 5.1, it is possible to choose an appropriate


region in the space of (r,p,q) In which the forecast error e is so
small that fitting the model makes sense. If the model with £ < 0.8
can be considered acceptable, then, as Figure 5.1 shows, the cross-
correlation of the mutually analyzed time series must be greater than
0.3-0.4 at lag one.
From the above reasoning, it follows that one of the ways of build-
ing multivariate climate models is to prepare corresponding samples
of spatially and temporally averaged time series (of modeling to-
gether climate characteristics), the cross-correlations of which satisfy
the above requirements. More generally, it makes clear the necessity
of developing different filtering methodologies. From application of
such filters, the resulting time series of the means must be analyzed
together in order to evaluate the possibility of building their multi-
variate model. If for such time series the estimated cross-correlations
for the non-zero lags are greater than 0.3-0.4, then building the mul-
tivariate climate model makes sense. Of course, in this case it is
necessary to conduct a theoretical study of the influence of the spa-
tial averaging (transformation) procedure on the cross-correlation
function of the samples of averages, analogous to those studies that
were carried out in Section 2.4. The calculations and algebraic rear-
rangements in this case arc more complicated than those presented
in Section 2.4, and the interpretation of the results is not so obvious.
The character of the statistical dependence 7 of the forecast errors
on parameters p,r,q is demonstrated by Figure 5.2.
This example helps to represent some of the requirements of the
statistical structure of the multiple time series, the multivariate cli-
mate model of which makes practical sense. In the general case of the
first-order multivariate AR process, the analogous prior theoretical
analysis is more complicated. However, we can notice that signifi-
cantly increasing the number of the model's variates is necessary only
for very accurate investigations. Having a small number of modeling
variates, we do not lose much in terms of the forecast's accuracy; but,
of course, we certainly can miss some of the subtle spectral features
of the analyzed processes.

5.3 Algorithm for Multivariate Model


The above consideration of the theory of the multivariate autoregres-
sive processes represents the standard approach because it directly
follows the univariate case. When developing an algorithm for fit-
ting a multivariate ARM A model to the system of the time series,
Figure 5.2: Correlation 7 (as a function of q, r, and p) of the forecast errors of
the multivariate AR(1) process determined by (5,37) and (5.38).

219
220 5 Multivariate All Processes

the most efficient methodology is the recursive scheme, which is anal-


ogous to that in Section 4.2.
The forecast for I steps ahead can be presented by formula

where

1 is the identity matrix of size q X q, 0 is the matrix of size q X q with


all zero elements. The covariance matrix S/ of the forecast errors for
/ steps ahead is denoted as

where W is determined by (5.15). Formulas (5.43) and (5.46) yield


the following recursive procedure:

Programming this algorithm in an algorithmic language, which per-


mits the immediate use of the matrix expressions, is very simple.
Fitting the sequence of models to the given system of the time
series by consecutive increase in the order p leads to the necessity of
inverting the sequence of matrices Wp (p = 1,2,...).. If the model
order is increased by one (from p — 1 to p), then to obtain Wp the
blocks of the last row and the last column must be added to the ma-
trix W P _I. The inversion of Wp is significantly simplified by taking
into account the fact that WJ1, has already been obtained. The
5.3 Algorithm for Multivariate Model 221

most convenient algorithm for inverting matrix Wp is partitioning


(Berezin and Zhidkov, 1965). Using this method one can obtain the
recursive equation for Wp"1.
Let

where

The matrices W J^ and V 1 (0) have already been computed in the


previous steps.
The partitioning of matrix W-1, corresponding to the blocks of

According to the partitioning method (Berezin and Zhidkov, 1965),


the expressions for the separate blocks of matrix W"1 (5.53) are

where 1 is an identity matrix.


Thus, in each step, a matrix of qxq size must be inverted (instead
of matrix Wp of pq X qp size), thereby significantly saving time and
increasing accuracy, compared with the direct inversion of Wp.
Despite the efficiency of the algorithm considered, its representa-
tion in the form of a computer routine for arbitrary p and q requires
a significant volume of computer memory.
222 5 Multivariate AR Processes

The following comments are useful for software development and


application of the algorithm considered.
1. In a multivariate case, an application of the (4.43)-type cri-
terion leads to some difficulties because the scales (or the units) of
measurement of the mutually analyzed processes can significantly
differ in the numerical values. For example, eliminating the statis-
tically insignificant estimates of parameters «;jk according to (4.43)
can lead to very different results, depending on whether the model
is developed for the anomalies or for the standardized anomalies.
Therefore, the computational aspects, in addition to the statistical
nature of the data, can determine the final inferences. It is possible
that processing the normalized anomalies is the most rational way of
performing computations, but its theoretical foundation is not abso-
lutely clear. The choice of the appropriate scales and units can be
simplified significantly if there are physical relationships connecting
the mutually analyzed processes.
2. In the case of normalization, the matrix autocovariance func-
tion in the above formulas must be replaced by the corresponding
matrix autocorrelation function, and the estimation of parameters ot
and matrices 231 must be carried out for the normalized time series.
During the final stage of computations, to obtain the results in real
scales or units, the forecast estimates and their confidence intervals
must be multiplied by the corresponding estimates of the standard
deviations. The analysis of variances {IS/};; of the normalized ob-
servations makes it possible to evaluate the accuracy of the results
in exactly the same way as was done for the univariate processes
(comparing the {X;};2 with 1). The range of variations of the fore-
cast accuracy, which were found for the univariate processes, remain
approximately the same for the multivariate cases.
3. The variances of the components of the q-variate white noise
Nt+l are located along the main diagonal of matrix £;. Therefore,
the estimation of S; is necessary to construct a confidence interval
of the forecast for / steps ahead. Sl is determined by the cross-
statistical structure [the matrices V(r)] of the processes xa on the
two intervals of lags: r 6 [~(p — i),p — 1] and T G [l — p, l + p],
which, of course, can overlap. However, by taking into account the
relationship (5.11), it is clear that the forecast depends only on V(r)
for r = 0, l , 2 , . . . , p - 1.
4. The identification process is developed by fitting the models
with successively increasing order and by analyzing the following
quantities:
5.4 AR(2) Process 223

a. values analogous to (4.43), vKctijk, where K is the sample


size, for each parameter (otijk) estimate to determine its statistical
significance;
b. values analogous to (4.44) for the residuals of each equation
of the consecutive models of the p and p + 1 orders;
c. the proximity of the normalized variances, £? = {S;};,-/V;;(0),
to 1. The estimation of the forecast accuracy is carried out by con-
structing the confidence intervals of the forecasts with the aid of the
variances {£/};; of the forecast errors.
By making a decision on the basis of these results, one can identify
the model's order and the number of variates, which must be jointly
analyzed.
5. In climatology the general problems of identifying multivariate
stochastic climate models have not been formulated. Even the appro-
priate models of the first order are not developed. But fitting such
models and appraising their physical and statistical characteristics is
an interesting area of the developing field of statistical climatology.
Of course, it is impossible to describe the climate only by AR
processes. However, developing such models is a scientifically sound
step in the right direction, which enables us to make accurate statis-
tical inferences about the detectability of climate change. It seems
that in the near future the first- and second order AR multivariate
climate models will have the most practical value.

5.4 AR(2) Process


The q-variate autoregressive process of the second order

is determined by the parameter matrices oti and c*2 and by the co-
variance matrix Si of the white noise N^I. Since each of these
matrices consists of q x q numbers, the second-order process is to-
tally determined by the 3 X q2 parameters. Their estimation may be
derived by the solution of the Yule-Walker system, with preliminary
estimation of the elements of the matrix covariance function V(r),

relative to matrices «i and « 2 -


224 5 Multivariate AR Processes

Then, Sj can be computed by the formula

After estimating matrices «j and 03, the forecast is made using the
recursive formula

where the unknown quantities of X must be replaced by their esti-


mates.
The computation is provided by (5.47) to (5.50), which are trans-
formed into the following equations:

where

1 is the identity matrix of order q X q; 0 is the matrix of size q X q


consisting completely of zeroes,

and

The main calculation is the inversion of the matrix WV If the


first—order model has already been built, the matrix V(0) has been
inversed. Having V~ 1 (0), matrix W2 can be inverted by applying
the scheme given in the Section 5.3. Let us consider
5.5 Examples of Climate Models 225

where Cjj are matrices of order q X q. Then, as we have already


mentioned, the following equalities can be used:

where « is the matrix of the parameter estimates of the first-order


model.

5.5 Examples of Climate Models


The statistical methodologies considered enable us to analyze the
statistical properties and predictability of many kinds of hydrome-
teorological information. Naturally, to fit the stochastic models to
climate data, the most voluminous and reliable sources of long-period
climatological observations must be used. One such source is differ-
ent atmospheric circulation statistics, obtained by Oort (1983), based
on the radiosonde measurements of the atmosphere.

5.5.1 Modeling of Atmospheric Circulation Statistics


From Oort's data base, six mean monthly time series, averaged for
the entire Northern Hemisphere, were selected (tables with data are
given in Polyak, 1989). The list of statistics (in Oort's notations)
includes:
1. TASFC—the 1000 mb level air temperature.

2. TAMN -the air temperature, obtained by averaging the ob-


servations over 20 standard isobaric levels from 1000 mb to 50
mb.

3. T850—the 850 mb level air temperature.


4. TS—the sea surface temperature.
5. QSFC—the 1000 mb level specific humidity.

6. Q850—the 850 mb level specific humidity.


226 5 Multivariate All Processes

Table 5.2: Estimates of the statistical structure of the Oort's at-


mospheric circulation statistics averaged over the entire Northern
Hemisphere.

Characteristics TASFC TAMN T850 TS QFSC Q850

Standard deviation
of the time series 0.27 0.23 0.24 0.28 0.17 0.20
p(l month) 0.49 0.74 0.58 0.88 0.76 0.86
p('2 months) O.32 0.63 0.50 0.82 0.67 0.74
Lead time (month)
for e(l) -- 0.8 1 2 1 4 2 3
Signal ratio (4.78) 0.77 0.87 0.68 0.95 0.86 1.00
e(l) for
univariatc A 11(1) 0.87 0.67 0.81 0.47 0.65 0.51
e(l) for
six-variate AR(l) 0.86 0.62 0.76 0.46 0.64 0.49

The observational time span is 15 years, from May 1958 to April


1973, with 180 terms in each time series. The temperature unit is
°C and the specific humidity unit is g/kg.
The seasonal cycle was removed by subtracting the mean esti-
mates for the corresponding months from each term of the time series
and dividing the anomalies by the standard deviation estimates for
the corresponding months. Therefore, the normalized anomalies are
analyzed.
The objective of this study is to compare the univariatc and mul-
tivariate modeling of the same observations.

5.5.1.1 Univariate AR Models


As a first step, each time series was approximated by the first-order
univariate autoregressivo model. The normalized standard deviation
£(l) of the forecast error for l steps ahead by the AR(1) of time
series zt with variance d2 and zero expected value was determined
by formula (4.26). The computation of l as a function of e(l) and
/)(!) was done by formula (4.27).
The estimates of the basic statistical characteristics are collected
in Table 5.2. The autocorrelations at one- and two-month lags reveal
the forecasting value of each of the considered time series. As one
might expect, the maximum value of the autocorrelation [ p ( 1 ) =
5.5 Examples of Climate Models 227

0.88] corresponds to the time series of the ocean surface temperature.


It is much less obvious that time series of specific humidity Q850
[p(l) = 0.86] and QSFC [/>(1) = 0.76] are rarely inferior to TS.
The autocorrelation coefficients of the temperature time series are
smaller: p(l) = 0,74 for the TAMN; p(l) = 0.58 for the T850; and
p(l) •= 0.49 for the TASFC] i.e., the surface air temperature, often
used as an indicator of climate change, is not the only or even best
indicator of such change. The normalized standard deviations of the
forecast error for the sea temperature s(l) = 0.47 and for the specific
humidity Q850 e(l) = 0.51, which means that their forecast for one
month ahead is about two times less variable than the anomalies.
In rows 5 and 6 of Table 5.2, the estimates of the lead time / (4.27)
corresponding to the normalized standard deviations [£(/) = 0.8] are
presented. These values are the most informative characteristics of
the time series under study. They show that, when the requirements
of forecast accuracy are minimal [e(l) = 0.8], the ocean temperature
TS and the specific humidity Q850 can be forecast for about 3 to 4
months ahead. The air temperature observations can be forecast for
not more than two months ahead.
The estimates of the signal ratio, given in Table 5.2, show that
the climatic signal consists of about 65 to 100% of the variance of
the corresponding time series; consequently, the white noise forms
about 35 to 0% of that variance. The signal level is so high because
the scales of spatial averaging are very large.
Figure 5.3 (a) presents the forecast estimates of the considered
time series for six months ahead and the corresponding 70% confi-
dence intervals. The forecasts smoothly converge to zero if the lead
time increases. The standard deviations of the forecast errors con-
verge to the corresponding standard deviations of the time series.
The ocean temperature and specific humidity observations cars be
forecast most accurately.
Finally, we must notice that a more elaborative study, using zonal
means time series of each of the elements, revealed that the possi-
bility of modeling is a result of the high autocorrelations (and small
variances) of the observations in the tropical regions of the Northern
Hemisphere. For most of the above elements, the forecasts in the
middle or higher latitude bands make sense for no more than one
month ahead. Indeed, it is only for TS and Q850 that the forecasts
in the middle and high latitudes make sense for a two-month lead
time.
Figure 5.3: Forecasts of the mean monthly anomalies of the atmospheric circulation statistics by
the univariate (a) and six-variate (b) AR(1) models. 1. Estimated forecasts. 2. 70% confidence
intervals. 3. Standard deviations of the time series. 4. Standard deviations of the forecast errors.
5. The last observations of the time series.
5.5 Examples of Climate Models 229

5.5.1.2 Six-Variate AR Model


A need for a mutual statistical analysis of various meteorological data
sets was clear at all stages of climatology development. However, this
approach in the case of multivariate statistical modeling calls for a
sufficiently large number of time series, which are to be sufficiently
long. The number (180) of terms in each of the above time series
is too small even to attempt to fit higher order multivariate models.
Discussed below are the main results of the AR(1) model parameters
estimation. Let us consider a six-variate AR model of the same time
series that were studied in the previous subsection. The results of
estimation are presented in Tables 5.3 and 5.4.
The diagonal elements (autocorrelations) of matrix V(l) (Table
5.3) have already been discussed when we considered the univariate
models. It is interesting to notice that some cross-correlations for
the one-month lag are also sufficiently large (0.35 to 0.66). Thus, it
is appropriate to build the multivariate model.
Table 5.4 gives the parameter estimates for the first order AR
model. The main component, which forms fluctuations of each of
the discussed processes, is the history of the process itself (diagonal
elements an of matrix a exceed all other elements of the line in terms
of magnitude). This fact is a confirmation of the reality and physical
significance of the model.
Along with its own history, the past of other processes influ-
ences the magnitude of the anomalies of each of the elements. Thus,
TASFC(t+l) is determined directly by not only TASFC(t), but also
by specific humidity QSFC(t) and Q850(t). Since the latter, to some
extent, depend on TS(t), the TASFC(t+l) indirectly depends on the
ocean surface temperature. Of course, both direct and indirect in-
fluences of the history of some processes on others are very small,
which causes major difficulties in their estimation and tracing.
Actually, statistically significant estimates in Table 5.4 represent
a matrix of interactions and feedbacks of observed processes. Having
the analogous estimates for the time series simulated by a GCM, we
could compare them, making statistical inference about the structure
of such AR models and the closeness of the GCM to reality. The
forecast estimates by this six-variate AR model presented in Figure
5.3(b) smoothly decline toward their mean values (to zero) with an
increase of the lead time, but not as rapidly as in the case of the first
order univariate autoregressive models.
The normalized standard deviations of the one-step-ahead fore-
cast errors (last row of Table 5.2) show that the most accurate results
230 5 Multivariate All Processes

Table 5.3: Matrix correlation function for zero and first temporal
lags.

T TASFC TAMN T850 TS QSFC Q850

TASFC 1.00 0.28 0.24 -0.03 0.42 0.01


TAMN 1.00 0.69 0.38 0.4.5 0.47
0 T850 1.00 0.12 0.11 0.17
TS 1.00 0.38 0.67
QSFC 1.00 0.52
Q850 1.00

TASFC 0.49 0.10 0.04 -0.04 0.35 0.02


TAMN 0.13 0.74 0.52 0.37 0.36 0.39
1 T850 0.08 0.42 0.58 0.13 0.01 0.1.3
TS -0.03 0.40 0.14 0.88 0.37 0.66
QSFC 0.24 0.48 0.07 0.38 0.76 0.48
Q850 -0.08 0.48 0.20 0.65 0.44 0.86

were obtained for the sea surface temperature and the 850 mbar level
specific humidity.
The discussion above shows that the meaning of multivariate
models is not limited to forecasts. Consideration of the parameter
matrix a as a matrix of interactions and feedbacks can be a pow-
erful tool for studies of different climate time series, their mutual
influence, and mutual conditioning.

5.5.2 AR Models of Some Temperature Time Series


Further, the results of the approximation of different systems of
monthly mean anomalies of the Northern Hemisphere temperature
time series by the first-order multivariate All models will be briefly
discussed. We will try to find the main regularities of interdepen-
dences and feedbacks of temperature fluctuations in different regions
of the troposphere and their mutual influence.
5.5 Examples of Climate Models 231

Table 5.4: Parameter estimates (exceeding 70% significance level) of


six-variate model.

TASFCt TAMN t T850t TS, QSFC, Q850t

TASFQ+1 0.44 „.,...._ _ 0.21 -0.27


TAMNt+1i -0.15 0.71 — — 0.30 —
T850i+i -0.21 0.29 0.45 — — —
TSt+1 — — — 0.80 — 0.11
QSFCt+1 — — 0.08 0.68 —

Q850t+1i ~ 0.14 ~ 0.76

The first three models are built with the aid of radiosonde data
(Oort, 1983); the observational time span is 15 years (1958-1973),
with 180 terms in each time series.

5.5.2.1 AR Model of Zonal TAMN Temperature.


Consider Oort's data, which present the system of 9 time series of
the troposphere air temperature obtained by averaging data within
different 10 latitude zones and over the entire atmosphere from 1000
to 50 mb. For example, the time series corresponding to 80° was
obtained by data averaging between the 85° and 75° latitudes and
over the isobaric levels: 1000, 950, 900,..., 50. The second time series
(70°) was obtained by data averaging inside of 65° and 75° latitudes
and over the same levels; and so on.
Table 5.5 shows the estimates of the parameters of the corre-
sponding nine-variate model.
The main qualitative feature of this matrix is that most of the
estimates, located below the main diagonal, are not statistically sig-
nificant and can be considered as equal to zero. This means that
climate fluctuations of zonal temperature are formed, on average, by
the previous month's anomalies of this and southern latitudes. De-
pendence on the anomalies of the north regions (feedbacks or inverse
influences, north to south) are practically undetectable with these
time series.
Therefore, climate fluctuations at southern latitudes influence the
232 5 Multivariate AR Processes

Table 5.5: Parameter estimates (exceeding 95% significance level) of


nine-variate zonal TAMN temperature.
Lat.
zones 80° 70° 60° 50° 40° 30° 20° 10° 0° £

80° .13 .16 -.10 -.12 .71 -.15 .95


70° — .44 -.30 .33 -.23 .14 .63 -.20 .88
60° .10 .41 -.19 .10 .48 -.24 .86
50° — .53 -.20 .44 -.15 .81
— — —
40° .23 .51 -.19 .27 -.17 .84

30° .16 — .19 .70 .11 .77
— — — —
20° .51 .19 .62
— — _ —
10° -.22 .21 .62 .14 .S3


0° — — — .77 .57

fluctuations of the North or, more generally, temperature anomalies


of tropical regions are the main source of climate temperature anoma-
lies of all other latitudes. This feature is related to the meridional
atmospheric circulation.
In particular, as the coefficient estimates (of the first two equa-
tions) in Table 5.5 show, the climate fluctuations of the Arctic zone
are formed by the temperature anomalies of almost all other latitude
bands. According to the results of Polyak (1989), the point gauge
time series of temperature anomalies of the Arctic regions are close
to white noise. So, the first two rows of Table 5.5 show the process of
formation of white noise samples (whitening) of Arctic temperature
with the aid of a linear combination of, generally speaking, autocor-
related data of southern regions. For the surface air temperature this
process will be discussed in Chapter 7.
It is interesting that approximately the same feature (that the
observed fluctuations in the tropical regions are the main source of
climate anomalies for the northern latitudes) were obtained for other
time series (TASFC, T850, TS, QSFC, Q850) (Polyak, 1989).

5.5.2.2 AR Model of Troposphere Temperature


Here, we consider four troposphere temperature time series from the
same Oort data base averaged horizontally on each level for the en-
tire Northern Hemisphere and vertically in the four isobaric layers:
950-850 (Tl), 800-400 (T2), 350-150 (T3), 100-50 (T4) mbar
(see Matrosova and Polyak, 1990). The parameter estimates of the
corresponding four-variate model are given in Table 5.6.
5.5 Examples of Climate Models 233

Table 5.6: Estimates (exceeding 95% significance level) of parameters


of four-variate AR model.

Layers 950-850 800-400 350-150 100-50


Tl(t) T2(t) T3(t) T4(t) £

950-850, Tl(t+l) 0.34 _ 0.82


800-400, T2(t+l) _0.24 0.83 — — 0.63
350-150, T3(t+l) — — 0.81 — 0.62
100-50, T4(t+l) -0.16 0.2 0.76 0.71

Table 5.6 shows that temperature fluctuations of the lowest level


are determined only by its own history. Anomalies of the second layer
depend on its own past and on the fluctuations of the first layer of
the previous month.
Interrelation of the temperature fluctuations of the third layer,
including tropopause, with others is statistically insignificant and
not detected. So, the time series of the tropopause temperature can
be considered and modeled independently, which corresponds to the
understanding of the special physical nature of the tropopause.
As the last row of Table 5.6 shows, climate temperature fluctua-
tions of the highest layer are formed by the temperature fluctuations
of the previous month of all lower layers except the tropopause. We
can see that the main source of the troposphere temperature fluc-
tuations is fluctuation in its lowest level. In spite of the relatively
small vertical climatic temperature variations, these fluctuations can
be traced and estimated using multivariate stochastic models.

5.5.2.3 AR Model of Zonally Averaged Surface Air Tem-


perature of the Midlatitude and Northern Regions
The next results were obtained using mean monthly surface air tem-
perature anomalies averaged in three extra-tropical latitude bands:
27.5°-37.5° (Ts), 37.5°-67.5° (Tm), and 67.5°-87.5° (Tn). The ob-
servational interval is 95 years (1891-1985), and each time series has
1140 terms. One of the most complete surface air temperature data
bases from the World Data Center, Obninsk (Russia), was used.
')'\A
£jt.j .< 5 Multivariate AR Processes

Table 5.7: Estimates (exceeding of 95% significance level) of param-


eters of three-variate surface air temperature All model.

Time series Tn(f) Tm(t) Ts(t) e

Tn(t + l) 0.37 0.14 0.29 0.92


Tm(t + l) 0.37 0.10 0.91
Ts(t-fl) 0.34 0.90

The statistically significant parameter estimates, given in Table


5.7, are situated on and above the main diagonal of the parame-
ter matrix. Therefore, in spite of a different source, nature, and
time interval of these measurements (compared with Oort's data),
qualitative consideration of the results are approximately the same
as in Section 5.5.2; i.e., monthly mean anomalies of different lati-
tude bands are formed on average by the fluctuations of the previous
month in the band under consideration and in southern latitudes.

5.6 Climate System Identification


The AR models considered above appear not as a special physical
wording of stochastic climate modeling problems but were basically
evoked by the data bases at hand. Accurate physical formulation of
such problems must be done. A stochastic model should have the
basic statistical properties of the system under study and should rep-
resent our knowledge about this system in a convenient and compact
form. Generally, the model must be a concentrated expression of
modern achievements of the theory which facilitates better under-
standing of mutual conditioning of observed processes. The method-
ology of fitting the model and the estimation of its closeness to the
real system forms a special branch of mathematical statistics, sys-
tem identification. Contemporary identification methods (Eikhoff,
1983) of complicated stochastic systems show that for an adequate
description of such systems, parallel building of their physical and
stochastic models (SM) is necessary.
5.6 Climate System Identification 235

In climatology, the building of physical models (general circula-


tion models, GCM) predominates. However, there are grounds (Has-
selmann, 1976; North and Kim, 1991; Polyak, 1989) to suppose that
SM, taught by qualitative observations, can offer an equally valid
form of climate description and for this reason should be of great
scientific and practical significance. Moreover, from an economical
viewpoint, SM are very effective.
According to EikhofF (1983), a general scheme of linear identifi-
cation of climate system can be presented graphically (Figure 5.4).
Neither the physical nor the stochastic approach alone can be the
only method for model development. Figure 5.4 clearly shows that
perfection of GCMs must be connected with developing the stochas-
tic modeling and incorporating it into the process of building the
GCM.
Of course, the identification procedure in Figure 5.4 presents an
idealized picture because the physical climate models (GCM) were
basically developed without any linearization and stochastic mod-
eling. But this scheme may be helpful in the infinite process of
diagnosis and perfection of GCMs.
Specifically, the identification of a linear stochastic model of cli-
mate includes the solution of the following problems:
1. climate data presentation;
2. determination of model order;
3. determination of dimensionality and the number of variates;
and
4. determination of variates initial in nonstationary variations and
selection of appropriate approximations for them.
Let us briefly discuss these questions.
1. Among the many possibilities (i.e., sampling, averaging, fil-
tering, transformation) of presenting the multitude of global
and local climate time series, we shall mention those that are
given in two papers. Oort (1983) estimated many time series
of global atmospheric circulation statistics by averaging aero-
logical data over different latitude bands and altitude levels of
the earth's atmosphere. Some of these data were used to build
the above models. Another possibility was given by Matrosova
and Polyak (1990), who showed that troposphere temperature
observations of the Northern Hemisphere can be presented by
Figure 5.4: Linear identification of climate system.
5.6 Climate System Identificationn 237

monthly mean time series for 12 fixed spatial regions, inside of


which statistical structures of data are approximately identical.
Of course, a wide range of additional studies must be carried
out to reveal spatial regions where observations have similar
statistical structure and, thus, where averaging is permissible.
2. Determination of the appropriate order of a climate model may
be based on physical and statistical concepts. As to the for-
mer, it is evident that basic laws of hydrodynamics (equations
of continuity, movement, energy, etc.), used for developing the
GCM, include time derivatives of different variables not greater
than second order. Therefore, we should hardly expect that
the stochastic climate model would be greater than the sec-
ond order. Moreover, temperature, as a basic characteristic
of climate system, is included in the equations of energy and
diffusion as a first-order time derivative. This allows us to use
first-order autoregressive models as the first step of climate sta-
tistical modeling. Since the volume of climate data is limited,
the level of complexity of the stochastic model is limited, in
principle.
3. Determination of a number of variates is based on the concept
of causality, as defined by Kashyap and Rao (1976). Building
the above climate models we used a simple empirical approach,
choosing from Oort's data base the time series with the largest
first autocorrelation coefficients.
4. Identification of the essentially nonstationary elements is based
on statistical and physical analysis of tendencies of variation
within a given observational interval. The problem of choos-
ing an appropriate form of nonstationary stochastic models de-
pends upon the development of methodology for solution of the
classical problems of nonlinear modeling (Tong, 1990).
Concluding this chapter one can emphasize that multivariate SM
is a powerful methodology, which can be used in a wide range of
climatological studies, including diagnostic analysis and verification
of GCM. The results in Chapters 7 and 8 and in Polyak (1989, 1992)
show that the major advantage of multivariate AR models over the
univariate case is in the possibility of interpretation of the parame-
ter matrix as a matrix of interactions and feedbacks, even when the
forecast has low accuracy or makes no sense at all. This model's core
(matrix of parameters) may be used as a basis for physical analy-
sis in the process of comparison with analogous estimates obtained
238 5 Multivariate AR Processes

using simulated data. Reliable estimation of each parameter a^ al-


lows us to determine the weight of fluctuation Xjt contributed to the
variability of Xi(t-\-i)-
Analyzing sets of time series of different processes that character-
ize local or global climate makes it possible to estimate their mutual
influence on the variability of each and, for example, to separate
natural and man-induced variability,
For verification and evaluation of GCM, one must compare mul-
tidimensional probability densities of observed and simulated vari-
ables. In practice, different first- and second-order moments must be
compared. The second moments of data generated by ARM A models
are identical to corresponding moments of samples, which is impossi-
ble to guarantee for those simulated by GCM. Moreover, the matrix
of the estimated parameters of the multivariate ARMA process (the
matrix of interactions and feedbacks) contains all of the informa-
tion about the second moments in a short and physically interpreted
form, and it can be used for diagnostic analysis in the GCM verifi-
cation problem as a standard for comparison with analogous results
obtained using data simulated by GCM (see Chapter 7). In other
words, the Intel-comparison of the stochastic models of the analo-
gous systems of time series, both observed and simulated, gives the
unique possibility of making a statistical inference about the identity
or distinction of the observed and simulated climate processes.
Such an approach presents an opportunity to draw conclusions
and to make recommendations concerning identification of the num-
ber of variates (and their types) and the order of models, for choosing
appropriate regions of spatial averaging of time series, for comparing
different stochastic or physical models (or results of different sim-
ulations by the same GCM), and for studying physical regularities
and features of observations. Even the results obtained above with
minimum observations can be used for qualitative comparison with
analogous estimates found using data simulated by GCM. Therefore,
multivariate ARMA models accumulate both physical and statistical
qualities. This agrees with the requirements of the system identifi-
cation theory.
This preliminary study demonstrates the great theoretical and
practical possibilities of statistical modeling, which allows us to solve
modern climatological problems on a new qualitative and quantita-
tive level.
6

Historical Records

In this chapter, the historical records of annual surface air tempera-


ture, pressure, and precipitation with the longest observational time
series will be studied. The analysis of the statistically significant
systematic variations, as well as random fluctuations of such records,
provides important empirical information for climate change studies
or for statistical modeling and long-range climate forecasting. Of
course, compared with the possible temporal scales of climatic vari-
ations, the interval of instrumental observations of meteorological
elements proves to be very small. For this reason, in spite of the
great value of such records, they basically characterize the climatic
features of a particular interval of instrumental observations, and
only some statistics, obtained with their aid, can have more general
meaning.
Because each annual or monthly value of such records is obtained
by averaging a large number of daily observations, the corresponding
central limit theorem of the probability theory can guarantee their
approximate normality. In spite of this, we computed the sample
distribution functions for each time series analyzed below and eval-
uated their closeness to the normal distribution by the Kolmogorov-
Smirnov criterion. As expected, the probability of the hypothesis
that each of the climatic time series (annual or monthly) has a nor-
mal distribution is equal to one with three or four zeros after the
decimal point.

6.1 Linear Trends


As seen in this section, the straight line least squares approximation
of the climatic time series (see Table 6.1) enables us to obtain very

239
240 6 Historical Records

simple and easy-to-interpret information about the power of the long


period climate variability. Carrying out such an approximation, we
assume that the fluctuation with a period several times greater than
the observational interval will become apparent as a gradual increase
or decrease of the observed values. Using only a small sample, it is
impossible to determine accurately the amplitude and frequency of
such long-period climate fluctuation. Consequently, the straight-line
model is the simplest approach in this case.
Let us begin with an analysis of the annual surface air temper-
ature time series, the observations of which are published in Bider
et al., (1959), Bider and Schiiepp, (1961), Lebrijn (1954), Manlcy
(1974), and in the World Weather Records (1975). The units of the
observations are °C. The stations' names, the observational inter-
vals, and the linear trend parameter estimates are given in Table 6.1
(<TO and cr\ are the standard errors for constant and straight-line ap-
proximations, respectively). The results in this table reveal that for
seven (of the eighteen time series) the estimates of the parameter /^
are statistically significant at the 99% level (the t statistic value is
greater than 3). For ten of the time series, the estimates of the slope
parameter are statistically significant at the 95% level (the t statistic
value is greater than 2). Therefore, in such an approximation a
stochastic model of the variations of the annual temperature time
series analyzed is

where /?0 is the annual mean for a given station, /3i is the annual
rate, and X ( t ) is the random fluctuation.
Most of the estimates of /3j in Table 6.1 are very small, but posi-
tive, which makes it clear that some kind of small global warming has
taken place long before the industrial revolution and the subsequent
increase of CO2 in the atmosphere. The mean, of the estimates of
/?i in Table 6.1 is about 0.002 (that is, two Celsius degrees per 1000
years); the mean of the corresponding standard errors apl is 0.001.
Careful consideration of the results presented in Table 6.1 shows that,
as a rule, the statistically significant fti estimates are obtained for the
longer observational intervals. Therefore, the time series must have
a considerable duration so that the parameter ft\ can be accurately
estimated. This point, is more fully illustrated in Table 6.2, where
the linear trend parameter estimates and their accuracies are given
for the different subintervals of the central England time series with
a 315-year record of surface air temperature. The results reveal that
approximately 90 years of observations were necessary to obtain a
statistically significant estimate of the parameter fl\.
6.1 Linear Trends 241

Table 6.1: Linear trend parameter estimates of the annual mean


surface air temperature (°C) time series.

Station Observ. k+1 en ft 0-/3, t ff


r ffo
interval

Strasbourg 1806-1955 150 .73 .002 .001 1.1 .06 .73


Prague 1775-1955 181 .87 -.001 .001 0.8 .06 .87
De Bilt 1755-1955 201 .70 .002 .001 2.4 .05 .71
San Bernard 1818-1955 138 .65 .002 .001 1.7 .06 .66
Berlin 1756-1955 200 .83 .002 .001 2.3 .06 .84
Trieste 1803-1955 153 .61 .002 .001 1.4 .05 .61
Hoenpisinberg 1781-1955 175 .78 --.001 .001 0.8 .06 .78
Basel 1755-1957 203 .69 .003 .001 3.6 .05 .71
Iena 1770-1955 186 .76 .003 .001 2.5 .06 .77
Vienna 1775-1955 181 .83 -.001 .001 1.0 .06 .82
Swaneuburg 1735-1940 206 .73 .002 .001 1.9 .05 .07
Geneva 1755-1955 201 .64 .002 .001 3.0 .05 .65
Stuttgart 1792-1955 164 .81 .004 .001 3.3 .06 .83
Karlsruhe 1799-1955 157 .68 .008 .001 6.3 .06 .76
Paris 1757-1953 197 .70 .004 .001 4.4 .05 .73
Turin 1755-1911 157 .68 .001 .001 0.5 .05 .68
St.
Petersburg 1752-1966 215 1.04 .004 .001 3.2 .07 1.08
Central
England 1659-1973 315 .60 .002 .000 4.9 .04 .62

More elaborate studies of the annual temperature linear trend


estimates for different stations and separate subintervals of the pe-
riod of the instrumental observations reveal that the most noticeable
increase in temperature has taken place during the first 35 to 40
years of our century, which corresponds to the growth of large cities
such as Karlsruhe, Paris, and Berlin; naturally, the local climate was
affected by the development of urban industry.
The results show that the long-period temperature fluctuations
can be powerful enough for their rate to be statistically significant, or
the model (6.1) can be approximately valid for the limited periods of
time. This model is, of course, inappropriate for forecasting because
the variance of such forecasts is greater than the variance of the
observations (see, for example, Figure 1.2). Later, a more accurate
242 6 Historical Records

Table 6.2: Linear trend parameter estimates of the central England


annual surface air temperature (C°),

Observation k+1 o\ A) ft a
f>Q °A t "y o-Q
interval

1659 1973 315 .595 9.140 .002 .034 .0004 4.9 .037 .618
1664-1973 310 .599 9.142 .002 .034 .0004 4.8 .035 .621
1674-1973 300 .604 9.146 .002 .035 .0004 4.8 .036 .627
1684 1973 290 .609 9.166 .002 .036 .0004 4.1 .037 .626
1694-1973 280 .600 9.188 .001 .036 .0004 3.2 .036 .610
1704 -1973 270 .584 9.217 .001 .036 .0005 2.0 .036 .588
1714-1973 260 .589 9.218 .001 .037 .0005 2.1 .037 .593
1724 1973 250 .596 9.218 .001 .038 .0005 2.2 .038 .601
1734-1973 240 .590 9.202 .002 .038 .0005 3.1 .039 .601
1744-1973 230 .564 9.199 .002 .037 .0006 3.6 .038 .579
1754-1973 220 .569 9.205 .002 .038 .0006 3.5 .039 .584
1764 1973 210 .568 9.204 .003 .039 .0006 3.8 .040 .587
1774-1973 200 .578 9.226 .002 .041 .0007 3.1 .042 .590
1784-1973 190 .566 9.219 .003 .041 .0007 3.7 .042 .585
1794-1973 180 .569 9.237 .003 .042 .0008 3.2 .044 .584
1804 1973 170 .569 9.244 .003 .044 .0009 3.2 .045 .585
1814-1973 160 .573 9.258 .003 .045 .0010 2.9 .046 .587
1824-1973 150 .557 9.281 .003 .046 .0011 2.4 .046 .566
1834-1973 140 .540 9.266 .004 .046 .0011 3.3 .047 .560
1844 1973 130 .534 9.288 .004 .047 .0012 2.9 .048 .550
1854- -1973 120 .529 9.301 .004 .048 .0014 2.8 .050 .545
1864-1973 110 .516 9.316 .004 .049 .0015 2.7 .051 .531
1874-1973 100 .507 9.308 .006 .051 .0018 3.5 .054 .536
1884-1973 90 .496 9.340 .006 .052 .0020 3.1 .055 .520
1894-1973 80 .465 9.409 .003 .052 .0022 t.4 .052 .467
1904-1973 70 .474 9.420 .004 .057 .0028 3.1 .057 .476
1914 1973 60 .482 9.445 .002 .062 .0036 0.6 .062 .480
1924-1973 50 .443 9.490 -.002 .063 .0043 0.4 .062 .440
1934-1973 40 .461 9.520 -.010 .073 .0063 1.5 .074 .470
1944 1973 30 .476 9.497 -.016 .087 .0.100 1.6 .089 .489
1954-1973 20 .486 9.390 -.001 .109 .0189 0.0 .106 .473
1964-1973 10 .231 9.390 .027 .073 .0255 1.0 .074 .233
6.1 Linear Trends 243

approximation based on the spectral characteristics of the time series


will be presented.
Consider the linear trend parameter estimates (Table 6.3) of the
longest historical records of the annual atmospheric pressure time
series in ten European cities. For any data set, no /3i estimates are
statistically significant at the 95% level. The fl\ estimates are not
statistically significant even for any ten-year (or greater) subinterval
of observations of any of the above time series. Therefore, it is clear
that atmospheric pressure had no sufficiently powerful long-period
climatic fluctuations. Indeed, the results reveal that atmospheric
pressure is the most stable characteristic of climate; moreover, as will
be shown shortly, their annual time series can be used as a sample of
a white noise process like a sample produced by any random number
generator.
Considering the two longest historical records of precipitation in
Swanenburg (Lebrijn, 1954) and in Paud Hole, East England (Crad-
dock and Wales-Smith, 1977), one can see that there are no statisti-
cally significant estimates of the parameter /?j (see Table 6.4). The
standard errors are large, and the precipitation variability is greater
than the variability of the air temperature or atmospheric pressure.
The additional analysis of linear trends on separate (10-, 20-, and
30-year) subintervals in Swanenburg did not yield any statistically
significant estimates of the parameter j3\.
Before concluding our consideration of the climatic trends, it is
important to emphasize that the problem of their interpretation is
significantly distinct from analogous problems in industry, astron-
omy, etc., where observations of certain variables are distorted by
random errors. Of course, meteorological observations also contain
errors; but when processing data which is a result of a large-scale av-
eraging (such as annual values), one can assume that the fluctuations
generally characterize the natural variability of the analyzed climatic
element. This variability presents the unity of short- and long-period
fluctuations; consequently, the notion of the trend (for which the ap-
proximation is obtained), and the random fluctuations (deviations
from the trend) are very conditional. A variation that looks like a
trend on the 10-year time interval can be a part of a random fluctu-
ation on the 100-year time interval. Similarly, the secular variations
of meteorological observations can be part of a random fluctuation
on the 1000-year interval. Unity, the deep internal interdependence
of the climate fluctuations of different time scales, reveals the condi-
tional and subjective character of distinguishing between a trend and
244 6 Historical Records

Table 6.3: Linear trend parameter estimates of the annual atmo-


spheric pressure (mm) time series.

Station Observ. k+1 a\ 0i


ff
ft1 t O-y <n>
interval

Basel 1755-1959 205 0.83 -.000 0 .001 0.1 .06 0.83


Geneva 1768-1960 193 0.82 -.0000 .001 0.1 .06 0.82
Copenhagen 1842 1960 119 0.97 .000 .003 0.1 .09 0.96
Happarauda 1860-1960 101 1.22 .000 .004 0.1 .12 1.22
Greenwich 1854-1949 96 1.37 .004 .005 0.8 .14 1.36
Valencia 1866-1960 95 1.25 .006 .005 1.3 .13 1.25
Aberdeen 1866 1960 95 1.63 -.0077 .006 1.1 .17 1.63
Bordeaux 1868-1960 93 1.77 .001 .007 0.2 .18 1.76
Torshavnn 1873 1959 87 1.25 -.003 3 .005 0.5 .13 1.25
St.
Petersburg 1881-1950 70 1.32 -.002 2 .008 0.3 .16 1.31

a random component. But the distinction must be imposed method-


ologically when wording the climatological trend analysis problems
because it can affect the selection of statistical methods and criteria.
In the study of climate variations, it is often presupposed implic-
itly that smaller microscale temporal fluctuations are a result of a
large number of separate factors, and it is possible to consider their
summary effect to be random. The larger fluctuations, such as the
secular trend, are, supposedly, a result of only a few (or even one)
physical phenomena, which can be monitored and explained. For ex-
ample, one of the hypotheses about global warming during the first
third of this century assumed that the refinement of the atmosphere's
transparency resulted from a single factor: the lack of large volcanic
eruptions at that time (Budyko, 1974). Today, there is an attempt
to explain global warming by another single factor: the increase of
CO2 and other trace-gas concentrations in the atmosphere.
This reasoning shows that the way in which the climatic trend
problem is studied represents an attempt to separate large- and
small-scale climate fluctuations. In the framework of some prob-
lems, our main interest could be in analyzing long-period variations
and their possible models, while in the framework of other problems,
our primary interest could be in analyzing short-period fluctuations.
6.2 Climate Trends over Russia 245

Table 6.4: The linear trend parameter estimates of the annual mean
precipitation ( m m ) time series.

Station Observ. k+1 "\ 00 01 "Pi t °Y ffo


interval

Swanenburg 1735-1944 210 124 743 .1 .1 .9 8.5 124

Paud Hole 1726 1975 250 501 2433 3


"^ .4 .7 28.4 449

Where exactly, though, is the boundary between long- and short-


period climate fluctuations? Does the use of a t-statistic criterion
make it possible to impose a distinction that has any physical sense?
At any rate, the results of the approximation of the annual mean
time series of meteorological elements can be considered as one of the
characteristics of the power of long-period variations, part of which
are observed on a given temporal interval.

6.2 Climate Trends over Russia


Let us begin by considering the annual surface air temperature time
series for the 1891 1987 period of 104 stations, more or less evenly
distributed over the territory of Russia (of course, the network is
more dense in the European part than in Siberia). The data, base
was collected and maintained in the Main Geophysical Observatory
(MGO), St. Petersburg, Russia. Citing the station names and corre-
sponding estimates would take too much space, so only the summary
results are given in Table 6.5.
Least squares straight-line approximations were carried out not
only for the whole observational period, but for various subperiods
(see the first line of Table 6.5), and the ft\ estimates were averaged
over the stations ranging in accordance with the amount of popu-
lation (see the first column of Table 6.5). It is clearly seen that
there is a noticeable difference in the mean estimates of parameter
/?i between large cities with a population of more than 1,000,000 and
smaller cities, especially when the observational period lasts more
than 50 years. For such periods, the rate of temperature increase in
the large cities (about 8°C/1000 years) is twice more that that of the
Table 6.5: Summary results of the temperature trend based on the observations of
the 104 Russian stations for the period 1891 to 1987. The numerator presents the
BI estimates in a C per 1000 years; the denominator is the number of estimates with
/?i f-statistics greater than 2.
Pei-iods of abservat ons
Averages over 1891- 1901- 1911- 1921- 1931- 1941- 1951- 1961-
Pi estimates of 1987 1987 1987 1987 1987 1987 1987 1987
12 largest cities
with population 8/7 8/6 8/5 ?/3 8/3 17/3 5/0 -7/0 0
over 1,000,000
21 cities with
population from 4/10 3/3 5/4 4/4 •f i o
5/ 3 11/5 /3
0U / Ce -19/2 2
0.3 to 1.0 million
34 cities with
population from 4/10 3/7 5/8 4/6 5/7 14/9 3/2 — 14/1
0.1 to 0.3 million
i
37 cities with !
1

population less 5/17 4/9 4/3 4/7 7/T 15/9 7/1 -12/0 0
than 100,000

all 104 stations 5/44 5/25 5/25 5/20 6/20 15/26 5/6 -12/33
6.3 Periodograms 247

smaller ones (about 4°-5°C/1000 years).


For fewer than half of the stations, the parameter /?i estimates
are statistically significant (for the large cities the number of such
estimates is more than half) at the 95% level (^-statistic is greater
than 2) for the data on the entire observational interval; and only 20
to 25% of such estimates are statistically significant for the smaller
subintervals. Since 1951, more than 95% of the fi\ estimates for the
individual stations are not statistically significant.
We can conclude this brief consideration with the following re-
marks:
1. The results obtained show that, for about a century, surface air
temperature increased in almost all Russian cities at an average
rate of about 5°C per 1000 years.
2. Temperature increases in the large cities (with population larger
than 1,000,000) were greater (about 8°C per 1000 years).
3. Since there are only a few large cities, a greater temperature
increase there does not substantially influence the rate of warm
ing over the large territories. When analyzing the global tem-
peratures using spatial averaging of the entire global network,
one probably should exclude data from the large cities because
their weights are too great (with respect to the size of their
territories) compared with the entire surface of the earth.
4. The greatest temperature increase was observed in the first
third of our century.
Now consider the annual sums of precipitation for 104 time series
of the most important agricultural regions of Russia for the same
period of time. The precipitation data base was also collected and
maintained in the MGO. The results, given in Table 6.6, show the
tendency to decrease in the first third of our century and then to
increase slightly. Statistically significant estimates of (3\ are obtained
in about one fourth (or less) of the cases.

6.3 Periodograms
6.3.1 Central England Surface Air Temperature
Studying a pedodogram of time series is really a deterministic ap-
proach since Fourier transform and computing amplitudes are not
248 6 Historical Records

Table 6.6: Summary results of the precipitation trend based on the


observations of the 104 regions in Russia for the period 1891 to 1987.
The numerator consists of averages over all /?i estimates (mm per
year); the denominator is the number of estimates with the /^ t-
statistics greater than 2.

Periods of observations

1891- 1911- 1931- 1951- 1971-


1987 1987 1987 1987 1987

-0.01/27 -0.15/11 0.20/7 0.12/2 1.17/13

accompanied by any statistical estimation procedure. In fact, a pe-


riodogram presents objective empirical information about the power
of the harmonics with the various frequencies that compose the time
series on the observational interval.
Let us begin our consideration with the central England annual
mean surface air temperature (Manley, 1974). This time series spans
315 years (from 1659 to 1973), thus making it the longest temper-
ature record known in climatology. The periodogram of this time
series is shown in Figure 6.1 (the estimates are presented in the per-
centages of the sample variance value). Therefore, the periodogram
values depend on the number of points in the time series, but the
advantage of such a unit is in the possibility of comparing very explic-
itly the power of the fluctuations with different periods. Figure 6.1
shows that the periodogram values fluctuate significantly for differ-
ent frequencies. To facilitate examination of Figure 6.1, the periods
(105, 24, and 15 years) for the significant peaks are marked, and the
percentages (8%, 5%, and 4%) of the variance corresponding to these
peaks are also shown.
To understand whether or not these peaks have any physical
meaning, it is necessary to answer the following two questions:
1. To what extent are these maxima stable in space? In other
words, do other stations' periodograrns of the temperature ob-
servations have exactly the same features?
6.3 Periodograms 249

Figure 6.1: Petiodogram (1) and spectral density estimates (2) of the central
England annual mean surface air temperature time series (1659-1973).

2. To what extent are these maxima stable in time? In other


words, are the peak periods and powers the same for any subin-
terval of this and other time series?
To answer the first question, it is necessary to compare periodograms
that have been computed with data from a variety of stations. The
results of such a comparison for the 18 longest temperature histor-
ical records, listed in Table 6.1, are given in Polyak (1975). These
results show that in the low frequencies (with periods greater than 12
years) relatively large-power spikes are present on the periodograms
of any station. But, in general, the frequency values of these spikes
vary significantly for the various stations; or, in other words, the
periodogram features do not have a regular spatial character.
Such irregular peaks on the periodograms of the various stations'
temperature time series make it clear that, if the number of stations
is sufficiently large, it is possible to find "cycles" at practically any
frequency, each of which is inherent only to the data chosen. This
250 6 Historical Records

phenomenon explains why so many cycles with a wide variety of peri-


ods were found in early publications of climatologists. A discussion of
this question can be found, for example, in Monin (1969). There are
several reasons for the development of the various long-term powerful
fluctuations of the air temperature. For instance, volcanic eruptions
lead to a decrease of air temperature with a gradual return to normal
levels over several years; global processes connect with atmospheric
carbon dioxide fluctuations; the earth orbit parameters change (as-
tronomy climate theory); the intense growth in the industry of large
cities affects temperature and leads to the gradual variation of the mi-
croclimate; and so on. Consequently, the cycles (large periodogram
peaks) of the climatological temperature time series of different sta-
tions must be random and unstable in space.
To answer the second question about the spike temporal stability,
it is possible to conduct a more detailed analysis of periodograms
on the separate subintervals of the entire interval of observations
and to compare the frequencies of their most powerful peaks with
those presented in Figure 6.1. Such an analysis can help to locate
in time both the most powerful fluctuations and the possible trans-
fer of power from one frequency to another. This kind of analysis
has been done, for example, for the St.Petersburg annual surface
air temperature time series with the 1752-1966 observational period
(Polyak, 1975). In that study, the periodograms were considered on
the various overlapping 128-year subintervals. The results showed
that the frequencies of the most powerful fluctuations are different
for different subintervals, or, in short, the cycles mentioned above
are temporally unstable.
Similar results were obtained for the historical temperature records
of some other stations (Polyak, 1975). Consequently, separate pow-
erful spikes on the periodograms (i.e., cycles) have a random char-
acter, and their appearance and disappearance in space and time is
the manifestation of the sample variability. This means that it is
senseless to use these cycles for long-period forecasts. A similar con-
clusion was obtained, for example, in the program "Understanding
Climate Change" (1975).
Therefore, the primary feature of the periodograms of the climate
temperature time series is a relatively significant part of the variance
corresponding to some of the low frequency harmonics.

6.3.22 Atmospheric Pressure and Precipitation


This study of the mterannual variability of atmospheric pressure and
precipitation is based on the historical records of the atmospheric
pressure for a 205-year span from 1755 to 1959 in Basel (Bider and
6.4 Spectral and Correlation Analysis 251

Schiiepp, 1961) and of the precipitation for a 210-year span from


1735 to 1944 in Swanenburg (Lebrijn, 1954).
The periodograrn of the atmospheric pressure (Figure 6,2, top
has three separate peaks with periods of 4, 3.5, and 2.5 years, re-
spectively. The power of these peaks reaches about 4% of the sample
variance. For periods of more than 4 years, there are no amplitudes
with power exceeding 3% of the variance.
The periodograrn of the precipitation time series (Figure 6.2, bot-
tom) also has separate peaks, the maximum of which (6% of the
variance) corresponds to a period of approximately 4.1 years. But
generally, the distribution of the variances along the frequency axis
has more or less even character for both time series. Multiple pe-
riodograms analogous to Figure 6.2 of many different time series of
atmospheric pressure and precipitation (Polyak, 1975; 1979) reveal
that the random cycles of pressure and precipitation with relatively
large amplitudes are unstable in time and space. In other words,
they appear as result of the sample variability.

6.4 Spectral and Correlation Analysis


6.4.1 Annual Mean Surface Air Temperature
In this subsection the spectral and correlation analysis of the Central
England temperature time series will be carried out. The correspond-
ing periodograrn has already been studied
Figure 6.3 presents the cumulative spectral function of this time
series and 99% confidence intervals for the white noise. The estimates
exceed the confidence interval above. This means that the time series
considered is not a sample of a white noise.
The spectral density estimates, given in Figure 6.1 (2), were ob-
tained by smoothing the periodograrn using a regressive filter (1.83)
with m = 3 and a width 2r + 1 = 35 points.
This spectrum's most noticeable feature is that it has a maximum
at low frequencies (near the origin). This maximum has also been
found in the spectral estimates of a variety of climate temperature
time series (Landsberg et al., 1959; Monin and Vulis, L969; Monin et
al., 1971; Polyak, 1975; Madden, 1977; and so on). Most of the va
ability was caused by the white noise component of the fluctuations.
It was shown (Polyak, 1975) that the shapes of the spectral density
estimates for the longest temperature historical records of different
stations and different observational intervals are almost identical.
252 6 Historical Records

Figure 6.2: Periodogram (1) and spectra! density estimates (2) of the Basel
annual mean atmospheric pressure (top) time series (1755-1959) and of the Swa-
nenburg annual precipitation (bottom) time series (1735-1944); (3) is white noise
spectrum.
6-4 Spectral and Correlation Analysis 253

Figure 6.3: Cumulative spectral function (1) of the central England surface air
temperature time series and the 99% confidence interval (2) for the white noise
(3).

Figure 6.4 gives the central England time series autocorrelation


function, which corresponds to the spectrum in Figure 6.1 (2). The
first several autocorrelation estimates are small (less than 0.3), pos-
itive, and statistically significant at the 99% level. The white noise
level, which is determined by the difference between one (value at
the zero lag) and estimates 0.2-0.3 at lag equal to one year, is large
(about 70 to 80 % of the variance). Approximately the same auto-
correlation estimates were obtained for other temperature historical
records listed in Table 6.1 (Polyak, 1975).
These results enable us to make the following comments:

1. The spectra of the climate air temperature time series have a


small maximum in the vicinity of the origin, which corresponds
254 6 Historical Records

Figure 6.4: Autocorrelation function of the central England temperature time


series.

to the periods of decades or centuries. Such "red spectrum


character" reveals the possibility of the existence of relatively
powerful low-frequency temperature fluctuations.
2, The temporal variability of climate air temperature primarily
results from white noise, which makes up approximately 70 to
80% of the variance.
3. In contrast to the periodograms, the spectral and correlation
estimates are spatially and temporally stable.
The above nonparametric consideration (together with the stud-
ies by Polyak, 1975; 1979) reveals that an approximate model for the
6.4 Spectral and Correlation Analysis 255

annual mean air temperature time series of a point gauge can be the
random process

where N(t) is a white noise, S(t) is the Markov random process


[independent of N(t)] with the autocorrelation function e~a\T\, and

The value of p = c,~a varies from station to station; for the middle
latitudes the range of its variations is about 0.1 to 0.3. Each of the
empirical approximations considered [(6.1) and (6.2)] has different
features. Although (6.1) realistically describes the long-period tem-
perature variation during the last century, it proves inadequate for
forecasting even in the near future.

6.4.22 Atmospheric Pressure and Precipitation


Let us look at the spectral and correlation characteristics of the at-
mospheric pressure and precipitation time series, the periodograms
of which are given in Figure 6.2. The parameters of the regressive
filter (1.83) that was used for the spectra estimation are the same for
both time series considered; the order equals three, and the width
equals 25 points (m = 3,r = 12).
Figure 6.2 (top, 2) gives the spectral estimates of the Basel at-
mospheric pressure along with the white noise spectrum (3). These
estimates are not smooth. However, by increasing the filter width,
they can be made smoother and closer to the white noise spectrum,
The minimum of the spectrum estimates in Figure 6.2 (top, 2) cor-
responds to the interval with periods greater than ten years, which
reveals that powerful climatic fluctuations of atmospheric pressure
are unlikely, just as one would expect. Nevertheless, it is clear that
the spectrum estimates are very close to the white noise spectrum.
Figure 6.5 (top) compares a theoretical cumulative spectral func-
tion of the white noise with the estimated cumulative spectral func-
tion of the Basel atmospheric pressure.
The estimates are within the 99% confidence interval; that is,
the analyzed time series presents a sample of white noise. This same
conclusion is derived when one considers the autocorrelation function
[Figure 6.6 (left)], whose values are very small and whose deviations
from zero are statistically insignificant at any standard level of sig-
nificance for all the non-zero lags.
256 6 Historical Records

Figure 6.5: Basel atmospheric pressure (top) and Swanenburg precipitation


(bottom) cumulative spectral functions (1) and the 99% confidence intervals (2)
for the white noise (3).
6.4 Spectral and Correlation Analysis 257

Figure 6.6: Basel atmospheric pressure (left) and Swanenburg precipitation


(right) autocorrelation functions.

The analogous consideration of the Swanenburg precipitation time


series reveals that all the estimates of the cumulative spectral func-
tion [Figure 6.5 (bottom)] are within the 99% confidence intervals for
white noise. Consequently, the time series analyzed are close to the
white noise sample. The same conclusion can be made analyzing the
corresponding correlation function [Figure 6.6 (right)], the values of
which for the non-zero lags are very small and statistically do not
differ from zero.
The deviation of the spectral density estimates [Figure 6.5 (bot-
tom)] from the white noise spectrum is slightly greater than the cor-
responding deviation for the atmospheric pressure, but the closeness
of it to the white noise is indisputable.
Along with Basel and Swanenburg data, precipitation and atmo-
spheric pressure time series of many other stations were also analyzed
258 6 Historical Records

in Polyak (1975; 1979). Because the results are similar to those ob-
tained above, it makes no sense to cite them here. We will only make
the following comments:

1. The spectral and correlation characteristics of the climatic time


series of the atmospheric pressure and precipitation are stable
in space and in time.

2. It is quite probable that the annual mean atmospheric pressure


and precipitation time series are samples of white noise. Even
on the separate subinlervals of each of these time series, the
closeness of the observations to the white noise sample is ob-
vious (Polyak, 1975; 1979). Therefore, any attempt to forecast
these time series is meaningless,
3. Since the duration of the records considered is unique, the spec-
tral and correlation estimates obtained not only carry the value
of appropriate examples but also show the general character-
istics of the atmospheric pressure and precipitation time series
for the middle latitudes of the Northern Hemisphere.

The possibility of forecasting climate by statistical methods is de-


termined by the character of the statistical and harmonic structure
of the long-period fluctuations. The results in this section (together
with those published in Polyak, 1975; 1979; 1989) reveal that the
anomalies of the monthly and annual mean climate time series have
a very high level of additive white noise. Based on these publications,
Table 6.7 gives the range of estimated values of the autocorrelation
coefficient for one month lag for the middle latitude Northern Hemi-
sphere time series of different climatological variables.
The air temperature observations have the largest value of the
autocorrelation coefficient. Table 6.7 provides information necessary
for the temporal statistical modeling of corresponding climate data
and also offers its forecasting value because the first autocorrelation
is all of the information required for building the A R ( l ) model.
The autocorrelated component of the climate time series of many
meteorological elements (with the exception of precipitation and at-
mospheric pressure) is conditioned by the small maxima in the low-
frequency part of the spectra with periods of several decades or cen-
turies. These maxima offer evidence of the possibility of the existence
of Jong-period climate fluctuations with relatively powerful ampli-
tude. Such fluctuations, which are usually referred to as cycles, have
6.5 Univariate Modeling 259

Table 6.7: Range of the autocorrelation coefficient estimates (for


one month lag) of the normalized anomalies of the monthly mean
clirnatological time series in the middle latitudes of the Northern
Hemisphere.

Element Autocorrelation coefficient,

Air temperature on the tropopause level 0.20-0.40


Surface air temperature 0.20-0.35
Thickness (500-1000) 0.15-0.30
Atmospheric zone 0.10-0.30
Geopotential of 500mb level 0.10-0.20
Direct radiation 0.10-0.20
Heat budget 0.10-0.20
Precipitation 0.00-0.10
Zonal component of the geostrophic
wind velocity of the 500mb level 0.00-0.15
Surface atmospheric pressure 0.00-0.10

a random spread in time and space. Considering only natural vari-


ability, it is important to recognize that the forecasting value of any
cycle is equal to zero.
From the example in Section 2.5, it follows that temporal averag-
ing of daily temperature observations with the first autocorrelation
coefficient equal to about 0.8 must lead to a monthly time series with
the first autocorrelation coefficient equal to about 0.07. The values in
Table 6.7 are greater than 0.07, which shows that there are long-term
variations beyond the synoptic scale fluctuations that affect climate
variation.

6.5 Univariate Modeling


Application of the nonparametric methodology of spectral and cor-
relation analysis to the climate (monthly or annual mean) time series
of hydrometeorological elements has provided a way to understand
260 6 Historical Records

the physical and statistical nature of long-period climate fluctuations


and to explain such phenomena as multiple climate cycles.
By analyzing the graphs of the periodograms of the various hy-
drometeorological observations at the various points of the earth,
one can see that each and every one of them has several separate,
relatively powerful peaks corresponding (in its totality) to virtually
any period ranging between several months to several hundred years.
These peaks are unstable in space and in time; their frequencies are
different for different stations or subintervals of the same time se-
ries. The part of the variability that corresponds to each maximum
does not exceed 5 to 8% of the sample variance, which excludes any
forecasting value. Such peaks clearly demonstrate the random char-
acter of the periodogram, or, in other words, the sample variability
of the time series. The periodogram of the climate time series (and
its peculiarities) does not carry general theoretical value. Rather, it
represents a basis for the consequent estimation of the spectral den-
sity. Any attempt to give a profound meaning to the separate peaks
(that is, cycles) is equivalent to attempting to build a law of behavior
of a random variable based on only one observation. Smoothing (or
averaging) of periodogram values in separate frequency subintervals
is the key feature of spectral density estimation methodology.
Even the first attempts (for temperature data) to estimate cli-
matic spectra have revealed the statistical nature of climate fluctua-
tions from the viewpoint of the theory of probability: White noise, a
random process with statistically independent separate values, dom-
inated in the composition of such iluctuations.
The theoretical spectrum of white noise is constant for all fre-
quencies; but, of course, the periodograms of its samples (for exam-
ple, those generated by the computer random number generator) are
similar to the periodograms of the climate time series. However, in
spite of the high level of white noise in climate fluctuations, some
spectral estimates of the separate stations' temperature time series
enable us to establish with a high degree of confidence the existence
of the single stable spectral feature: a small maximum in the vicinity
of the zero frequency.
This spectral maximum at the origin is more pronounced for the
time series of stream flow, air temperature, and radiative observa-
tions (see Table 6.7) and is significantly smaller (or does not exist at
all) for the observations of the precipitation, the atmospheric pres-
sure, the velocity and the direction of wind, the geopotential of the
standard isobaric surfaces, and some other elements, the estimated
spectra of which do not differ significantly from the white noise spec-
trum. Generally speaking, then, the assertion about the pure noisy
6.5 Univariate Modeling 261

character of the local climate's variations cannot be accepted. The


more probable conclusion is that corresponding climatic observations
are presented by the sum of a very small signal (autocorrelated com-
ponent) and a significant portion of white noise, against the back-
ground of which the signal is often not detectable at all.
It is clear that slow variations of local climates during the earth's
history must be associated with the low-frequency spectral maxi-
mum. Therefore, we must recognize that the study of this extreme
has decisive meaning in the statistical description of climate change.
This maximum provides the evidence for the possibility that there are
climate fluctuations of a nonregular character on time scales ranging
from decades to thousands of years.
The above analysis of meteorological data for the instrumental
period of observations using the random process theory presents one
possible description of the statistical nature of climate. Such a de-
scription shows that white noise assumes the primary role in the
formation of climatic fluctuations. The autocorrelated component
is very small, but the spectral maximum in the vicinity of the zero
frequency demonstrates its existence.
However, in the framework of stationarity only the autocorre-
lated component can carry any information about climate change.
Only this component determines the climate predictability (climatic
memory). The white noise component is unpredictable, and the cli-
mate would also be unpredictable and unchangeable if the climatic
fluctuations were entirely formed by the white noise.
White noise forms 70 to 100% of the monthly and annual mean
anomalies of the point gauge climate time series in the middle lati-
tudes. But in spite of the fact that the autocorrelated component is
drowned out in the white noise, its small value (for some elements)
reveals two peculiarities. First is the step from one for zero lag to
the 0.0-0.3 values for the lag that is equal to one month or one year.
This step demonstrates the possibility of presenting the climate time
series as the sum of the signal plus the white noise.
Second, the closeness to zero of the autocorrelations for the lags
greater than two, or sometimes three or four, months (or years for
the annual data) enables us to interpret the signal not as a pure
deterministic component, but as, in the first approximation, a ran-
dom process with several non-zero autocorrelations. (If a signal were
a pure deterministic component, the corresponding autocorrelation
function would not approach zero). It is natural to think that the
signal characterizes the physics of the observed process. Its reliable
detection stands as one of the problems of statistical climatology.
262 6 Historical Records

The discussion above of spectral and correlation estimates, to-


gether with results of the theoretical study of the forecast error of
different kinds of tmivariate ARMA processes presented in Chapter
4, enables us to draw certain conclusions about the possibility of
identifying ARMA models of separate stations' climate time series
and the limits of their statistical predictability without having to fit
such models at all.
Indeed, if the maximum value of autocorrelations does not exceed
0.3, then according to the results of Section 4.7 it makes sense to
forecast such a time series for, at most, one step ahead independent
of the order and type of the fitted ARMA model. In the best case,
the standard deviation of the forecast error would consist of 90% of
the standard deviation of the observations, and the forecast estimates
would be very close to the time series mean. The practical uselessness
of such a forecast is quite clear.
Therefore, formulating the problem of univariate statistical ARMA
modeling of the climate time series of point gauges presents no spe-
cial interest, no deep physical matter. As for the type and order
of such models, it is most likely that only the following three cases
should be considered.
1. The white noise model, which matches the point gauges time series
of precipitation, atmospheric pressure, wind, and perhaps some other
variables.
2. The first order autoregressive model (annual river stream flow,
air temperature, radiation), whose spectrum estimates have a small
low-frequency maximum that characterizes the A!.t(l).
3. The signal plus white noise model, where it is possible to accept
the AR(l) process for a signal. This latter model, ARMA(1.,1), is
considered the most general, since the white noise and the AR(i)
model represent; its particular cases.
Finally, nonparametric analysis of the historical records enables
us to establish the reality of the signal-plus-white-noise character for
the climatic time series of the point gauges. It also points out the dif-
ficulty of reliably detecting a signal that is masked by this noise. In
other words, the climatic time series of point gauges must be further
processed by an appropriate procedure of spatial and temporal aver-
aging or filtering in order to evaluate the character of the long-period
fluctuations (the signal) responsible for climate change.
6.6 Statistics and Climate Change 263

6.6 Statistics and Climate Change


As we have already discussed, the periods of temporal fluctuations
of meteorological elements range from a fraction of a second to many
millennia.
However, climate change study concerns variations in periods
that range from several weeks to many millennia. But by estimating
the climate statistics (that is, the statistics of data with large-scale
temporal averaging), we are actually analyzing the resulting state
of the climate system, which is conditioned by the interactions and
feedbacks of multiple physical factors and processes taking place in
the atmosphere and ocean in real time. Such interactions and feed-
backs, which are generally represented by the hydrodynamic laws,
determine the character of fluctuations of the meteorological ele-
ments of any spatial and temporal scales. Estimated statistics pro-
vide a qualitative picture of this resulting state. Because for most
of the instrumental period of meteorological observations the mea-
surements have been done on the earth's surface, our knowledge of
climate has been limited to the surface climate. The description
of the local climate has had a relatively qualitative character based
on estimates of climatic means (norms) and standard deviations for
the separate months, seasons, and years of different regions of the
earth. Unfortunately, these norms and variances do not give the
entire representation about the large-scale spatial and long-period
temporal variability of meteorological elements. In order to obtain
a more complete description of climate, it is necessary to study the
physical and statistical nature of the spatial and temporal climatic
fluctuations in the form of multivariate stochastic models.
With the development of the system of radiosonde stations and
corresponding data bases, several projects have been undertaken to
describe the meat) climate regime of the upper atmosphere (Oort,
1978; 1983; and others). These investigations, which are not yet
completed, have significantly widened our knowledge of the atmo-
spheric processes and their possible variations.
The primary subject under investigation is the cause of climate
change. Various investigations have suggested several possible ex-
planations (hypotheses), though none has become a generally recog-
nized theory. The reason for this is that the climatic fluctuations
are forced not only by separate powerful phenomena (for example,
volcanic eruption, and so on) and processes, but also by interactions
and feedbacks of multiple factors of a mechanical, physical, chemical,
and biological character. Nevertheless, it is important to enumerate,
at least briefly, the possible causes of climate change discussed by
264 6 Historical Records

Budyko, 1974; Hays et al. 1976; Monin, 1969, 1977; Monin and
Vulis, 1971; and others.
One set of theories is based on elements external to the climatic
system, which include the following:
1. The solar radiation power variation.
2. The amount of solar energy that reaches the earth surface due
to variation of the concentration of intercelestial dust.
3. The variation of the seasonal and latitudinal distribution of
incoming radiation due to variation in Earth's orbit geometry.
4. The content of volcanic dust in the atmosphere.
5. Earth's magnetic field and the tectonic motion of Earth's crust.
Another set of theories is based on elements internal to the climatic
system. This set of theories assumes that there are physical mecha-
nisms with sufficiently long response feedbacks to stimulate climatic
fluctuations with periods of the order of thousands of years. These
internal elements are:
1. The growth and decay of ice shields.
2. Snow cover variations.
3. Cloudiness variations.
4. Variation of the evaporation from the ocean surface near the
equator.
5. Ocean deep-water circulation.
6. Deforestation.
There is a set of theories which assumes that the climatic system
can have several different stable states and that it can move from
one state to another (a scientific assumption otherwise known as
the transitivity of climate). In the last few decades, anthropogenic
causes, which result from the atmospheric C02 (and other gases)
have increased, with the heat budget variation has dominated other
theories.
Even this very brief list of possible causes of climate change
reveals the complexity and multiplicity of climatological problems.
6.6 Statistics and Climate Change 265

Many papers (such as Lorenz, 1977; Monin, 1972; Flon, 1977; Has-
selmann, 1976; and others) have considered the different physical as-
pects of climate change. There were also attempts (Mitchell, 1976)
to build a spectrum of climatic fluctuations with periods that range
from one hour to 109 years and to explain the spectrum's features.
Mitchell differentiates between the deterministic and stochastic mech-
anisms of climate variations, estimates the temporal scales of differ-
ent processes of nature, and presents them as features of his spec-
trum.
In many publications that analyze the nature of climate fluc-
tuations, estimates of different climate statistics (for example, the
spectral and correlation characteristics) are considered.
The continuous evolution of the living world for millions of years
enables us to assume that earth's climate variations have taken place
primarily evolutionarily and, therefore, have occurred relatively slowly
(though, some regions, of course, experience catastrophes).
Indeed, the value of a trend (see Section 6.1) of any climatic pro
cess, such as surface air temperature, on the time intervals, compared
with the period of instrumental measurements, is very small. It is
several orders less than the climatic noise level or the accuracy of the
instruments used today. For example, if for a period of 100,000 years
the mean surface air temperature of a certain region has changed by
about 10°C, the mean annual rate is equal to approximately 10-4 °C.
It is clear that no observations with such accuracy currently exist.
Therefore, the systematic variations of climate tend to be smaller on
several orders than the accuracy of measurements. Consequently, it
is very difficult to obtain a rigorous physical or statistical description
of such long-period fluctuations.
7

The GCM Validation

In this chapter the observed and simulated (by the Hamburg GCM)
Northern Hemisphere monthly surface air temperatures, averaged
within different latitude bands, are statistically analyzed and com-
pared. The objects used for the analysis are the two-dimensional
spatial-temporal spectral and correlation characteristics, the multi-
variate autoregressive and linear regression model parameters, and
the diffusion equation coefficients. A comparison shows that, gener-
ally, the shapes of the corresponding spectra and correlation func-
tions are quite similar, but their numerical values and some features
differ markedly, especially for the tropical regions. The spectra re-
veal a few randomly distributed maxima (along the frequency axis),
the periods of which were not identical for both types of data. A
comparative study of the estimates of the diffusion equation coef-
ficients shows a significant distinction between the character of the
meridional circulations of the observed and simulated systems.
The approach developed gives approximate stochastic models
and reasonable descriptions of the temperature processes and fields,
thereby providing an opportunity for solving some of the vital prob-
lems of theoretical and practical aspects surrounding validation, di-
agnosis, and application of the GCM. The methodology and results
presented make it clear that formalization of the statistical descrip-
tion of the surface air temperature fluctuations can be achieved by
applying the standard techniques of multivariate modeling and multi-
dimensional spectral and correlation analysis to the data which have
been averaged spatially and temporally.

266
7.1 Objectives 267

7.1 Objectives
The idea of the statistical approach to the problems of GCM variabil-
ity validation is contained in the comparison (observed vs. modeled)
of the probability distributions of the different atmospheric and ocean
processes and fields. At first, such a statement sounds like a standard
statistical approach, and its solution would be obvious and simple if
the number of climate processes taking place jointly were not huge
and if they did not present a tremendously complicated (in its in-
terrelationships and feedbacks) deterministic-stochastic system. As
is known, the Stochastic System Identification Theory (see Eikhoff,
1983) deals mostly with the methodology for identifying linear sys-
tems, The interdependences of climatic processes and fields are not
linear, and the application of this theory can give only highly approx-
imate results. But it seems that such results can present very helpful
information for many areas of climate studies and can also serve as
an important step in the right direction for generalizing the approach
to and solution of the GCM confirmation problems. Moreover, on
the first stage of GCM variability study, even rough theoretical (sta-
tistical) approach is a principle way to refrain from some empirical
validation schemes which are not recognized by the statisticians, and
the results of which are impossible to compare statistically. Further
developments in this direction will help us to define more accurately
the details of such approaches and to identify appropriate observa-
tional structures.
This study is an attempt to compare the variability of one of
the most important climate processes, surface air temperature, for
the two complicated stochastic systems: real and modeled. As test
statistics of such a comparison, the estimates of the corresponding
second-order moments (or their functions) are used.
As is well known, some studies (e.g., Katz, 1992; Oort, 1983; San-
ter et al., 1990; Santer and Wigley, 1994) have provided a comparison
of the means and standard deviations of different atmospheric char-
acteristics derived from the observed and simulated data. In our
study, the spatial temporal climatic variability of the fluctuations
(the deviations from the means) of the surface air temperature pro-
cesses and fields is considered. Without studying and statistically
modeling such deviations, it is impossible to make any statements
about climate change (to forecast, to reason about the warming or
cooling of the atmosphere, and so on) because in all these predictions
one is reasoning about a deviation from the mean climatic regime.
In the lights of the results provided it will become clear that uti-
lization of the GCMs for different physical experiments without any
268 7 The GCM Validation

variability validation is not acceptable. But this is the situation on


this moment: hundreds of papers have been published (and continue
to release) where the GCMs with unknown variability are used for
the analysis of the real climate variability (climate sensitivity, green-
house effect, reconstruction of paleoclimates, El Nino study, and so
on).
When comparing the second moments, one can use the estimates
of the correlation matrices or the parameter matrices of the mul-
tivariate All models as standard test statistics. Additionally, the
parameter matrices as the matrices of interactions and feedbacks
present interesting diagnostic information about both systems.
The methods in Chapter 3 and 5 have obviously illustrated ways
in which to design and develop simple multivariate climate models,
as well as ways in which to estimate the multidimensional spectra
and correlation functions of different climatological structures. In
this chapter these methodologies are used not only for estimation
of the second-moment climate statistics, but also for conducting a
sequence of their comparative studies by calculating some physically
interpreted characteristics (the coefficients of the diffusion equation).
Consequently, then, this study will help us to answer this important
question: To what extent can GCM describe real large-scale climate
variations?
In short, this study presents the estimates and the results of a
comparison of a complex of the multiple second-moment statistics of
the observed surface air temperature and the 1000 mb air tempera-
ture simulated by the Hamburg GCM.

7.2 Data
There are many ways in which to represent the air temperature in
the form of time series and fields by spatial-temporal averaging and
filtering. Because our first objective is to compare the climate statis-
tics (or the statistics of the large-scale fluctuations), the following
two different structures of data are used. The first is the 31-year
zonally averaged monthly air temperature time series of the North-
ern Hemisphere (NH) at the 1000 mb isobaric level simulated by one
of the Hamburg coupled global atmosphere-ocean general circulation
models, sec Cubasch et al. 1992, Max Plank Institute for Meteo-
rology, Hamburg), The Hamburg GCM has described in details in
many papers, for example, in Roeckner, 1989; Santer and Wigley,
1990; Santer et al., 1994. The temperature time series, available for
various 5.625° width latitude zones, were additionally averaged for
7.2 Data 269

each pair of the adjacent latitude zones. As a result, eight time series
were obtained for each of the eight latitudinal bands of 11.25° width.
This structure can be considered as eight variate time series (or
two-dimensional latitude-temporal field)

where i and t are the latitude and temporal subscripts respectively.


The source of the observed data is the United Kingdom's global
surface air temperature observations for 1891-1990 (Jones et al.,
1986), archived at the NCAR The sample, used in this study, presents
surface air temperature of the NH for 31 years (1959-1989) aver-
aged within the same latitude bands of 11.25° width. The latitude-
temporal field of these eight time series have the same structure as
(7.1).
For ease of reference, Table 7.1 enumerates the time series of
different latitude bands and identifies their boundary values.
Before the spectral analysis and multivariate modeling, the an-
nual cycles [Ti(t) ) and Si(t)]] in the mean and variances were removed
and the time series of the normalized observations,

were analyzed.
The standard methodology (periodogram —> smoothing —> spec-
trum and periodogram—* Fourier transform —> correlation function)
with a two-dimensional Tukey spectral window (see Section 1.11) was
employed. Because the values of the periodograms (and the spectral
estimates) are small, they were multiplied by the factor (1000) to be
conveniently presented by illustrations. As a result of the estimation,
the normalized sample spectral density and its standard deviations
increased 1000 times will be analyzed systematically.
The observed and simulated sets compared are not completely
identical climatic structures: The first is the surface observations
and the second is the 1000 mb simulated data. As will become clear
shortly, these facts are of no import, at least at this stage of the
study.
But in principle, this discrepancy restricts the possibility of mak-
ing the quantitative statistical inferences (by hypotheses testing based
on corresponding test statistics) about parameters of the compared
distributions, and it constrains us to carry out a basically qualitative
study. If this approach and the form of the results representation
270 7 The GCM Validation

Table 7.1: Estimates of the linear trends /^(oC/year) and t-statistic


values for the observed and simulated data.

Simulated
Observed data data

Time Latitude 100 years 31 years 31 years


series band
number (degrees) A /, Pi / Pi t

1 90.00-78.75 .024 10.4 .005 0.8 .008 0.9


2 78.75-65.50 .009 8.4 .1)09 1.7 .010 1.7
3 67.50-56.25 .007 8.5 .012 2.6 .004 0.9
4 56.25-45.00 .006 9.9 .010 3.0 -.0033 -1.1 1
5 45.00-33.75 .005 13.6 .000 0.2 -.0077 -4.0 0
6 33.75 -22,50 .005 24.1 .001 1.3 .004 -3.0 0
7 22. 50- -11. 25 .005 23.9 .006 4.6 -.003 3 -2.99
8 11.25-00.00 .004 16.7 .006 4.7 .000 0.4

proves acceptable to the scientific community involved in climate


analysis and modeling, it will be possible to carry on a special exper-
iments for purposefully recollecting both kinds of data on the iden-
tical level to apply more rigorous statistical inferences in multiple
comparison and hypotheses testing.
To get a comparative representation about the long-term vari-
ability of the two 31-year data structures with respect to the entire
interval of 100 years of the surface air temperature observations, the
estimates of the parameter fi\ (of the simple linear regression lines
flo + Pit, which approximate each of the latitudinal time series) were
found together with corresponding t-statistic values (Table 7.1). Al-
though for the observed data almost all of the /3} estimates are posi-
tive, the linear trends for 31 years of observations are not as persistent
as for the entire 100-year interval. For the latter, each of the eight /3[
is statistically significant (t-statistic values are from 8 to 24), while
for the former, at most four of them are statistically significant, with
smaller t-statistic values lying in the interval (2.6, 4.7). As for the
7.3 Zonal Time Series 271

simulated data, four of the fti estimates are negative, and at most
three of them are statistically significant. So quantitatively as well as
qualitatively, there is no strict consistency in these simple character-
istics of the long-period fluctuations. But one remark must be made
here: The scales of variations of the fl\ estimates are approximately
the same for both types of data (less than about 1°C per 100 years).

7.3 Zonal Time Series


When one has a monthly time series of observations, it is possible
(in contrast to the traditional univariate technique) to apply a two-
dimensional methodology of spectral and correlation analysis. The
time series has to be presented in table (field) form with 12 observa-
tions in each row and with the number of rows equal to the number
of years (31) of observations (see Subsection 3.8.3). In our case, the
time series for each latitudinal band is presented as a two-dimensional
field of the 31x12 size. Such an approach makes it possible to sepa-
rate the interannual and intraannual fluctuations and to obtain their
description more expressively.
Before considering the results, the following remark must be made:
We will begin with a descriptive consideration of the estimates, men-
tioning their peculiarity (e.g., possible spectral maxima) as a reflec-
tion of the sample variability. After that the speculation about the
possible character of the spectra will be provided. Then, the statis-
tical inferences about the equality of the correlation coefficients will
be done.
Figure 7.1 gives estimates of the two-dimensional spectral den-
sities of the observed and simulated time series for each latitudinal
band.
Visual comparison of the pictures can lead to the preliminary
conclusion that there is some general similarity in their shapes. The
intraannual parts of them for the monthly frequencies have specific
bell-shape graphs with maxima corresponding to the zero monthly
frequency, which are expressed by the crest spreading along the an-
nual frequency axis. The heights of these maxima (the heights of
the crests) vary for different annual frequencies. The most power-
ful monthly fluctuations correspond to the periods of more than six
months. Fluctuations with periods of less than 3 or A months have
the least power. So the shape of the intraannual part of the spectra
seems to be very close to the shape of the spectra of the first-order
univariatc autoregressive process. It clearly shows that for periods
of more than a year, the long-scale spatial fluctuations predominate
272 7 The GCM Validation

Figure 7.1: Two-dimensional spectra of the observed (left) and simulated (right)
surface air temperature time series of different latitude bands. (Correspondence
between the picture number and the latitude band boundary is given in Table
7.1.)
7.3 Zonal Time Series 273

Table 7.2: Standard deviations of the spectral estimates and periods


(years) of the spectral maxima within the annual frequency interval
(see Figure 7.1) of the time series of different latitudinal bands.

Time Observed data Simulated data


series
number Standard Periods Standard Periods
deviations of deviations of
s1 s2 maxima si s2 maxima

1 0.10 0.41 3.2 0.11 0.53 3.2; 8.


2 0.12 0.52 3.6 0.12 0.58 2.4; 5.3
3 0.14 0.48 3.2 0.12 0.68 5.3
4 0.15 0.49 5.3 0.12 0.40 6.2
5 0.17 0.90 5.3 0.14 0.72 2.3
6 0.31 0.44 oo 0.15 0.38 2.9
7 0.31 0.93 3.6 0.15 0.59 8.0
8 0.31 1.95 3.6 0.17 0.80 2.9; 5.9

in the variability of the surface air temperature. Therefore, the most


powerful part of the spectra is the crest (for zero monthly frequency),
with several maxima. This is a brief description of the similarity of
the spectra.
The details of the spectra are quite different. First, the interan-
nual maxima periods are different for different latitude bands (their
numerical values are given in Table 7.2).
For the observed data there are about three clearly distinguish-
able latitudinal regions, with the periods of the maxima being about
3.6 years for the tropics, 5.3 years for the midlatitudes (35° to 57°),
and 3.2 to 3.6 years for the northern and Arctic regions. Roughly
speaking, this means that within each of these three regions, the
spectral structure of the observed temperature fluctuations can be
considered as approximately identical.
The specific frequency composition of the simulated data differs
from that of the observed data: The periods of the maxima (see Fig-
ure 7.1 and Table 7.2) vary randomly for the time series of different
274 7 The GCM Validation

Table 7.3: Correlations (see Figure 7.2) for the first two lags (months)
and corresponding N statistic values.

Time Observed data Simulated data


series N
number Cor relations Correlations
1 2 1 2 1 2

1 0.16 0.02 0.15 0.12 0.1 1.2


2 0.20 0.12 0.25 0.05 -0.7 7 0.9
3 0.30 0.1.9 0.28 0.10 0.3 1.1
4 0.35 0.20 0.32 0.12 0.4 1.0
5 0,32 0.15 0.41 0.16 -1.4 4 -0.1 1
6 0.55 0.41 0.40 0.13 2.5 3.8
7 0.76 0.65 0.44 0.23 6.8 6.7
8 0.82 0.68 0.43 0.25 9.1 7.1

bands. There is no noticeable separation of the statistical structure


of the atmosphere into three regions, as was seen for the observed
data.
The principal question which arises in the process of interpret-
ing the maxima is that of their statistical significance. As is known
(3.9), in general, the accuracy of the spectral estimates is different
for different frequencies because the variance of such estimates is
proportional to the corresponding squared value of the spectra. The
greater the spectral value, the greater the error of its estimate can
be. To get some representation about accuracy, approximate values
of the standard deviations were found with the aid of the residuals:
periodogram —spectral estimates. The value of such a standard de-
viation depends on the region of the frequency domain over which
the squared residuals are summed up. For each spectrum in Figure
7.1, the two standard deviations, si and s2, are given in Table 7.2:
s1. was computed with the aid of the residuals for all the points of
the frequency plane, while s2 was found only with the aid of the
residuals along the major crest of the spectrum estimates. Like the
other characteristics analyzed here, these standard deviations are
7.3 Zonal Time Series 275

latitude dependent, with greater values for the tropical bands where
the spectral estimate variability for the observed data is about two
times greater than that for the simulated data. By comparing the
standard deviations in Table 7.2 with the heights of the maxima in
Figure 7.1, it is easy to notice that the estimates of the crest values
are certainly statistically significant (for instance, at the 95% signif-
icance level). The question about the statistical significance of the
separate maxima with respect to the mean level of the corresponding
crest, however, is more complicated. Although some low-frequency
maxima of the tropical bands are certainly significant, at least at
the 70% level, it is likely that for the rest part of the interannual
frequency interval and for the time series of the nontropical regions,
they are simply the result of the sample variability. This means
that for the considered period of observations, the variability in the
vicinity of these maxima was certainly slightly higher than for other
frequencies; however, the possible theoretical spectral density of the
corresponding processes exhibits no such features. Therefore, the es-
timates in Figure 7.1 demonstrate two kinds of features: The first is
the specific shape of the random process spectrum under study with
only small maxima in the vicinity of the zero frequency; the second is
the specific character of the observed sample Fourier decomposition.
Of course, it is possible to increase the width of the two-dimensio-
nal filter used and to get rid of the most of the maxima that we
can see in Figure 7.1. But the empirical study of some spectral
features of the analyzed samples, which represent real and simulated
31 year climate, can be useful for the comparison. For example,
having the results in Figure 7.1, one can conclude that the separate
observed climate fluctuations are significantly more powerful than
the simulated ones, especially in the tropical regions.
In both cases, the above-mentioned periods of the spectral max-
ima are randomly distributed along the frequency axis. But it seems
that for the observed data, the low frequency maxima are the re-
sults of some systematic long-period forcing (may be of the CO2 or
El Nino type), whereas for the simulated data the maxima are the
results of the nonlinear transformation [of the amplitude modulation
type (Polyak, 1975)], which occurs in the process of the numerical
integration of the hydrodynamic equations. The most important
distinctions of the spectra for the two types of data are in the white
noise component level, which can be seen in the illustrations as the
distance between the minimal spectral estimates and the plane with
zero values. (Because normalized data is considered, none of the dis-
crepancies in the level of this component are absolute values; they are
276 7 The GCM Validation

respective to the values of the corresponding standard deviations.)


For both systems, the white noise component increases as the lat-
itude increases, but the starting level (at the equatorial region) is
significantly greater for the simulated data.
Therefore, Figure 7.1 shows the "whitening" of the temporal tem-
perature fluctuations in the process of spatial northward transport.
The value of the white noise component and the character of its
variation from one latitude band to another are different for both
types of data, but the observed spectra have one very specific addi-
tional feature: a dramatic change of the frequency structure on the
boundary (35°) between the tropics and the middle latitudes (see
Figure 7.1, 6), where the heat balance is reached. For this latitude
band (Figure 7.1, 6), the powerful maximum formed in the origin
shows that something unusual abruptly occurred in the frequency
composition of the temperature fluctuations on the boundary of this
region. Nothing like this appears in the simulated spectra, where the
latitudinal variation of the spectral structure is gradual and smooth.
The shapes of both types of the two-dimensional correlation func-
tions corresponding to the spectra in Figure 7.1 do not change mark-
edly from one latitude band to another, but their values vary sig-
nificantly (see Table 7.3). These shapes are approximately similar
to those in Figure 7.2, which provides examples of such correlation
functions in the 56.25° to 45° latitude band. The intraannual cor-
relations are non-zero only for up to 4 monthly lags, while the in-
terannual statistical dependence of the monthly fluctuations is close
to zero. Therefore, the main feature of the shape of such correla-
tion functions is that they are almost flat (nonzero correlations are
concentrated in the narrow strip about the origin of the annual lag
axes).
The distinction between the correlation functions is reflected in
the different character of their latitudinal variation, which is illus-
trated by the trend of the first two correlations, given in Table 7.3.
For the observed data, the first correlations decrease from 0.82 in
the equatorial region to 0.16 in the Arctic, while for the simulated
data such a decrease is from about 0.43-0.44 to 0.15. There is a dra-
matic change (from 0.76 to 0.32) in the observed correlation value in
the sixth (33.75° to 22.5°) latitude band for the transition from the
tropics to the midlatitudes. This change in the span of the narrow
11.25° latitude area is greater than the total latitudinal variations of
the simulated correlations.
For the observed data, the noise level changes from almost zero
for the tropics (where there are especially powerful low frequency
maxima) to the significant level of about 80% of the total variability
7.3 Zonal Time Series 277

Figure 7.2: Two-dimensional correlation functions of the observed (bottom)


and simulated (top) time series of the fourth latitude band.
278 7 The GCM Validation

of the fluctuations in the Arctic area. The noise component for the
simulated data is significantly high for the tropical regions (about
40% of the variability) and approximately the same (as for the ob-
served data) for the Arctic. Therefore, the character of whitening
(that is, increasing the white noise component in the data of the
northern regions) of the long-period temperature fluctuations, ex-
pressed as a latitudinal trend of the statistical structure, is the most
distinctive feature of the temperature fluctuations, and it is quite
different for both kinds of data.
The abrupt change in the statistical structure of the observed
fluctuations on the boundary between the tropics and the middle
latitudes corresponds to the theoretical understanding of the merid-
ional types of the atmospheric circulation (see, for example, Palmen
and Newton, 1969, where Figures 1.5, 1.6, 1.11, and some others show
a specific character of the tropical and midlatitude mass circulation
in the NH and the boundaries between them for different seasons).
Thus, it is not unusual that this character has been reflected in the
observed second-moment statistics; what is really unusual is that it
is not clearly distinguishable in the simulated data.
To compare the correlations more accurately let us consider a
test statistic (Rao, 1973)

(where r1 and r2 are the sample correlation coefficients and n\ and


n2 are the sizes of the samples), of a simple asymptotic criterion for
testing the equality of the two correlation coefficients. If n1 and n2
are large, then N has an approximately normal distribution.
The N values for the first two correlations (Table 7.3) show that
the difference in the correlations compared are statistically significant
(at 95% level) for the three tropical bands (values of N are from 2.5 to
9.1). As for the other five bands, such differences are not statistically
significant.
These results show that the temporal variability in the middle
and northern regions of the observed and simulated data is quite
similar or, more accurately, the spectral and correlation structures
of these two types of temperature fluctuations for these regions are
statistically undistinguished. But such variability in the tropics and
on the boundary between the tropics and midlatitudes is different.
In this region the abrupt change of the observed temporal statistical
structure is accompanied by the slow and gradual variations of the
analogous simulated statistics.
7.4 Multivariate Models 279

7.4 Multivariate Models


Evaluation of the similarities of or distinctions between both types
of second-moment climate statistics is also possible through the in-
tercomparison of the parameter estimates of the multivariate autore-
gressive models fitted to their multiple time series. The AR stochas-
tic models are an approximation of the stationary time series by a
system of linear differential equations (more exactly, by correspond-
ing difference equations); or, put another way, they are a numerical
linearization of a system under study. This approach proves very
interesting because the linear approximations of the analogous ob-
served and simulated data structures must be very close if the sam-
ples are from the same distribution. But to build a multivariate
AR model, one needs sufficiently long multiple samples. The vol-
ume of the data (372 points in each monthly time series) is enough
to consider only the first-order multivariate AR models. It is also
important to study the multivariate AR climate models, for they
offer an alternative approach to the climate system description and
to the analysis of the interactions and feedbacks of the atmospheric
processes.
A description of an algorithm used for building first-order mul
tivariate autoregressive models is given in Section 5.2. The dataset
of the normalized anomalies for the eight fixed latitude bands (see
Table 7.1) is assumed to be a multiple (eight-variate) time series with
372 temporal terms.
The estimates of the autocorrelation and cross-correlation coeffi-
cients for the first two temporal lags of the multiple time series are
presented in Tables 7.4 and 7.5. For the observed data, the matrix
correlation function estimates for T — 0 (the spatial correlation field)
show that when the lags are equal to about 1260 km (one latitude
zone separation) there is a noticeable spatial statistical dependence
with correlations from 0.54 to 0.83.
If spatial lags greater than 2520 km, most of correlation estimates
are not statistically significant. For a temporal lag equal to 1 month
(Table 7.5), it is possible to see the crest along the main diagonal with
the correlation maximum for the equator (in the lower right corner).
The smallest value (0.15) of the autocorrelations is in the Arctic
region, while the largest (0.93) is in the equator. As for the analogous
simulated data estimates (Tables 7.4 and 7.5), spatial correlations for
the 1260 km lag are sufficiently fewer (from 0.22 to 0.39) as well as
autocorrelations for the one-month lag in the tropical region.
To make accurate statistical inferences, one has to compute the N
statistic values for the corresponding correlations in Tables 7.4 and
280 7 The GCM Validation

hation functions for the zero lag (r = 0).forthe zero lag (T=o).

|| Ml | «2 | «3 | «4 MS M6 U7 M8
Observed data
MI 1.00 0.68 0.31 -0.10 0 -0.21 1 -0.16 6 -0.033 -0.04
«2 0.68 1.00 0.79 0.14 -0.13 3 -0.07 7 0.07 -0.011
«3 0.31 0.79 1.00 0.61 0.07 -0.03 3 0.05 0.00
M4 -0.10 0 0.14 0.61 1.00 0.61 0.133 -0.033 0.00
«5 -0.21 1 -0.13 3 0.07 0.61 1.00 0.54 0.03 -0.022
M6 -0.166 -0.077 -0.03 3 0.13
3
0.54 1.00 0.67 0.41
"7 -0.03 3 0.07 0.05 -0.03 3 0.03 0.67 1.00 0.83
«8 -0.044 -0.01 1 0.00 0.00 -0.02 2 0.41 .083 1.00
Simulated data
MI 1.00 0.22 -0.01 1 -0.044 -0.07 7 r 0.03 0.03 -0.03
u-2 0.22 1.00 0.39 -0.01 1 -0.20 0 0.03 0.09 0.00
U3 -0.01 1 0.39 1.00 0.27 -0.17 7 0.00 0.05 -0.05
(«4 -0.04 4 -0.01 1 0.27 1.00 0.27 0.09 0.06 0.00
M5 -0.07 7 -0.200 -0.17 7 0.27 1.00 0.31 0.02 0.01
W6 0.03 0.03 0.00 0.09 0.31 1.00 0.24 0.12
U7 0.03 0.09 0.05 0.06 0.02 0.24 1.00 0.25
us -0.03 3 0.00 -0.055 0.00 0.01 0.12 0.25 1.00

7.5. Such values show that the difference in the spatial correlations
(see the first matrix in Table 7.6) when the lag is about 1260 km
are statistically significant at any standard level of significance. The
largest value (12.7) of N corresponds to the equatorial band. This
comparison shows that the spatial variability of the observed and
simulated climate are sufficiently different.
As for the temporal-spatial correlations, the N statistic values
(the second matrix in Table 7.6), show the statistically significant
difference for the three tropical bands as well as for some of the
midlatitude bands.
In spite of a certain qualitative similarity between the latitudi-
nal trend of the estimated correlations (for the one-month lag) for
both kinds of data, it is obvious that the values of the corresponding
correlations in Tables 7.4 and 7.5 differ markedly, which means that
the observed and simulated samples considered are from different
distributions.
Figure 7.3 (bottom) illustrates with special clarity the fundamen-
tal distinction in the observed statistical structures (matrix correla-
tion function from Table 7.5 for the one-month lag) of the tropical,
7.4 Multivariate Models 281

Table 7.5: Matrix correlation function for the first lag (r = 1 month).

II "1 «2 «3 «4 «5 «6 U7 «8
Observed data
MI 0.15 0.08 0.06 0.06 -0.02 -0.02 -0.01 -0.05
U2 0.12 0,17 0.20 0.20 0.11 0.07 0.06 -0.05
us 0.09 0.19 0.30 0.38 0.24 0.13 0.05 -0.04
U4 0.08 0.16 0.23 0.41 0.38 0.17 0.02 0.01
U5 0.01 0.09 0.08 0.19 0.36 0.27 0.06 0.01
U6 -0.07 7 0,10 0.11 0.05 0.18 0.61 0.56 0.38
U7 -0.09 9 0.03 0.05 0.00 0.03 0.56 0.84 0.74
us -0.088 -0,03 0.02 0.06 0.01 0.42 0.80 0.93
Simulated data
u\ 0.17 0,15 0.00 0.04 0.04 0.02 -0.06 -0.07
U2 0.11 0.23 0.14 0.11 -0.144 -0.02 0.04 -0.03
U3 0.06 0 17 0.28 0.25 -0.04 0.06 0.06 0.00
U4 0.02 0.01 0.04 0.32 0.26 0.09 0.11 0.08
U5 -0.05 -0.06 0.01 0.00 0.42 0.18 0.06 -0.02
U6 0.02 0.00 -0.01 0.08 0.20 0.43 0.16 0.07
U7 -0.07 -0,01 0.08 0.09 -0.02 0.09 0.49 0.23
U8 0.03 0.03 0.05 0.06 -0.03 0.11 0.18 0.49

midlatitude, and polar regions with abrupt and significant changes on


the boundaries, which separate three major zones of the meridional
atmospheric circulation (Oort and Rasmussen, 1971; Oort, 1983; Pal-
men and Newton, 1969).
An analogous picture for the simulated data in Figure 7.6 (top)
shows the gradual trend of the correlation structure, whose values in
the tropical region are about two times less.
The parameter estimates (exceeding the 99% significance level) of
eight- variate AR and linear regression models for both kinds of data
are given in Tables 7.7 and 7.8. Each row in these tables presents an
equation in finite differences (for the autoregressive model) or a linear
dependence (for the regression model) for the normalized anomalies
«,-(<)•
The standard error ei+1 1 for the autoregressive equation presents
the normalized standard deviation estimate of the forecast for one
month ahead as a proportion of the standard deviation of the original
anomalies.
282 7 The GCM Validation

Figure 7.3: Matrix correlation functions for the one month lag of the observed
(bottom) and simulated (top) temperature time series.
7.4 Multivariate Models 283

Table 7.6: N statistic values for the comparison of the correlations


in Tables 7.4 and 7.5.

MI~T w2 «3 U4
r=0 (month)
U5 U6 UT U8

Ui — 8.2 4.5 -0.88 -1.9 9 -2.66 -0.88 -0.1 1


U2 8.2 — 9.0 2.1 1.0 -1.44 -0.3 -0.1 1
uy, 4.5 9.0 — 5.9 3.3 -0.44 0.0 0.7
U4 -1.99 2.1 5.9 — 5.9 0.6 1 ') 0.0
L • £j

U5 -1.99 1.0 3.3 5.9 — 3.9 0.1 -0.4 4


U6 -2.6 6 -1.44 -0.4 0.6 3.9 — 7.7 4.3
UT -0.88 -0.33 0.0 -1.22 0.1 7.7 __ 12.7
U8 -0.1 1 -0.1 1 0.7 0.0 -0.44 4.3 12.7 —
r—l (month)
Ui -0.33 -1.00 0.8 0.3 -0.8 8 -0.55 0.7 0.33
U'2 0.1 -0.88 0.8 1.3 3.4 1.2 0.3 -0.33
V-3 0.4 0.3 0.3 2.0 3.9 1.0 -0.11 -0.55
U4 0.8 2.1 2.6 1.4
4 1.8 1.1 -1.22 -1.00
U5 0.8 2.0 1.0 2.6 -1.0 0 1.3 0.0 0.4
u6 -1.22 1.4 1.6 -0.44 -0.33 3.4 6.4 4.5
U7 -0.33
0.5 -0.44 -1.2 2 0.7 7.4 9.3 9.7
U8 -1.55 LZ°-8 -0.44 0.0 0.5 4.6 12.4 15.2 5.2

Even a brief examination of the estimates in Table 7.7 of the


model parameters for the observed data reveals that the number
of statistically significant values above the main diagonal is greater
than the number of those below the main diagonal. In other words,
the temperature anomaly values are mainly determined by their own
past and by the anomaly values of the more southern bands (north-
ward transport of heat), We have already found this feature, while
considering the example in Subsection 5.5.2. The observed tempera-
ture fluctuations in the moment of t -f 1 of the middle and northern
latitude bands depend on almost all southern bands' anomalies (the
spreading of the atmospheric temperature anomalies takes place, on
average, northward; the southward transport is weaker because most
of the corresponding parameter estimates are not statistically signif-
icant). The more or less regular change of the signs of the parameter
284 7 The GCM Validation

estimates in Table 7.7 shows that some kind of stable spatial waves
are responsible for the northward temperature fluctuation transport.
Therefore, the approximation of the multiple time series by the mul-
tivariate All model has the potential to detect some regularities in
the second-moment structure in spite of the significant variability
and noise associated with the fluctuations.
For the simulated data in Table 7.7, the temporal dependence
is determined primarily by the past of the time series itself. This
means, firstly, that the spatial-temporal statistical dependence of the
temperature fluctuations of different latitude bands is very low, and,
secondly, their statistical structure is simple and can be presented by
a univariate first order autoregressive model.
In tropical regions, the approximation accuracy (£t+i) for the ob-
served data is much higher than the analogous accuracy for the sim-
ulated data, and the corresponding' forecasts for a couple of months
ahead proves relatively accurate, while such forecasts for the cor-
responding simulated time series do not make sense for more than
one month ahead. The forecast in the polar and midlatitude regions
proves virtually useless for both types of data.
The linear regression model parameters in Table 7.8 also indicates
a significant difference in the character of the spatial dependence of
the compared anomalies. For the observed data, each anomaly is
determined by the values on two to four closest latitudinal bands from
both sides, while for the simulated data such spatial dependence can
be seen mainly on the values of one adjacent band from both sides.
For any of the latitudinal bands, the accuracy of the linear regression
approximation of the observed data is significantly higher (s is from
0.23 to 0.55) than for the simulated data (e is from 0.85 to 0.97). The
observed zonal monthly temperatures considered can be accurately
interpolated by the linear combination of the observations for the
same month of several adjacent latitude bands, while the spatial
interpolation of the analogous simulated data cannot be accurately
done at all.
Therefore, comparison of the estimated parameters of the mul-
tivariate models showed that the temporal and spatial statistical
dependence of the observed data is greater than that of the simu-
lated data, especially in the tropical regions. This discrepancy clearly
shows a fundamental distinction between the statistical qualities of
the temperature fluctuations for the two kinds of data, which means
that both multiple samples are from different distributions.
7.5 The Diffusion Process 285

Table 7.7: Estimates (exceeding the 95% significance level) of the pa-
rameters of the eight-variate autoregressive models for the latitude-
temporal data on the 31-year interval (e is the normalized standard
error).

Observed data
«li U'2t «31 "4* «5* «6* «7« "8* £
MK+I — 0.24 -0.31 0.34 -0.17 7 — -0.122 — .98
«2J+1 --- 0.44 -0.34 0.42 -0.288 0.33 0 -0.177 ._.. .96
W3t + l 0.15 0.25 -0.288 0.41 -0.32 2 — .93
«4t + l — — 0.27 0.28 — 0.19 -0.33 3 0.25 .89
U5t + l — — — 0.19 0.19 — — — .91
"6*4-1 ._ 0.19 -0.122 0.54 0.12 — .75
Ujt + l — — — — 0.13 0.45 0.37 .50
"8!+l — — — — — — -0.16 6 1.03 .35
Simulated data
Ult + l 0.16 — — — ----- .98
»2i+l — 0.17 — — — — .96
u-.it + 1 — 0.30 — — — .95
«4*+l — - 0.14 0.29 — — ._.. — .92
WSH-I — — 0.16 0.34 — — — .88
Met+i — — — 0.41 ._.. — .89
«7H-1 — — — _.. — 0.46 — .86
W8* + l — — — — — — 0.46 .86

7.5 The Diffusion Process


The observed temperature fluctuations are most strongly dependent
(temporally) upon their own past as well as upon the fluctuations
of the southern regions' past since the computed matrix of the pa-
rameters has a special form with mostly non-zero elements above the
main diagonal (Table 7.7). This shows the northward climatic (with
the monthly temporal scales) meridional transport of temperature
anomalies. To present an approximate physical description of this
transport let us assume that one of the appropriate models for such
a study can serve the diffusion process. According to Monin and Ya-
glom (1973, p. 580), the advection velocity practically coincides with
the instantaneous flow velocity, and obviously, it cannot be greater
than the How velocity.
286 7 The GCM Validation

Table 7.8: Estimates (exceeding the 95% significance level) of the pa-
rameters of the eight-variate linear regression models for the latitude-
temporal data on the 31-year interval (e is the normalized standard
error).

Observed data
Ml M2 MS M4 MB «a 117 M8 e
m -- 1.69 -1.47 7 0.83 -0.49 9 0.19 -0.211 — .55
«2 .35 - 1.03 -0.644 0.35 -0.177 0.14 .25
«3 -.255 0.83 0.67 -0.37 7 0.17 — .23
M4 .25 -0.922 1.19 0.63 -0.32 2 0.18 — .30
«5 -.244 0.83 -1.08 8 1.03 -- 0.69 -0.50 0 — .39
U& — -0.477 0.57 -0.61 1 0.79 0.91 -.322 .42
UT 0.27 0.24 0.24 -0.400 0.63 — .56 .35
Us — -0.13 3 — -- 0.18 -0.466 1.14 — .50
Simulated data
MI — 0.26 -0.133 — .97
11 2 .21 — 0.39 — .88
MS 0.37 0.32 -0.20 0 — .85
«4 — — 0.36 0.32 — .90
MS -— 0.21 0.30 — 0.31 ----- .87
M(i ----- 0.34 0.21 — .9]
«7 0.22 — .23 .94
Us -- -- — 0.24 .96

Before the corresponding methodology is applied, the following


remarks must be made. The character of the temperature advection
under study is determined by many factors; for example, the direc-
tion and strength of the wind. If, for example, one is interested in the
mean advection velocity of the synoptic scales fluctuations associated
with the eastward wind, one must collect and analyze temperature
observations that correspond to the kind of wind chosen. Therefore,
careful planning is needed to determine what kind of advection must
be evaluated. Given this situation, it was decided first to evaluate
(where possible) the mean northward advection velocity for each lat-
itude band, and then to find its mean value for the entire Northern
Hemisphere.
One of the ways to explore the climatic diffusion, of the tempera-
ture fluctuations and to estimate the advection velocity is to find the
relationship between the AR models and a diffusive description of
the temperature fluctuations (Polyak et al., 1994). Spatial averaging
7.5 The Diffusion Process 287

of data reduced dimensionality and brought the temperature fields


to the coordinate system (x, t). Although this is the only case we will
consider here, our reasoning will hold for a field of any dimension.
Therefore, one can approximate each of the latitude-temporal
fields by the following diffusion equation:

However, a multivariate model is the system of difference equations


corresponding to a system of ordinary differential equations. In
order to exploit this relationship, the diffusion equation must be
approximated by a system of ordinary differential equations. The
straight — linemethod (Berezin and Zhidkov, 1965) achieves such an
approximation. If

then replacement of the derivatives O 2 u/Ox 2 and du/dx by the sim-


plest finite differences gives

This system of ordinary differential equations approximates the dif-


fusion equation and corresponds to the autoregressive equations of
the multivariate model.
Any other form of the difference representation for the deriva-
tives d2u/dx2 and du/dx leads to a different, but linear, system of
differential equations. The greater the order of the finite differences
used, the more accurate an approximation can be achieved.
Thus, estimating the AR model parameters offers one of the pos-
sible solutions for the inverse problem of these differential equations.
The relationship between the parameters of the autoregressive
model and the coefficients of the diffusion equation can be established
by replacing the derivative du/dt in (7.4) by the simplest difference.
We have

where / is the time step, and the parameters a,b,c are interconnected
with l,h,q,s, and v by the following relationships:
288 7 The GCM Validation

The autoregressive parameters (a,b,s) present appropriate samples


for the subsequent estimation of the coefficients of the diffusion equa-
tion. This estimation makes sense when there is a real diffusive pro-
cess in a given x direction, as is the case in the above spatial diffusion
of temperature.
Let us associate the finite difference autoregressive equations (Ta-
ble 7.7) for the obtained eight-variate AR models with the coefficients
of the diffusion equation through formulas (7.5-7.8). For example,
for the seventh time series (24° to 13° latitude band) in Table 7.7,
one has a = 0.13, b = 0.45, and c — 0.37. By substituting these val-
ues in the equations (7.6 7.8) and having in mind that l = 1 month
and h ~ 1260 km, the estimates

are obtained (only two significant digits are presented in the final
coefficient estimates). The negative value of v means southward ad-
vection.
The results for each of the AR equations in Tables 7.7 are given
in Table 7.9. In the calculations by formulas (7.6-7.8), both the sta-
tistically significant and insignificant estimates (which are not shown
in Table 7.7) of the AR model parameters were used. The means in
Table 7.9 were obtained using the means a, b, and c of the AR pa-
rameter estimates which were found by averaging the corresponding
values in Tables 7.7 on the main diagonal and on the two adjacent
diagonals parallel to the main diagonal.
The estimates in Table 7.9 reflect the distinction in the Northern
Hemisphere's three major atmospheric meridional circulation regions
(Palmen and Newton, 1969; Oort and Rasrnusson, 1971; Oort, 1983).
For these three regions, the sign of the estimated values of the ad-
vection velocities are different, thereby reflecting the changes of the
advection directions on the boundaries between the regions. The
very crude estimation of the boundary between the tropics and the
midlatitude circulation zone is 30-35° for the observed data and
7.5 The Diffusion Process 289

Table 7.9: Diffusion equation coefficients (s is in m2/sec; v is in


m/sec; q is in sec - 1 ) of the observed and simulated data for different
latitude bands.

Time Observed data Simulated data

series
s V q s V <l
number

2 -3622 0.12 0.0000004 2622 0.00 0.0000002


3 3102 -0.09 0.0000002 -1012 0.05 0.0000003
4 2402 0.18 0.0000002 1542 0.10 0.0000002
5 29522 0.05 0.0000002 2802 0.03 0.0000002
6 -44 -0.12 0.0000002 792 0.03 0.0000002
7 3912 -0.12 0.0000000 1712 - 0.01 0.0000002

mean 2092 0.00 0.0000002 1812 0.04 0.0000002

24—30° for the simulated data; between the arctic and the midlat-
itude circulation zone is 60—70° for the observed data and 70—80°
for the simulated data.
Let us compare our estimates with the corresponding results ob-
tained by Oort and Rasmussen (1971, pp. 71 and 228) using wind
observations. Their approximate latitude points of the wind sign
change (boundary values for the midlatitudc region of the merid-
ional atmospheric circulation) are 31° and 67° with the wind values
from —0.9 to 0 m/sec for the tropics, 0 to 0.5 m/sec for the mid-
latitude, and 0 to —0.2 for the arctic. Roughly speaking, therefore,
the advection velocity estimates in Table 7.9 for both kinds of data
are approximately within the limits of accuracy of the Oort and Ras-
mussen (1971) results.
Relating to the turbulent diffusion coefficient s, Monin and Ya-
glom (1973) assumed that its value has to be of the order about 1
to 10 m2/sec. Our results for both kinds of data are significantly
greater, which means that the theoretical and experimental study of
turbulence is a problem of great significance.
290 7 The GCM Validation

Generally, the numerical values of the advection velocities (as well


as other coefficients of the diffusion equation) of different latitude
bands for both types of data differ markedly. The estimate (—0.01
m/sec) of v in the tropics for the simulated data is many times less
in absolute value than the corresponding estimates for the observed
data, which means that the observed and simulated meridional at-
mospheric circulations in the tropics are very different. The reason
for this difference needs to be given special consideration, which can
be accomplished by averaging the data within latitude bands more
narrow than 11.25° and, perhaps, with a time step (of data) of less
than one month.
By accepting the above mean values in Table 7.9 for the observed
data as the diffusion equation coefficient estimates, one gets the fol-
lowing:

This equation describes the mean meridional diffusion of the tem-


perature fluctuations for the entire Northern Hemisphere obtained
using the 31-year interval (1959 1989) of observations.
The questionable point in the above analysis, especially in the
midlatitudes for the observed data, is the order of the finite dif-
ferences used. As the estimated parameters of the autoregressive
equations show (see Table 7.7), each value of ukt+1 must be a linear
combination of the three to four values of u in the moment of t, which
is not the case for the simplest finite difference schemes used in for-
mulas (7.4) and (7.5). But increasing the finite differences order in
(7.4) and (7.5) requires a special development of the methodology as
well as a corresponding analysis of the numerical integration schemes
used by the Hamburg GCM.
It is difficult to conclude from the above results what estimates
(observed or simulated) most accurately represent the mean bound-
ary values between the Ross by three-cell meridional circulation zones
(Palmen and Newton, 1969). However, the analysis of the temporal
variations of these latitude boundaries (which can be completed by
applying the above methodology) can serve as an important source
of information about climate change, and the 100-year set of sur-
face air temperature observations can be used in such a study as an
appropriate data base.
Comparison of the results obtained for the observed and simu-
lated data shows that in different latitude bands the monthly tem-
perature fluctuation advection velocities, as well as other coefficients
7.6 Latitude-Temporal Fields 291

of the diffusion equation, are markedly different. The absolute values


of the v estimates are greater for the observed data, so the real pro-
cesses of the transport of the climatic fluctuation are more powerful
than those described by the model.

7.6 Latitude-Temporal Fields


The eight zonal time series (observed or simulated) can be analyzed
jointly as a two-dimensional latitude-temporal field of the normalized
temperature anomalies of the NH.
But the two-dimensional spectral and correlation analysis method-
ology is valid only for the homogeneous random fields. The investi-
gations that were carried out in the previous sections showed tha,t
neither of the two latitude-temporal fields above is homogeneous.
Their statistical structure depends not only on the differences be-
tween coordinates of the grid points, but also on the values of these
coordinates; that is, their statistical structure is latitude-dependent.
The application of the statistical methodology, developed for the
homogeneous field (Chapter 3), to the nonhomogeneous field could
bring about some very general characteristics which would not corre-
spond to different, maybe approximately homogeneous, subdomams
of the entire field. Therefore, such an approach can serve only as
a very crude estimation of the real statistical properties, and in
this case the spatial-temporal spectral and correlation characteristics
could give only a rough, approximate description of the variability
of such a latitude- temporal field.
Keeping the existence of the three different meridional atmo-
spheric circulation regions of the NH in mind, one can try to overcome
the inconvenience related to the nonhomogeneity by analyzing data
on each such region; that is, by comparing the statistical structure
of the corresponding observed and simulated fields for each region
separately.
In addition to dividing the entire field into three different sub-
fields, it is especially helpful to have high spatial resolution. To
satisfy this requirement, we will consider the zonal temperature time
series for different 5.625°-width latitude bands.
The first two subfields to be compared will correspond to the trop-
ical region (0-28.125° latitude). The size of each subfield is 5[5.625°
(or 630 km) width latitude bands]x372 months.
The main distinctive features of the two-dimensional spectral es-
timates (Figure 7,4) are the low frequency maxima which are cen-
tered in the origin. The non-low frequency part of the simulated
Figure 7.4: Spectra of the observed (bottom) and simulated (top) latitude-
tomporal temperature fields of the tropical region.
7.6 Latitude-Temporal. Fields 293

spectrum looks more irregular than that of the observed one. The
observed spectrum is smoother because the number of feedbacks and
interdependences in nature are significantly greater than what are
included in the model; therefore, the power of separate randomly dis-
tributed pikes in the observed periodogram must be smaller, whereas
the smoothing procedure leads to more uniform estimates in this case,
The spectral maximum for the observed data is significantly more
powerful than that for the simulated data. This indicates a very low
level (the distance between the spectral surface and the horizontal
plane) of the white noise component for the observed data. There-
fore, it is possible that the observed long-period climate fluctuations
are caused by some natural forces that are not identically described
by the model. The low-frequency spectral maximum of the surface air
temperature, which is well-known from the analysis of the historical
records, reflects the possibility of the large-scale climate fluctuations.
(But for the historical records this maximum is not very powerful be-
cause most of the meteorological stations with the long observational
intervals are located in the middle latitudes.)
The existence of such a maximum (in spite of its small value) for
the simulated data means that, generally speaking, the GCM ren-
dered some climate features more or less realistically. Of course, we
are far from attempting to identify the physical nature of the max-
ima for both types of data: they must have a different power as well
as, for example, spectral maxima obtained using two sets of temper-
ature data given on two different observational intervals. But the
difference in the power, which is concentrated in the low-frequency
part of the spectra in Figure 7.4, is too large to be attributed only
to the sample variability.
The approximate values of the standard deviations of the spectral
estimates were obtained using the residuals: periodogram — spectral
estimate. The values of the standard deviations vary significantly
for different regions of the frequency—wave-number domain. For
example, for the entire domain of the spectral estimates of the ob-
served data (Figure 7.4, bottom), this standard deviation is about
0.06, while for its low-frequency part it is about 0.20. Analogous
values for the simulated data (Figure 7.4, top) are about 0.02 and
0.10. These estimates show that the low-frequency spectral maxima
of both types of data are statistically significant at the 95% level.
The distinction in the distributions of the observed and simulated
data becomes especially clear when one compares the correlation
functions (see Figure 7.5 and Table 7.10). Such a comparison reveals
that the numerical values of the estimates are absolutely different.
294 7 The GCM Validation

Figure 7.5: Correlation functions of the observed (bottom) and simulated (top)
latitude-temporal temperature fields of the tropical region.
7.6 Latitude-Temporal Fields 295

Table 7.10: Correlations (see also Figures 7.5 and 7.7) for the first
two spatial and temporal lags and corresponding statistics N.

Lags of the Observed Simulated N


correlations data data
Latitudes 90.00°-78.75°

1 month, 0 km 0.13 0.16 -0.4


2 months, 0 km 0.04 0.08 -1.6

0 month, 630 km 0.45 0.32 2.1


0 month, 1260 km 0.06 0.06 0

Latitudes 78.75°-33.75°

1 month, 0 km 0.26 0.29 1


2 months, 0 km 0.20 0.09 4

0 month, 630 km 0.79 0.58 15


0 month, 1260 km 0.46 0.14 12

Latitudes 28.125° -0°

1 month, 0 km 0.82 0.45 26


2 months, 0 km 0.75 0.24 28

0 month, 630 km 0.70 0.46 13


0 month, 1260 km 0.40 0.06 12
2 9 6 7 The GCM Validation

For example, for the one-month temporal lag (and the zero spatial
lag) the compared correlations arc 0.82 and 0.45, and the N value is
approximately 26. For the 630 km spatial lag (and the zero temporal
lag) the correlations are 0.70 and 0.46, and N « 13. The differences
of the correlations are statistically significant at any standard level.
The spatial and temporal dependences of the observed data are high,
which was more difficult to foresee.
The estimates of the correlation functions confirm the inferences
derived from the spectral analysis: There is a significant qualita-
tive and quantitative distinction in the power of the white noise and
the low-frequency components in the observed and simulated climate
variability for the tropical region, or the two samples analyzed are
from different distributions.
The second pair of the subfields to be considered corresponds to
the midlatitude region (33.75 — 78.75° latitude). The size of each field
is 8(5.625° (or 630 km) width latitude bands]x372 months.
The most general feature of the frequency —wave-number spectra
of these snbfields (Figure 7.6) is the multiple crests spread along the
frequency axis. This means that the temporal variability is reflected
by the fluctuations with all possible periods from two months to many
years. It seems that (in the case of simulated data) the small crests
for different frequencies result from the large number of nonlinear
transformations of the amplitude modulation type produced in the
process of the numerical integrations of the hydrodynamic equations,
while for the observed data such crests result from multiple interac-
tions and feedbacks of many processes of nature. It is possible to
eliminate all these irregularities by increasing the spectral window
width. However, to keep the statistical methodology identical with
that used in the tropical region, this was not done.
The spatial variability is realized mainly by the large spatial
waves (the spatial part of the spectra has maxima corresponding to
the zero wave number). The concentration of the power in the low-
frequency part of the observed spectrum is also greater than that
for the simulated data, but the maxima themselves are significantly
smaller than those for the tropical region.
Spatial-temporal correlation functions (see Figure 7.7 and Table
7.10) slightly resemble each other especially in their temporal parts,
but the differences in the spatial correlations are still large, and, of
course, statistically significant (Table 7.10). The overall comparison
of the two correlation's fields for the midlatitude suggests that the
corresponding samples are from different distributions.
The volume of the available observations (as well as the size of
the area) of the arctic region (78.75-90.00° latitude) is too small to
Figure 7.6: Spectra of the observed (bottom) and simulated (top) spatial-temporal temperature
fields of the midlatitude region.
298 7 The GCM Validation

Figure 7.7: Correlation functions of the observed (bottom) and simulated (top)
spatial-temporal temperature fields of the midlatitude region.
7.7 Conclusion 299

be appropriately represented by the two-dimensional field, but some


conception about statistical structures of its observed and simulated
temperatures can be gathered from the estimates of the temporal and
spatial correlations given in Table 7.10. The temporal variability of
both types of data is almost identical and are very close to a white
noise process. Their spatial variability is still sufficiently different.
In short, the analysis provided in this section is an approximation
of different latitude-temporal data structures by the homogeneous
fields. Yet even though the analysis is rough, it makes possible to
find the sufficient distinction of the sampled distributions. This dis-
tinction shows that there is much to be done before the practical
application of the GCM for the analysis of the real climate variabil-
ity can be scientifically justified.

7.7 Conclusion
There were two kinds of results obtained. The first includes the
features of the observed spatial and temporal climate variability (the
second-moment climatology) which can be used to validate any GCM
and for some other purposes. The second includes the comparison
of these features with the analogous characteristics of the climate
simulated by this particular version of the Hamburg GCM.
The most general conclusion that can be made about the vari-
ability of the real climate is that all of the kinds of second moments
considered (spectra, correlations, parameters of the stochastic mod-
els) are latitude-dependent. There are three clearly distinguishable
latitude regions (arctic, midlatitude, and tropics) within which such
variability can be assumed as approximately homogeneous.
The comparative study above showed the general similarity of
the spectral and correlation spatial-temporal structures of the ob-
served and simulated second moments. Some estimates, such as lin-
ear trends and advection velocities, have approximately the same
orders for both types of data. However, the character of the latitu-
dinal distribution of the white noise component in the fluctuations
is quite different. The latitudinal trend of the observed spectra and
correlation functions has a very specific feature: the dramatic change
of its structure on the boundary between the tropics and the middle
latitudes. This change is in complete agreement with the theoretical
understanding of the character of the meridional atmospheric circu-
lation. In contrast, such a latitudinal trend of the simulated data is
gradual and smooth.
300 7 The GCM Validation

The autoregressive model for the observed data shows that the
temperature fluctuations are most strongly dependent upon the fluc-
tuations in the southern regions, which is not the case for the sim-
ulated data. This abrupt change of the second moments on the
boundary between midlatitude and tropics, the different character
of the northward transport of heat (parameters of the ARMA mod-
els), the level of a white noise component, and some other features,
show that the observed meridional circulation differs from that in the
GCM. A relationship between the coefficients of the diffusion equa-
tion and the parameters of the stochastic model has been found that
has led to the estimation and comparison of the advection velocity
for the corresponding latitude bands. Such advection velocities are
different for both types of data.
Comparison of the statistical structures of the two-dimensional
spatial-temporal fields for tropics and midlatitudes revealed that
their main distinction is in the redistribution of their low-frequency
composition, where the power of the observed spectrum is signifi-
cantly greater than the power of the simulated one. This implies a
relatively lower level of the white noise component for the observed
data and provides evidence that the possible causes of the observed
long-period climate fluctuations can be explained by some natural
phenomena similar to El Nino or increasing of C02 concentration.
The methodology developed may help in the verification and com-
parison of many observed and simulated processes and fields. More-
over, problems identical to those presented in this study arise in
comparison of different GCM models or the results of different sim-
ulations by the same model with different dynamical components or
boundary conditions. The problem lies in identifying appropriate
priorities and in determining the scale of the spatial and temporal
averaging of data, as well as in revealing the possible uses of different
kinds of statistical information obtained.
On the whole, the comparison of the observed and simulated
second moments identifies that many details of their structures differ
markedly in quality and quantity; and the largest differences in the
compared statistics, found for the tropical regions, can, possibly,
serve as a measure of such a distinction. This fact also shows that
physical modeling of the heat fluxes must be done more accurately.
Therefore, the problems of the GCM diagnoses and verifications
can be stated and solved in terms of standard statistical inference
procedures concerning the distribution parameters associated with
observed and simulated atmospheric systems.
The statistical methodology developed here can also be used for
generating multiple samples of any size with statistical properties
7.7 Conclusion 301

that are identical to the properties of the observations; for finding


the inverse solution of some of the hydrodynamic equations; and for
studying some other problems of meteorological and climatological
data analysis and modeling. Of course, the GCM verification is a
continuous and complicated process (because the physical and com-
putational methodologies are constantly being perfected and because
the volume of observations is constantly being increased), but even-
tually the number of compared variables will become greater.
8
Second Moments of Rain

The first part of this chapter presents a description of the GATE rain
rate data (Polyak and North, 1995), its two-dimensional spectral and
correlation characteristics, and multivariate models. Such descrip-
tions have made it possible to show the concentration of significant
power along the frequency axis in the spatial-temporal spectra; to de-
tect a diurnal cycle (a range of variation of which is about 3.4 to 5.4
mm/hr); to study the anisotropy (as the result of the distinction be-
tween the north-south and east-west transport of rain) of spatial rain
rate fields; to evaluate the scales of the distinction between second-
moment estimates associated with ground and satellite samples; to
determine the appropriate spatial and temporal scales of the simple
linear stochastic models fitted to averaged rain rate fields; and to
evaluate the mean advection velocity of the rain rate fluctuations.
The second part of this chapter (adapted from Polyak et al., 1994)
is mainly devoted to the diffusion of rainfall (from PRE-STORM ex-
periment) by associating the multivariate autoregressive model pa-
rameters and the diffusion equation coefficients. This analysis led to
the use of rain data to estimate rain advection velocity as well as
other coefficients of the diffusion equation of the corresponding field.
The results obtained can be used in the ground truth problem for
TRMM (Tropical Rainfall Measuring Mission) satellite observations,
for comparison with corresponding estimates of other sources of data
(TOGA-COARE, or simulated by physical, models), for generating
multiple rain samples of any size, and in some other areas of rain
data analysis and modeling.

302
8.1 GATE Observations 303

8.1 GATE Observations


8.1.1 Objectives
For many years, the GATE data base has served as the richest and
most accurate source of rain observations. Dozens of articles present-
ing the results of the GATE rain rate data analysis and modeling
have been published, and more continue to be released. Recently,
a new, valuable set of rain data was produced as a result of the
TOGA-COARE experiment. In a few years, it will be possible to
obtain satellite (TRMM) rain information, and a rain statistical de-
scription will be needed in the analysis of the observations obtained
on an irregular spatial and temporal grid. When satellite measure-
ments over a certain region take place only twice a day, there may
be significant losses of information, leading to inaccuracies in the cli-
matological averages of rain. For example, if a rain time series has a
diurnal cycle (hidden periodicity), the specific moments (of the day)
in which the two daily readings are taken determine a bias in the
spatial and temporal averages of the satellite data. A standard sta-
tistical technique for detecting such hidden periodicity is spectral and
correlation analysis. Knowledge of a diurnal cycle of a non-zero rain
rate is important in operational practice for retrieving data from the
records of brightness temperature when one needs to know a range
of variations of a rain rate.
These developments will lead to a broadening of the scope of
comparative statistical description and modeling of rain in order to
evaluate the scales of possible distinctions in rain climate character-
istics for different geographical regions of the earth.
For a long time, the univariate autoregressive model has been one
of the most widespread methodologies for modeling rainfall time se-
ries. The multivariate methodology applied here is statistically and
physically richer than the univariate approach. Moreover, multivari-
ate modeling enables us to study the diffusion of rain because there
is a simple interconnection between diffusive and multivariate AR
presentations of multiple rain time series.
Second-moment estimates provide the most appropriate way for
comparing the climatic features of rain because the theory and prac-
tice of statistical inferences and hypotheses testing have already been
developed for such moments.
On the whole, the objectives of this study are as follows:
1. To study the features of rain variability in the spaces of the
temporal-spatial lags and of the frequency-wave numbers and
to evaluate a possible distinction in estimates corresponding to
radar and satellite observations.
304 8 Second Moments of Rain

2. To estimate the diurnal cycle of a non-zero rain rate.


3. To build multivariate AR models and to study the diffusion of
rain.
4. To document the initial data structures and to demonstrate
the real volume of the non-zero observations used for obtaining
different statistics.
Multiple analogous studies (Bell and Reid, 1993; Chiu et al., 1990;
Crane, 1990; Graves et al., 1993; Hudlow, 1977, 1979; Kedem et al.,
1990; McConnell and North, 1987; Nakamoto et al., 1990; North and
Nakamoto, 1989; Simpson et al., 1988; and so on) with the same data
set were carried out. But our approach is different in presenting a
detailed pictorial description of data as well as in its application of
standard statistical technique of two-dimensional and multivariate
analysis to spatially and temporally averaged fields. Such averaging
of the original three-dimensional array helps to reduce the noise com-
ponent and to show the second-moment characteristics of rain more
distinctly. Whenever possible, we also tried to find the simplest
stochastic model of one or two parameters, because the comparison
of a small number of parameter estimates is the most effective way to
solve the ground truth problem as well as to provide a comparative
climatological description of rain.

8.1.2 Data Documentation


The GATE data base, created and discussed in detail by Arkell and
Hudlow (1977) and by Patterson et al. (1979), offers the most com-
plete archive of rain observations. The two 19-day intervals (June 28
to July 16 of GATE1 and July 28 to August 15 of GATE2) of rain rate
observations used in this study were collected over a, 400 km diameter
hexagonal area in the Atlantic Ocean, centered on 8°30'.W, 23°3()'W
off the west coast of Africa. The data were obtained by averaging the
15 minute rain rate (mm h-1) radar and surface ship measurements
over 4 X 4 km pixels. Therefore, the temporal step of the data is 15
minutes and the spatial step in both directions is 4 km.
Table 8.1 documents the number of observed spatial fields (100 X
100 size) within each hour for the time interval of the GATE1 exper-
iment. Because there are only 1,7.16 terms of the two-dimensional
time series in the GATE1 archives, the omitted or missed fields were
replaced by zero fields. After such replacement, the initial structure
obtained is the following matrix time series:
8.1 GATE Observations 305

Figure 8.1: Number of non-zero observations of the GATEl rain rate data base.

where i,j, and t are the subscripts in the directions of north-south


(NS), west-east (WE), and time, respectively.
To obtain the general representation of the GATEl array, the
number of the non-zero observations for each spatial point (i,j) was
counted and presented in Figure 8.1 (its contour lines are given in
Figure 8.13).
This figure shows that the number of readings with rain for dif-
ferent points of the ocean varies from zero to 600; in other words, the
proportion of readings with rain varies from zero to about one-third
of the total number of terms. The farther out in the Northern part
of the Atlantic Ocean, the fewer were the number of observations
with rain. In the background of this significant north-south trend,
there are two maxima of the amount of non-zero observations corre-
sponding exactly to the locations of the two ships (the Oceanographer
at the center and the Researcher at the most southern point of the
GATE area) whose data were used for creating of the data base.
Such a dependence (which does not really exist) between the rainfall,
amount and the ship locations clearly shows that not all of the rains
in the GATE spatial area were recorded for the observational inter-
val and that it is possible that not all of the non-zero readings were
really rain.
Analysis of data averaged over all spatial points i and j for each
moment t of the readings shows that the spatial mean rain rate varied
from zero to about 13 mm/hr and that there were only a few 15-
306 8 Second Moments of Rain

minute intervals when rain was not recorded. Some of the general
statistical characteristics of the GATE1 data appear in Table 8.2.
This study deals with the spectral and correlation analysis of
three types of two-dimensional rain rate fields obtained by averaging
data sequentially along axes (subscripts) i,j, and t (8.1). Averag-
ing, as well as estimating the sample standard deviations, was done
by adding only the non-zero observations and dividing the sum by
their number. Then, the anomalies were divided by the correspond-
ing estimates of standard deviations. The zeroes were left as zeroes
after normalization. The elements of such two-dimensional fields of
normalized anomalies will be denoted as Uij or uit.
The standard methodology with two-dimensional Tukey spectral
window was employed. The values of the periodograms and the spec-
tral estimates were multiplied by some factor to be conveniently pre-
sented by illustrations. The factor used is 10,000.

8.1.3 Spatial Field


Averaging data (8.1) along the t-axis over the entire observational
interval for each point (i, j) of the surface gives the square spatial field
of 100 X 100 size with a 4 km step along each axis. This field reveals
that the range of variations of the temporal means is 0-20 mm/hr.
Because the sample mean of the entire data is about 4.4 mm/hr, the
sample distribution of the observations must be positively skewed.
The spatial spectrum estimates (Figure 8.2) show the concentra-
tion of almost all power in the vicinity of zero wave numbers with a
significant peak centered in the origin.
Therefore, this spatial field has a two-dimensional trend possibly
caused in part by the decreasing trend of non-zero observation num-
ber with distance from the ship (see Figure 8.1) and in part by the
natural latitudinal variations of rain. Most of the spectral estimates
are zeroes. The estimates also show that the shortest wavelengths
with non-zero power are about 32 km. According to the Nyquist the-
orem, this means that it is useless to choose a spatial resolution of
the observations that is less than 16 km for such a spectral analysis.
Thus, the first simple inference to be drawn from the spectrum
above is that the optimal sampling of the rain rate observations must
have spatial steps of about 16 km, which, as compared to the GATE1
data sampling, would reduce to 4x4 times a volume of the analyzed
data without losing any information.
The two-dimensional spatial correlation function is given in Fig-
ure 8.3, its contour lines in Figure 8.4.
Table 8.1: Count of the GATEl observed spatial fields within each hour.
Day —
Hour 179 180 181 182 183 184 185 186 187 188 189 190 191 192 193 194 195 196 197

0 3 4 3 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4
1 4 4 4 4 4 4 4 4 4 4 4 0 4 3 4 3 4 4 4
2 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4
3 4 4 4 4 4 4 4 4 4 4 4 4 3 3 4 4 4 4 4
4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 3 4 4 4
5 4 4 4 4 4 4 4 4 4 4 4 4 4 4 3 4 4 4 4
6 4 3 4 1 4 4 4 4 4 4 4 4 4 4 3 4 4 4 4
7 4 4 4 0 4 4 4 4 4 4 4 4 4 4 3 4 4 4 4
8 4 4 4 0 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4
9 4 4 4 0 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4
10 4 4 4 0 4 4 4 4 4 3 4 4 4 4 4 4 4 4 4
11 4 4 3 0 4 4 4 4 4 4 3 4 4 4 4 4 4 4 4
12 3 0 1 3 4 2 3 4 3 4 2 4 4 4 4 4 4 4 4
13 4 1 2 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4
14 4 5 4 4 3 4 4 4 4 4 4 4 4 4 4 4 4 4 4
15 4 4 4 3 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4
16 4 4 4 4 4 4 4 4 4 4 4 4 3 4 4 4 4 4 4
17 3 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4
18 0 4 2 4 4 4 4 1 4 4 1 3 2 3 4 4 4 4 0
19 3 4 4 4 4 4 4 4 4 4 2 4 4 4 4 4 4 4 0
20 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 0
21 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 0
22 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 0
i
23 4 3 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 0
308 8 Second Moments of Rain

Table 8.2: Some general characteristics of the entire GATE1 data


and 12-hour time-step sample.

Characteristics GATE1 data 12-hour sample

Grand mean (mm h-1)h-1 4 4.419 4.2000


Standard deviation 7.092 6.932
Number of non-zero
observations 1466271 24201
Percentage of non-zero
observations 9 8

The isolines reflect the predominance of the transport of rain rate


fluctuations along the west-east line that match the illustration (Hud-
low, 1979) of the surface streamlines of the mean wind directions in
the time and area of the GATE1. Of course, these isolines cannot
point out the direction (west or east) of this transport. Therefore,
the spatial dependence of rain rate fluctuations clearly show that
the synoptic transport of rain along the parallels is reflected in the
second moments and that the corresponding powerful propagating
waves dominate over all other waves. Meridional propagation is sig-
nificantly weaker, Although the prevalence of such transport of rain
was expected, its domination of the two-dimensional second-moment
statistics is not a trivial result.
The correlation function also shows that, for distances of about
50 km, the spatial correlation coefficient damps to values of about
0.3-0.38. Positive correlations, which can be traced up to distances
of slightly more than 200 km into the parallel direction and up to
distances of about 100 km into the meridional direction, show that
the spatial rain rate field is anisotropic. The anisotropy appears
especially clear for distances greater than 35 km, where the values
of the correlations into the NS line damp to zero faster than they do
along the WE line. However, for distances of less than 35 km, the
anisotropy is hardly distinguishable.
This statement can be clearly illustrated by taking cross-sections
of the above correlation function (Figure 8.3) at the two vertical
8.1 GATE Observations 309

Figure 8.2: Spatial spectrum of the rain rate field obtained by averaging the
GATE1 data over the time domain.

planes corresponding to the zero lags. The two resulting one-dimensi-


onal correlation functions (Figure 8.5) into the meridional and paral-
lel directions, respectively, show that, for distances of about 35 km,
their values are very close and can be approximated, for example, by
the exponential correlation function p(r) = 0.89r, where r is lag in
km.
Therefore, for lags less than 35 km, the deviation of the statistical
properties of the spatial rain rate field from the isotropic one is not
significant. For lags greater than 35 km, the difference in the esti-
mates of these one-dimensional correlation functions grows markedly,
reflecting the real distinction of the statistical character of the WE
and NS transports of rain rate fluctuations.
In order to find a better stochastic model of the spatial field con-
sidered, the approximation of the above one-dimensional correlation,
functions by the correlation functions of the second- and third-order
autoregressive processes was done. Because the results are not much
better than those for the first-order model with the correlation func-
tion p(r) = 0.89r, they are not shown.
310 8 Second Moments of Rain

Figure 8.3: Spatial correlation function of the rain rate field obtained by aver-
aging the GATE1 data over the time domain.

One other attempt at one-parameter modeling by the Laplace


equation in accordance with the methodology of a paper by Jones and
Vecchia (1993) should be mentioned (see Subsection 3.8.5). The two
correlation functions of the solution of the Laplace equation (3.100),
together with the above one-dimensional estimates, are given in Fig-
ure 8.6.
One can see that the correlation function of this model proves rel-
atively close to the estimates only for small distances. On the whole,
as follows from comparison of Figures 8.5 and 8.6, for distances of less
than 35 km the concave shape of the correlation function of the first-
order AR process is slightly more appropriate than the correlation
function of the solution of the Laplace equation.
According to Jones and Vecchia (1993), some anisotropic fields
can be transformed to the isotropic form by rotating the coordinates
by an appropriate angle a and by scaling the two coordinate axes
differently (see Subsection 3.8.5). However, careful consideration of
the isolines in Figure 8.4 reveals that the anisotropic character of
the rain rate field is very complicated and that it is impossible to
transform it into an isotropic one by rotating and scaling coordinates.
8.1 GATE Observations 311

Figure 8.4: Contour lines of the spatial correlation function in Figure 8.3.

One can show that the angle a and the scaling parameter are not
constant and that they differ for different points of the field. For
example, the a is almost zero for distances of less than 35 km and
grows significantly with increasing lag.
But both models considered are very simple (one parameter), so
for small distances the estimates of these parameters can be used
in comparing characteristics of rain obtained by different kinds of
measurements.

8.1.4 Ground Truth Problem


Any measurements of rain present different samples from the un-
known distribution of rain considered as a multidimensional random
function. The finite and discrete character of the samples limits the
312 8 Second Moments of Rain

Figure 8.5: The GATEl one-dimensional correlation functions into the NS and
WE directions and the function p(r) = 0.89r.

possibility of an accurate statistical description of rain in the spatial


and temporal domains. Specifically, according to the Nyquist the-
orem, the time step of the observations limits the high frequencies
that can be studied, while the sample size limits the low-frequency
resolution.
There are many different possibilities for transforming the GATEl
observations into different time series and fields by spatial and tem-
poral averaging and/or sampling. For a comparative study of the
dependence of variability on the time step and the volume of data,
a sample (which imitates satellite data) of spatial fields with a time
step of 12 hours was selected from (8.1) and considered separately.
The chosen fields correspond to exactly midnight and noon.
The counts, analogous to those given in Figure 8.1, of the non-
zero observations in this sample for each spatial point of GATEl area
show that the real number of readings with rain varies from zero to
15 and that there is very little data for the northern region. Some
of the statistical characteristics of the entire GATEl data and the
8.1 GATE Observations 313

Figure 8.6: The GATEl one-dimensional correlation functions into the NS


and WE directions and the correlation functions of the solution of the Laplace
equation, where a = 0.0041 (1) and 0.0005 (2).

12-hour time step sample are presented in Table 8.2. The estimates
of their grand means, given in the first row of Table 8.2, are rela-
tively close (4.4 mm/hr and 4.2 mm/hr), and it is easy to show that
the 99% confidence interval for the mean of the distribution based on
the entire GATEl data is (4.404 to 4.434), while that for the 12-hour
sample is (4.085 to 4.315). The small difference (0.2 mm/hr) of the
above two means can be interpreted as bias of the 12-hour sample
mean. The cause of these particular distinctions will be discussed
shortly. Of course, accurate statistical inference is not completely
appropriate in the above comparison because the samples are sta-
tistically dependent, but such inference will be valid when dealing
with real satellite observations, which are statistically independent
of ground data. The statistical significance of the distinctions in the
standard deviation estimates (7.092 and 6.932) can also be tested in
the case of independent samples.
Let us compare some of the results presented in Figures 8.2 and
8.3 with the analogous estimates for the field obtained by averaging
314 8 Second Moments of Rain

Figure 8.7: Spatial correlation function of the rain rate field obtained by the
12-hour time step sample averaging over the time domain.

the 12-hour time step sample over the time domain. The shape of
the spectrum, which is not cited here, is similar to the spectrum
in Figure 8.2, but the maximum in the origin is about three times
smaller than that for the entire data.
The comparison of the correlation functions (Figure 8.7 with Fig-
ure 8.3) shows that the spatial statistical dependence of the averages
of the 12-hour sample is significantly smaller, as expected, or that
damping is faster. Positive correlations can only be traced up to
distances of less than 100 km.
For example, for lags equal to 4 km, correlations in Figure 8.3 are
equal to about 0.83-0.85, while for the 12-hour sample such correla-
tions are only about 0.64-0.56. The values are different, but if the
samples were independent, it would be possible to test a statistical
hypothesis and make inferences about differences in the correlations
of the distributions.
8.1 GATE Observations 315

Another feature of the 12-hour sample that should be mentioned


is that instead of seeing the prevalence of the transport along the par-
allels we see the predominance of southeastward propagation. This
distinction could possibly be connected with the diurnal cycle of the
wind direction in this area in the time of the experiment. The corre-
sponding one-dimensional correlation functions (which are not cited
here), along the WE and NS lines, are close to each other, and they
can be approximated by the exponential curve 0.7T". But this model
is valid only for distances of about 20 km. Modeling by the Laplace
equation (see Subsection 3.8.5) is also appropriate only for the same
distances, but the accuracy in this case is slightly lower than that of
the above exponential approximation.
Concluding the comparative consideration of the statistics of the
two rain rate spatial fields makes it possible to assert that the values
of their spectral and correlation estimates differ markedly. Therefore,
the 12-hour time step sample, an imitator of the satellite data, is
not statistically representative for the rain rate field in the sense
that its second-moment spatial statistics differ significantly from the
corresponding estimates, which were obtained using the entire data
set.

8.1.5 Spatial-Temporal Fields


8.1.5.1 Latitude-Temporal Field
Averaging data set (8.1) along the north-south direction over the
subscript i gives the two-dimensional spatial-temporal rain rate field,
which can be considered as a latitude-temporal field. The size of the
field is 100 x 1824.
The central part of the spectrum of the normalized values of this
field (Figure 8.8) shows a concentration of almost all power in the
narrow wave number band along the frequency axis, which means
that the observed variations of data take place mainly in time.
The primary feature of the spectrum is a significant crest over the
frequency axis with a wide maximum that peaks in the origin. The
maximum reveals that the data has a spatial-temporal trend, which
can be associated with the trend of the non-zero observation number
along the west-east direction and, possibly, with the seasonal cycle.
On the background of the main crest, there are several smaller ones,
the most noticeable and interesting of which is a lowering crest cor-
responding to a period of about 24 hours. It is spread in the domain
of the positive (negative) frequencies and wave numbers quadrant
along almost all wave numbers with a non-zero power. Moreover, it
316 8 Second Moments of Rain

Figure 8.8: Central part of the spatial-temporal spectrum of the rain rate field
obtained by averaging the GATEl data along the NS direction.

can be seen that this diurnal maximum is more prominent for the
non-zero wave numbers than for the zero wave number. This feature
explains the cause of difficulties in detecting the diurnal cycle using
spectral analysis of univariate time series. The standard deviations
of the non-zero spectral estimates in Figure 8.8 vary from point to
point with approximate values in the interval (0.03-0.5); therefore,
the spectral maximum of a diurnal cycle is statistically significant.
The two-dimensional spatial-temporal correlation function (Fig-
ure 8.9) reveals that the temporal correlation for the 15-min lag is
equal to 0.77 and that the spatial correlation for the 4-km lag along
the west-east line is 0.81.

8.1.5.2 Longitude-Temporal Field


Averaging data set (8.1) along the west-east direction over the sub-
script j gives the two-dimensional spatial-temporal rain rate field,
which can be considered as a longitude-temporal field of the same
size as the previous one. The spectrum and correlation function
(which are not cited here) shapes of this field reveal their similarity
with those given in Figures 8.8 and 8.9. But it can be mentioned that
8.1 GATE Observations 317

Figure 8.9: Spatial-temporal correlation function of the rain rate field obtained
by averaging the GATE1 data along the NS direction (levels are 0.2, 0.4, and
0.6).

the peak of the spectrum in the origin is more powerful than that for
the previous latitude-temporal field. The most plausible reason for
this is the latitudinal trend of the amount of data. A small lowering
crest corresponding to the diurnal frequency (0.042) can be seen here
as well, although the transport of rain rate fluctuations is occurring
primarily along the WE line.
For a more descriptive comparison, the univariate autocorrelation
functions of the cross-sections by the plane with the zero spatial lags
of the above two spatial-temporal correlation functions were consid-
ered separately. It was found that these autocorrelation functions
can be approximated by a first-order autoregressive model [with cor-
relation function p(T) = 0.35T,T is hours] only for lags within the
interval of one hour. An attempt to fit the second- and third-order
318 8 Second Moments of Rain

AR models was not successful. In order to accurately approximate


the temporal correlation functions on the 10 to 15 hour lag interval,
the order of the AR model must be greater than 10. The most sig-
nificant features of the two spatial-temporal spectra are the lowering
crests corresponding to the diurnal cycle frequency. This means that
the two-dimensional spectral analysis shows the existence of a diurnal
cycle in the GATE1 rain rate observations with a small amplitude.
To gain insight into both the extent to which the results reflect
reality and the reliance upon applied methodology, one must realize
that the periodograms of spatial-temporal rain rate fields have a very
complicated random character with many discrete lines and peaks for
randomly spread frequencies and wave numbers. The smoothing pro-
cedure leads to the formation of one powerful peak (in the origin) and
a few other small, but noticeable, spectral crests, one of which can
be interpreted as a reflection of the diurnal cycle. The most likely
cause for the maximum being at the origin, as already mentioned,
is the trend of data imposed by the nonuniform number of observa-
tions (see Figure 8.1), the seasonal cycle, and the natural latitudinal
variability. The most likely cause of other tiny maxima and crests
are the sample variability.

8.1.6 Diurnal Cycle


Now, we can look at the diurnal cycle estimates obtained by spatial
and temporal averaging over all non-zero observations corresponding
to each of the 96 fixed moments of the day. The estimated curve
(Figure 8.10) has a relatively simple sine wave shape with minimum
values during the night and maximum values during the afternoon.
The range of variation (3.4 to 5.4 mm/hr) is small compared with
the standard deviation (about 7 mm/hr) of observations; this is the
cause of difficulties in its detection in the spectral estimates.
The diurnal cycle estimate accuracy is relatively high; their stan-
dard deviations are less than 0.1 mm/hr.
The comparison and verification of the GATE1 diurnal cycle es-
timates with the results derived from other sources of data must
be done very carefully because the diurnal cycle can have a differ-
ent shape in different regions and its seasonal and latitudinal trends
may obscure other patterns. But it is interesting that, in spite of the
obvious fact that the estimates in Figure 8.10 reflect the features of
the GATE1 sample, the rain rate diurnal cycle shape very closely fol-
lows the surface temperature diurnal cycle shape (see, for example,
Schmugge, 1977).
8.1 GATE Observations 319

Figure 8.10: Estimates of the diurnal cycle of the GATE1 rain rate data and
their averages within each hour.

The practical importance of taking the diurnal cycle into account


when solving different engineering problems can be shown in the
framework of the ground truth problem described above. Thus, av-
eraging the two separate estimates, corresponding to midnight and
noon (the two moments of the day for which the readings of the 12-
hour sample were taken) on a curve in Figure 8.10 gives exactly the
12-hour sample mean estimate of 4.2 mm/hr (see Table 8.2).
Averaging all 96 estimates in Figure 8.10 gives the entire GATE1
sample mean of 4,419 rnm/hr. The difference between these two
estimates is equal to the above-mentioned bias, and it is completely
conditioned by the presence of the diurnal cycle in the GATE1 data
set. It is clear that such a bias can have nothing in common with
the distinction in the probability distributions of errors of different
kinds of rain measurements.
Let us associate the rain rate diurnal cycle with the diurnal cycle
of number of non-zero observations. Such a diurnal cycle (Figure
320 8 Second Moments of Rain

Figure 8.11: Diurnal cycle of the number of non-zero observations of GATEl


data and the averages within each hour.

8.11) shows the reliability of the results as well as the approximate


character of the diurnal cycle of rainfall because, on average, the
greater the number of non-zero observations, the greater the amount
of precipitation. With the estimates of Figure 8.10 in mind, it seems
natural to expect that the maximum of the number of non-zero obser-
vations must correspond to the afternoon hours. Of course, relatively
small lunch and dinner minima on the curve in Figure 8.11 have noth-
ing in common with rain statistics, but they do reveal the remarkably
consistent character of data from the GATEl experiment.

8.1.7 GATE2 Statistics


The source of GATE2 rain rate observations is the radar and sur-
face measurements that took place from July 28 to August 15 of
1974 at the same region of the Atlantic Ocean. Table 8.3 gives a
general representation about the volume of observations in the cor-
responding data base. There are almost three days with entirely
missing data, and the total number of non-zero observations equals
only two-thirds of the corresponding number of the GATEl data.
8.1 GATE Observations 321

Figure 8.12: Spatial correlation function of the rain rate field obtained by
averaging the GATE2 data over the time domain.

This inevitably leads to lower accuracy of the second-moment esti-


mates. Nevertheless, some of them will be very briefly discussed.
The shape of the spatial spectrum (which we do not show here) of
the two-dimensional spatial GATE2 field is approximately the same
as that of Figure 8.2, but the peak in the origin of the wave number
coordinates is smaller.
Figure 8.12 shows the corresponding spatial correlation function,
analogous to that given in Figure 8.3. It is hard to believe that
such a beautiful structure reflects the properties of the most variable
random process of nature rather than the result of simulation by a
purely artificial deterministic model. It leaves hope that an accurate
description of rain rate in the form of an interpretable physical model
will be found. The anisotropy of the GATE2 spatial rain rate field
carries a different character than that of the GATE1 data; specif-
ically, the predominant transport of rain occurs along the NW-SE
line rather than along the parallels, as it did for the GATE1 data.
Moreover, the values of the correlations in both directions for the
Table 8.3: Count of the observed GATE2 spatial fields within each hour.
Day
Hour 209 210 211 212 213 214 215 216 217 218 219 220 221 I 222 223 224 225 226 227

4 3 3 0 0 0 4 4 4 3 2 4 4 3 4 4 4 4 4
?2 4
4
4 4 0
0
0
0
0 4 4 4
4
4 4 4 4 4
4
4 4 4 4
4 4 0 4 4 4 4 4 4 4 4 4 4 4
3 4 4 4 0 0 0 4 4 4 4 4 4 4 4 4 4 4 4 4
4 4 4 4 0 0 0 4 4 4 4 4 4 4 4 4 4 4 4 4
5 4 4 4 0 0 0 4 4 4 4 4 4 4 4 4 4 4 4 4
6 3 4 4 0 0 0 4 4 4 4 4 4 4 4 4 4 4 4 4
7 4 4 4 0 0 0 4 4 4 4 4 4 4 4 4 4 4 4 4
8 4 4 4 0 0 0 4 4 4 4 4 4 4 4 4 4 4 4 4
9 4 4 4 0 0 0 4 4 4 4 4 4 3 4 4 4 4 4 4
10 4 4 4 0 0 0 4 4 4 4 4 4 4 4 4 4 4 4 4
11 4 4 4 0 0 0 4 4 4 4 4 4 4 4 4 4 4 4 4
12 4 3 4 0 0 0 4 3 4 4 4 4 4 3 4 4 4 4
13
14
4
4
4
4
4
4
0
0
0
0
0
0
4
4
4
4
4
4
4
4
4
4
4
4
4 4 4 4 4
4
4 4
4 4 4 4 4 4
15 4 4 4 0 0 0 4 4 4 4 4 4 4 4 4 4 4 4 4
16 4 4 4 0 0 0 4 4 4 4 3 4 4 4 4 4 3 4 4
17 4 4 4 0 0 0 4 4 4 3 2 4 4 4 4 4 4 4 4
18 3 4 3 0 0 1 4 4 4 2 3 4 4 4 4 4 4 4 4
19 4 4 4 0 0 3 4 4 4 2 4 4 4 4 4 4 4 4 4
20 4 4 4 0 0 4 4 3 4 4 4 4 4 4 4 3 4 3 4
21 4 4 3 0 0 4 4 4 3 4 4 4 4 4 4 4 4 4 4
22 4 4 1 0 0 4 4 4 4 4 4 4 4 4 4 4 4 4 0
23 4 4 0 0 0 4 4 4 3 4 4 4 4 4 4 4 4 4 0
8.1 GATE Observations 323

GATE2 spatial field are slightly greater than those for the GATE1
spatial field (Figure 8.3).
In general, the shapes of the spatial-temporal spectral and corre-
lation characteristics of the GATE2 data (the illustrations of which
are not provided here) are similar to those of the GATE1 data in
Figures 8.8 and 8.9. However, the details are different. First, instead
of having a spectral maximum for the 0.042 frequency value (the 24-
hour period), there are a couple of small peaks for the slightly lower
and slightly greater frequencies. Therefore, it makes no sense to
evaluate the corresponding diurnal cycle of rain rate for the GATE2
data.
Comparing the diurnal cycles of the number of non-zero obser-
vations for both phases of the experiment shows that the volume
of GATE2 data equals approximately 60% of the volume of GATE1
data. The cause of this is not simply the three days with entirely
missing data. The lunch and dinner minima of the GATE2 data are
wide, and the corresponding losses of information are significant for
only the 19-day interval of observations, especially for evaluation of
the temporal characteristics. That is one of the reasons a diurnal
maximum was not found in the corresponding spectral estimates.

8.1.8 Multivariate Models


Fitting linear stochastic processes to the multiple rain time series
leads to the creation of a stochastic rain model, which has indepen-
dent meaning as one possible approach to its description.
In the case of rain observations, the goal of modeling is not, of
course, forecasting in the sense of predicting the onset or the end
of rain. Our intention is to build rain rate models and to use them
for diagnostic analysis of parameter matrices in terms of spatial-
temporal interactions and feedbacks of observed processes.
Physical interpretation of parameters of multivariate stochastic
models of spatially and temporally averaged data can provide the ba-
sis for a statistical description of a transport of rain fluctuations. To
solve this problem, one must take into account a high level of white
noise associated with fluctuations measured at separate points and
attempt to reduce it. The signal detection may be enhanced through
spatial and temporal averaging of data with a consecutive fitting of
multivariate autoregressive models to the resulting time series; so
averaging must be done to present the rain rate fields employed in
the form of multiple time series. A stochastic model of rain rate
fluctuations will generally be sensitive to the scales of spatial and
324 8 Second Moments of Rain

Figure 8.13: Contour lines of the number of non-zero observations of the


GATEl rain rate data base and the specific spatial bands of averaging for multi-
variate modeling.

temporal averaging, In the process of this work, several multivari-


ate rain rate models corresponding to different scales of spatial and
temporal averaging were constructed and analyzed, but only the two
simplest models are presented here.
The data was averaged temporally within an hour (the temporal
step employed is 60 minutes). Because of the distinctions in the
statistical characteristics of the rain rate field in the NS and WE
directions, the two different spatial averaging strategies, illustrated
in Figure 8.13, were chosen.
The first strategy (Figure 8.13, 1), averaging within three 50-km
width longitude bands with a maximum amount of data, is intended
to describe west-east interdependence. The second strategy (Figure
8.1 GATE Observations 325

Table 8.4: Matrix correlation function for the first two lags of the
three-variate latitude-temporal model (one-hour time step).

u1 u2 u3

T = 0

u1 1.00 0.40 0.23


u2 0.40 1.00 0.64
u3 0.23 0.64 1.00

T = 1

u1 0.65 0.45 0.23


u2 0.34 0.72 0.54
u3 0.26 0.64 0.77

8.13, 2), averaging within three 50-km width latitude bands with
a maximum amount of data, is intended to explore the extent of
meridional dependence of rain rate perturbations. Therefore, the
time step between the two terms ut+1 and ut of the time series studied
is one hour, and the space distance between two adjacent time series
is 50 km. Two three-variate models were constructed for each spatial
and temporal averaging procedure shown in Figure 8.13.
Here are some of the results. The matrix correlation functions
for the first two lags and the matrices of the estimated parameters of
these two AR models, as well as the parameters of the corresponding
multivariate linear regression models, are given in Tables 8.4, 8.5,
8.6, and 8.7. The statistical dependences presented by the matrices in
Tables 8.4 and 8.6 show that the spatial correlation reaches a value of
0.65 (Table 8.4) into the WE directions, and only 0.39 (Table 8.6) into
the NS directions. The temporal and spatial-temporal correlations
of these data reach a value of 0.77 (Table 8.4) into the WE directions
and of about 0.71 (Table 8.6) into the NS directions. These values
are statistically significant (at any standard level of significance) and
allow us to build a stochastic model.
326 8 Second Moments of Rain

Table 8.5: Estimates (exceeding the 99% significance level) of the


parameters of the three-variate autoregressive and linear regression
models for the time series of the first scheme of averaging (Figure
8.13, 1) and one-hour time step (e is the standard error).

Autoregressive model

ult u2t u3t e

u
it+i 0.61 0.75
u2t + l 0.20 0.44 0.31 0.63
u3t + l 0.71 0.64

Regression model

u1 u2 u3 £

u1 0.42 0.92
u-2 0.26 0.58 0.72
u3 0.65 0.77

Each row of Tables 8.5 and 8.7 with parameter estimates is asso-
ciated with the corresponding autoregressive or regression equation.
The values listed in column e of these tables are the standard er-
rors of the forecasts for one step ahead for the AR(1) model or the
standard errors for the regression model.
Consider the parameters (Table 8.5) of the latitude-temporal AR
model, which is fitted to the data, averaged as indicated in Figure
8.13, 1. With the results of the two-dimensional correlation analysis
(Figure 8.4) in mind, one would expect to find a statistical depen-
dence of the values at any moment upon the past rain rate fluctua-
tions to the west and to the east. In the matrix of the model (Table
8.5) the parameters on the main diagonals are largest and there are
only two statistically significant estimates, 0.20 and 0.31, out of the
8.1 GATE Observations 327

main diagonal. Therefore, the GATE1 rain rate fluctuations are


mainly determined by their own past and partially determined by
the history of the fluctuations of the closest region to the west and
to the east. So, on average, the predominance of east-west transport
of rain in the time of the GATE1 finds its reflection in the estimates
of the parameters of the AR model when the appropriate spatial and
temporal averaging are used to detect the signal.
The linear regression parameter estimates (Table 8.5) show the
determining statistical dependence of the GATE1 data on the values
in the two closest adjacent bands in any particular moment of time.
As can be seen from comparison of the standard errors, the autore-
gressive model accuracy is slightly higher than the regression model
accuracy.
Now the three-variate longitude-temporal model will be consid-
ered for the data obtained by the averaging procedure as shown in
Figure 8.13, 2. The results (Table 8.7) are simple, and the accu-
racy of this three-variate model is lower [standard errors e(t + 1) are
greater] than for the latitude-temporal models. In this case, a non-
zero estimate out of the main diagonal would be evidence of strong
north-south transport of rain. However, the estimates in Table 8.7
do not support such a hypothesis. Only the estimates on the main
diagonal are not zeroes. This means that the rain rate fluctuations
into the NS directions in the time of GATE1 were mainly determined
by its own past, or, in other words, the temporal statistical depen-
dence between the mean rain rate fluctuations of the two adjacent
50-km width bands was not strong enough to cause the correspond-
ing AR model parameter estimates to be statistically significant. So
the transport of rain fluctuations observed was mainly realized along
the parallels. The NS motions were significantly weaker.
Therefore, the approximation of the multiple time series by the
multivariate AR model has the potential to detect the structure of
second-moment rain rate fluctuations in spite of the significant vari-
ability and noise associated with the observations. The first autore-
gressive model shows that, on average, the rain rate fluctuations
were most strongly dependent upon advection along the west-east
line, since the computed matrix of the parameters has a special form
with partially non-zero elements outside the main diagonal. It is in-
teresting to estimate this advection velocity when one has only rain
rate observations.
It is obvious that the advection of rain is determined by the direc-
tion and strength of the wind. Therefore, careful planning is needed
for determining what kind of advection will be evaluated.
328 8 Second Moments of Rain

Table 8.6: Matrix correlation function for the first two lags of the
three-variate longitude-temporal model (one-hour time step).

u1 u2 u3

T = 0

u1 LOO 0.22 0.27


u2 0.22 1.00 0.39
u3 0.27 0.39 LOO

T= 1

u1 0.56 0.22 0.29


u2 0.18 0.44 0.37
u3 0.25 0.26 0.71

The direction of the transport of rain in the time and area of


the GATE1 was not stable and changed several times. As Hudlow
(1979) has shown, the surface streamlines of the mean wind directions
in the time and area of the GATE1 were curved from the northwest
to the east. The purposeful selection of the time interval of the
GATE1 observations with the fixed wind direction is complicated by
the fact that we have too little data for the appropriate statistical
analysis and modeling. In this situation, it was decided, first, to
evaluate the mean advection velocity separately into the WE and
NS directions for the entire observational interval of the GATE1, and
then to obtain such velocities for separate days (whenever possible).
Such a decision determined the strategy of data averaging (see Figure
8.13) for building multivariate autoregressive models and, finally, the
values of the estimated mean advection velocities, as well.

8.1,9 Diffusion
One of the appropriate methodologies for analysis of rain advection
is an approximation of the observed field by the diffusion equation
8.1 GATE Observations 329

Table 8.7: Estimates (exceeding the 99% significance level) of the


parameters of the three-variate autoregressive and linear regression
models for the time series of the second scheme of averaging (Figure
8.13, 2) and one-hour time step (e is the standard error).

Autoregressive model

uIt u2t u3t £

uli+1 0.53 0.82


u2t+l 0.38 0.89
u3t+l 0.64 0.69

Regression model

u1 u2 u3 £

ul 0.21 0.96
u2 0.35 0.91
u3 0.19 0.35 0.90

(see, for example, Cahalan et al., 1982; North and Nakamoto, 1989;
Polyak et al. 1994). This problem can be solved using the relation-
ship between the AR model and a diffusion equation (see Section 7.5).
The parameter estimates (Tables 8.5 and 8.7) give appropriate sam-
ples for subsequent estimation of the diffusion equation coefficients
because the AR model presents the equations in finite deferences of
the same kind as (7.5).
Let us consider the first three-variate AR model. The estimates
in Table 8.5 make it possible to assume that a w 0.20, b w 0.44, and
c w 0.31. By replacing these values in formulas (7.6)-(7.8) and by
performing the computations, one obtains the values of s, v, and q.
The mean advection velocity v for the time of the GATE1 equals
about -5.0 km/hr ( — 1.4 m/sec). A negative value means that the
330 8 Second Moments of Rain

Table 8.8: Mean advection velocity (km/hr) for different days of the
GATE1.

Day 179 180 181 182 183 185 188 189 192 193 194 195

vWE — 64 ~8 -50 —5 — 15 30 -13 0 — ~33 -13 -23

V
NS 0 0 -5 -30 -17 0 2 9 29 -50 ~8

mean rain rate advection took place into the westward direction.
The diffusion coefficients are s « 252 km 2 /hr (4172 m 2 /sec), and
q w 0.04 1/hr (0.00001 I/sec). It seems that these results are quite
reasonable for such computations. Accepting these values as the
estimates of the diffusion equation coefficients, one gets the following:

In using this approach, we attempted to evaluate the mean advection


velocities into the WE direction for each day of the GATE1. This
was possible only for 11 days, for which the volume of the non-zero
data was large enough to estimate the corresponding correlations.
Of course, one can unite the observations for the adjacent days
with little data, though not at this stage of analysis. The results
(Table 8.8) show that in the time of GATE1, the mean daily advec-
tion velocity estimates into the WE direction VWE fluctuate from 64
km/hr (18 m/sec) westward to 30 km/hr (8 m/sec) eastward with a
predominance of the westward transport of rain,
As the longitude-temporal model (Table 8.7) with zero param-
eters outside the main diagonal makes clear (without any calcula-
tions), the mean advection velocity into the NS direction for the time
interval of GATE1 was close to zero. The estimates of the mean daily
advection velocities into the NS direction (Table 8.8) fluctuate from
36 km/hr (10 m/sec) southward to 50 km/hr (14 m/sec) northward.
We can see that the wind was unstable during the GATE1 experi-
ment; and, as a result, the mean advection velocity in both directions
is close to zero. Therefore, it is interesting to estimate the rain advec-
tion velocity for the observations with stable direction of wind, which
can be accomplished by studying the PRE-STORM precipitation, as
shown in the following sections of this chapter.
8.2 PRE-STORM Precipitation 331

Notice here that, according to Monin and Yaglom (1973), the


advection velocity is practically coincident with the instantaneous
flow velocity, that is, with the wind velocity. By simultaneously
analyzing the wind and precipitation satellite data, it is possible to
verify and compare the results, if one realizes that the estimates of
the spatial correlation function of rain (Figure 8.4) show the lines
of transport of rain (or the isolines of the predominant wind) in the
area, while the AR models and diffusive description give the direction
and value of the mean wind velocity of the rain fluctuations.
Further development of this approach (by incorporating the sec-
ond spatial coordinate into the diffusion equation) will promote the
creation of a powerful methodology for describing the transport of
rain in a real three-dimensional space. Based on these results, the
problem of applying such a methodology to the mutual verification
of the rain and wind satellite observations can be formulated.

8.2 PRE-STORM Precipitation


This section presents the results of the two-dimensional spectral and
correlation analysis and multivariate modeling of PRE-STORM rain-
fall gauge data (Polyak et al., 1994).
Most of the previous analyses of rain have been based on the
GATE data. There is a need to study other sources of rain observa-
tions to ascertain whether there is a difference between the conclu-
sions drawn from GATE analyses and those based upon data from
other areas such as subtropical or midlatitude land areas. One in-
teresting experiment provides observations that can be used for such
an analysis—the PRE-STORM (Preliminary Regional Experiment
for STORM-Central) data from the Kansas/Oklahoma region. Dur-
ing the experiment the stable eastward or southeastward wind that
enable us to investigate horizontal propagation of rain anomalies in
these directions had been observed.
The intent of this study is, first, to develop a strategy for spatial
averaging of rainfall observations, and second, to explore the impli-
cations of these results for the estimation of the diffusion equation
coefficients.

8.2.1 Data
The PRE-STORM field experiment took place from May through
June of 1985. Aggregated five-minute rainfall observations were ob-
tained by the system of 42 Portable Automated Mesonet (PAM)
332 8 Second Moments of Rain

Table 8.9: Mean characteristics of PRE-STORM rainfall data.


Characteristic May June

Mean number of onsets of rain 35 63


Mean duration of rain 0.21hr 0.31hr
Mean interval between onsets of two consecutive rains 38hr 27hr
Mean relative frequency of observations with rain 0.004 0.008
Mean first autocorrelation coefficient (for 5 mm lag) 0.52 0.66
Mean second autocorrelation coefficient (for 10 min lag) 0.43 0.44

First autocorrelation coefficient (for 5 min lag)


of spatially averaged time series 0.69 0.81
Second autocorrelation coefficient (for 10 min lag)
of spatially averaged time series 0.48 0.61

stations, most of which were situated in Kansas, and 42 Stationary


Automated Mesonet (SAM) stations situated in Oklahoma. PRE-
STORM rain observations were described in detail by Meitin and
Cunning (1985) and analyzed by Graves et al. (1993). Considera-
tion in the present study is limited to rainfall data collected at 40
PAM stations as shown in Figure 8.14(a). Missing and spurious data
are replaced by zero.
Each of the time series has a very special property: More than
99% of their terms are zero. For a preliminary analysis, some statis-
tics were evaluated separately for May and June data. The estimates
for the point gauges of different stations vary significantly. Table 8.9
shows that the rains in June occurred, on average, almost two times
more often than in May, and their mean duration was 1.5 times as
long. A shortage of data in May is accompanied by higher variability,
and it is expected that the accuracy of different June statistics must
be higher than May statistics. As the values at the bottom of Table
8.9 show, spatial averaging over all points of rainfall observations
leads to an increase of the first two autocorrelations.

8.2.2 Fields of Monthly Sums


Sums of rainfall data were found for each point gauge for May and
June separately. The resulting two fields of monthly sums are consid-
ered as samples of the two-dimensional random field. The advantages
8.2 PRE-STORM Precipitation 333

Figure 8.14: (a) PRE-STORM mesonet sites over the central part of the United
States. The circles indicates the PAM stations, and the crosses indicates the SAM
sites, (b) The specific regions for each spatial-averaging scheme for the PAM
stations.
334 8 Second Moments of Rain

of such consideration are that there are no gaps in the monthly sums
and that the standard statistical technique can be applied to their
spectral analysis.
Figure 8.15 presents the periodograms of the fields. The values
have a random character, but a noticeable concentration of a signif-
icant power along the east-west axis can be seen. The estimation
of the spectra was done by applying a two-dimensional Tukey filter.
Two-dimensional spectra (Figure 8.16) show the main features of the
wave number composition of these fields, but their resolutions are too
low to reveal the details of the spectral structure.
Only the general picture is clear: Powerful westerly spatial prop-
agation waves dominate all other waves. Meridional propagations
are significantly smaller.
Correlation functions (Figure 8.17) show that for distances of
about 50 km the space correlation coefficient can reach values of
about 0.35 to 0.45.
In spite of limitations in spatial resolution, the spatial depen-
dence of monthly rainfall fluctuations is clear. Thus, the well-known
synoptic eastward transport of rain clouds is reflected in the spectral
representation of the second moments of monthly sums of rainfall.
This analysis also revealed that the rain data employed are rep-
resentative samples of a multivariate random process.

8.2.3 Multivariate Models


Interpretations of the parameters of the multivariate stochastic mod-
els of spatially and temporally averaged data can provide the basis
for a statistical description of a transport of rainfall fluctuations.
As in the case of GATE data, the signal detection may be enhanced
through spatial averaging and time aggregation of observational data
with consecutive fitting of multivariate autoregressive models to the
resulting time series.
Three different spatial averaging strategies are illustrated in Fig-
ure 8.14(b). The first strategy—averaging within four longitude
bands is intended for the detection of west-east interdependence.
The second—averaging within five latitude bands—is intended to
explore the extent of meridional propagations of rainfall perturba-
tions across latitude zones. The third strategy is useful for the anal-
ysis of southeastward transport. Five (5, 15, 30, 45, and 60 min)
temporal aggregation intervals were employed. Fifteen models were
constructed from the combinations of spatial averaging and temporal
aggregation assumptions. To reduce the seasonal contribution to the
8.2 PRE-STORM Precipitation 335

Figure 8.15: Two-dimensional periodograms of monthly sums of May (top) and


June (bottom) rainfall fields of PAM stations (X is north-south direction; Y is
west-east direction).
336 8 Second Moments of Rain

Figure 8.16: Estimates of two-dimensional spectra, computed by smoothing


the periodograms shown in Figure 8.15.

rainfall perturbation analyzed, these models were built separately for


May and June data. The most interesting results are presented here.
Each particular model is identified by number, month, and aggre-
gation, where "number" corresponds to one of the three strategies in
Figure 8.14(b), "month" is May (m) or June (j), and "aggregation" is
the aggregation interval. The aggregation time determines the time
step between the two terms ut+1 and ut. Matrices of the estimated
parameters of all the models are given in Tables 8.10, 8.11, and 8.12.
Each row of the tables with parameter estimates is associated with
an autoregressivc equation.
Consider, for example, model (l,m,5), which is fitted to the data,
as indicated in Table 8.10, The values listed in the columns e are the
8.2 PRE-STORM Precipitation 337

Figure 8.17: Estimates of two-dimensional correlation functions of monthly


sums of May (top) and June (bottom) rainfall fields of PAM stations, obtained
by Fourier transform of the spectrum fields shown in Figure 8.16 (X is north-south
direction; Y is west-east direction).
338 8 Second Moments of Rain

standard deviations of the forecast errors for one step ahead. The
autoregressive equations corresponding to model (l,m,5) are shown
below.

We would expect to find a statistical dependence of the rainfall fluc-


tuations (for each region of averaging) upon the past rainfall fluc-
tuations of the western regions. In the matrices of parameters for
the various models, this statistical dependence should manifest in
the closeness to zero of the parameter estimates above the main di-
agonals of the matrices. For May data—when aggregation intervals
arc small (5 and 15 min)—this effect is clear but not strong. For
temporal aggregations greater than 15 minutes, however, either pa-
rameter values above main diagonals are zeros or the number of sta-
tistically significant estimates (of above-main diagonals) is less than
the number below. For the models of June data, the results are more
dramatic, especially for the 30 and 45 minute aggregation intervals
when all the estimates above the main diagonals are zeros. We now
consider the five-variate models (Table 8.11) for zonal spatial aver-
aging [see Figure 8.14(b2)]. In this case, a large number of small
and zero estimates above main diagonals would be evidence that the
rainfall fluctuations from the north influence the rainfall fluctuations
of regions to the south.
However, the estimates in Table 8.11 do not suggest such an
influence. On the whole, it seems that the numbers of zeros above
and below main diagonals are approximately the same.
Models of the third type (Table 8.12)—based on diagonal band
spatial averages [see Figure 8.14(b3)]—provide a clear quantitative
picture of the strong dependence of rainfall fluctuations of any region
upon the southeastward advection of rain clouds. All but one of the
parameters above the main diagonals of June models and most such
parameters of May models are zeros.
Rainfall fluctuations of each region with aggregation intervals
larger than 15 minutes are partially determined by their own past
(parameters on the main diagonals are the largest) and by the his-
tory of fluctuations of the closest region to the northwest (the values
of the parameters of the lower second diagonals are not zeros). So,
on average, southeastward transport of rain clouds in May and June
of 1985 finds its reflection in rainfall statistics of the second mo-
ments when appropriate spatial averaging and temporal aggregation
are used.
8.2 PRE-STORM Precipitation 339

Table 8.10: Estimates (exceeding the 99% significance level) of pa-


rameters of four-variate rainfall autoregressive models for spatial av-
eraging scheme 1 [see Figure 8.14 (bl)].
May June
Parameters * Parameters *
(l,m,5) (l,j,5)

0.63 0.09 0.04 _ 0.76 0.72 0.09 _ 0.67


0.23 0.50 0.08 — 0.80 — 0.77 — 0.11 0.62
-0.08 0.12 0.56 0.05 0.79 0.05 0.05 0.75 — 0.64
-~-
— 0.60 0.80 — — 0.07 0.67 0.73

(l,m,15) (1,j.15)
_ __
0.49 0.13 -0.06 0.86 0.65 0.08 0.74
— 0.64 -0.23 — 0.78 0.09 0.60 — — 0.78
— -0.06 0.50 0.14 0.85 — 0.14 0.66 — 0.71
— — — 0.58 0.81 0.10 -0.14 0.21 0.56 0.77

(l,m,30) (l,j,30)

0.37 0.12 0.90 0.66 _ _ _ 0.73


0.23 0.41 -0.14 — 0.86 0.24 0.49 — — 0.81
— 0.27 0.42 — 0.83 — 0.21 0.63 — 0.71
— — 0.18 0.68 0.68 0.09 -0.15 0.28 0.58 0.71

(l,m,45) (l,j,45)
_ _
0.21 0.18 -0.15 0.94 0.59 0.79
0.26 0.38 — — 0.86 0.32 0.57 — 0.71

— 0.33 0.47 — 0.79 — 0.35 0.60 — 0.64
0.10 -0.15 0.35 0.64 0.65 — — 0.40 0.52 0.68

(l,m,60) (1,j,60)
0.31 0.12 .._ _ 0.93 0.50 _ _ _ 0.85
0.41 0.33 -0.15 — 0.80 0.26 0.58 -0.26 0.13 0.75
— 0.43 0.59 -0.18 0.58 0.14 0.34 0.54 — 0.68
0.22 -0.17 0.45 0.41 0.74 0.42 0.43 0.68
340 8 Second Moments of Rain

Table 8.11: Estimates (exceeding the 99% significance level) of pa-


rameters of five-variate rainfall autoregressive models for spatial av-
eraging scheme 2 [see Figure 8.14 (b2)].

May
Parameters | e
(2,m,5)

0.64 -0.06 6 0.04 0.77


— 0.48 -0.08 8 0.12 0.83
._.. 0.09 0.55 0.04 __. 0.82
0.04 -0.10 0 -0.04 4 0.54 0.08 0.82
— 0.16 0.04 -0.07 0.55 0.76

(2,m,15)

0.39 0.16 0.90


0.15 0.51 — — — 0.82
— 0.32 0.28 — -0.10 0 0.91
0.06 0.15 __ 0.50 -0.16 6 0.87
0.14 0.11 0.09 0.46 0.80

(2,m,30)
_
0.34 0.10 0.28 0.86
0.30 0.34 __ — 0.11 0.82
_. 0.67 0.17 0.14 -0.255 0.80
0.13 0.16 0.38 -0.122 0.90
0.17 0.24 0.12 0.15 0.19 0.85

(2,m,45)

0.27 0.12 0.21 0.15 0.88


0.44 0.41 — 0.80
__ 0.98 0.15 — -0.59 9 0.75
_._ -0.211 0.35 0.20 0.14 0.90
0.34 — 0.16 0.40 0.83

(2,m,60)

0.23 0.13 0.21 0.88


0.45 0.48 -0.14 4 0.12 -0.13 3 0.78
0.23 0.47 0.23 _ -0.144 0.78
0.35 0.32 0.49 -0.47 7 0.78
0.35 — 0.15 0.27 0.85

(Contiuned)
8.2 PRE-STORM Precipitation 341

Table 8.11 (Cant.)


June
Parameters *
(2j,5)

0.60 0.11 0.07 -0.09 -0.07 0.74


0.72 0.07 — — 0.66
— 0.08 0.71 — — 0.68
-0.11 — 0.11 0.63 — 0.74
-~ —- 0.08 — 0.68 0.71

(2,j,15)

0.37 0.19 0.09 0.86


0.16 0.45 0.17 — — 0.77
0.08 0.53 — — 0.81
-0.14 - 0.26 0.39 0.06 0.83
-0.09 0.08 — — 0.44 0.89

(2j,30)

0.20 0.27 0.18 _ 0.11 0.86


0.36 0.27 0.36 — — 0.65
— 0.18 0.33 0.18 — 0.83
-._ — 0.17 0.39 — 0.86
— — — — 0.32 0.94

(2,j,45)

0.46 0.15 0.15 0.79


0.33 0.34 0.25 — 0.14 0.65
— — 0.47 0.24 — 0.72
-0.13 0.15 0.20 0.36 — 0.85
— — — 0.43 0.88

(2j,60)

0.38 0.20 0.16 0.80


0.37 0.15 0.42 -0.15 0.12 0.67
-- — 0.34 0.30 — 0.76
__. — 0.24 0.45 — 0.79
-0.16 — __
0.21 0.39 0.88

Consider the model (3, j, 30) of data spatially averaged over ten-
gauge regions, as shown in Figure 8.14(b3), and aggregated over 30
minute time intervals (6 sequential points). The autocorrelation esti-
mates for the first two lags are represented in Table 8.13. Statistically
significant estimates of space-time correlations reach values of 0.33
and 0.39, for example, and allow us to build a stochastic model.
The constructed models describe the behavior of rainfall fluc-
tuations with different aggregations. They show that in May and
342 8 Second Moments of Rain

Table 8.12: Estimates (exceeding the 99% significance level) of pa-


rameters of four-variate rainfall autoregressive models for spatial av-
eraging scheme 3 [see Figure 8.14 (b3)].

May June
Parameters | s Parameters «
(3,m,5) (3j,5)

0.68 0.05 __„ -0.05 0.73 0.77 -0.03 0.63


— 0.58 — 0.81 0.06 0.80 — — 0.59
— — 0.65 — 0.76 — — 0.77 — 0.63
-0.12 0.05 0.04 0.63 0.76 — -0.06 — 0.69 0.71

(3,m,15) (3J.15)

0.57 -0.11 _ 0.81 0.61 _ __ _ 0.78


0.45 0.06 — 0.88 0.16 0.64 — — 0.71
0.07 0.49 0.06 0.86 -0.08 0.12 0.62 — 0.75
-0.14 0.14 0.52 0.81 — -0.08 0.08 0.56 0.81

(3,m,30) (3,j,30)

0.46 0.88 0.55 0.85


0.22 0.43 0.12 0.85 0.27 0.59 — — 0.69
0.23 0.19 0.44 — 0.82 — 0.25 0.49 — 0.81
— — 0.22 0.59 0.73 0.12 -0.11 0.17 0.52 0.80

(3,m,45) (3,j,45)

0.25 -0.12 0.96 0.51 0.84


0.41 0.55 — .._ 0.71 0.35 0.52 — — 0.69
0.30 0.27 0.43 __ 0.78 — 0.38 0.55 — 0.67
0.24 0.62 0.69 — -0.14 0.34 0.52 0.76

(3,m,60) (3,J,60) 3,j,60


1 1
0.28 _ __ 0.96 0.32 __ _ _ 0.93
0.55 0.35 __ 0.74 0.44 0.43 — — 0.71

0.30 0.29 0.47 — 0.75 — 0.44 0.42 — 0.74
0.20 0.40 0.34 0.79 — 0.43 0.49 0.71
8.2 PRE-STORM Precipitation 343

Table 8.13: Matrix autocorrelation function for the first two lags (0
and 30 min) of the model (3, j, 30).

Ul M2 U3 u4

T= 0

uI 1.00 0.29 -0.11 0.20


u2 0.29 1.00 0.17 -0.07
M3 -0.11 0.17 1.00 0.08
u4 0.20 -0.07 0.08 1.00

T = 1 (30 min)

uI 0.52 0.42 0.00 0.18


u2 0.11 0.68 0.33 -0.09
M3 -0.08 0.08 0.54 0.18
u4 0.07 -0.04 0.01 0.56

especially in June, when it rained more often, such fluctuations were


most strongly dependent upon advection from the northwest. The
computed parameter matrices of multivariate autoregressive models
for longitudinal and diagonal band averaging (with temporal aggre-
gations of a half hour or more) have a clear and stable triangular
form with zero elements above main diagonals. Therefore, as in the
case of GATE data, appropriate spatial averaging and the approxi-
mation of multiple time series by multivariate autoregressive models
has the potential
s to detect the structure of second-moment rainfall
fluctuations. s

8.2.4 Diffusion
With sets of the AR parameter estimates, we can compute (exactly
the same way as in Section 7.5) the diffusion equation coefficients
(for temporal aggregations of a half hour or more and for eastward
and southeastward horizontal atmospheric motions).
Consider the numerical scheme (7.2) to (7.8). We assume that
h & 100 km (the approximate distance between the centers of the
344 8 Second Moments of Rain

regions of averaging) and c = 0. The value of b was taken as the mean


of the four estimates on the main diagonals of the parameter matrices
(Tables 8.10 and 8.12) of corresponding models for the regions of
averaging under schemes 1 and 3 [Fig. 8.14(b)]. The value of a is
the mean of the three estimates on the diagonals below the main
diagonals of the same models. The results of such computations are
shown in Table 8.14. The advection velocity v varies between 34 and
48 km/hr with a mean of 44 km/hr (about 12 m/s). Such a result,
as it seems, is a reasonable approximation of the mean velocity of the
synoptic fronts for the considered region of the United States. Also
note that the value of the diffusion coefficient d = y/s varies from 41
to 49 with mean of 47. If the mean values in Table 8.14 are accepted
as the estimates of the coefficients of the diffusion equation, we get
the following:

Therefore, the mean values of the diffusion equation coefficients can


be determined from appropriate modeling of averaged and aggre-
gated systematic rainfall data when the wind direction is stable.

8.3 Final Remarks


The formalization of the rainfall statistical description can be achieved
by applying standard techniques of multivariate modeling and mul-
tidimensional spectral and correlation analysis to data averaged spa-
tially and temporally. The analysis can be briefly summarized as
follows.
1. The fitted simple linear stochastic rain field models (as well
as two dimensional spectral and correlation estimates) can be
used in the ground truth problem and in the comparative cli-
matological studies connected with statistical inferences con-
cerning the rain distribution parameters.
2. The structure of the multivariate AR model parameter matrix,
as well as that of the two-dimensional spectrum and correlation
function, paints a clear and stable picture of the predominance
of the propagation of the GATE1 rain rate fluctuations along
the parallels (as well as the eastward and southeastward wave
propagation of the PRE STORM rainfall fluctuations).
8.3 Final Remarks 345

Table 8.14: Estimates of the coefficients of the diffusion equation for


different months, regions, and aggregation intervals /(hours).

/ a b V 9

Region 1, May

0.50 0.23 0.47 48 46 0.60


0.75 0.31 0.42 45 41 0.36
1.00 0.43 0.41 46 43 0.16

Region 1, June

0.50 0.24 0.59 49 48 0.34


0.75 0.36 0.57 49 48 0.09
1.00 0.34 0.51 41 34 0.15

Region 3, May

0.50 0.21 0.48 46 42 0.62


0.75 0.31 0.46 45 41 0.31
1.00 0.41 0.36 45 41 0.23

Region 3, June

0.50 0.23 0.54 48 46 0.46


0.75 0.36 0.54 49 48 0.13
1.00 0.44 0.42 47 44 0.14

Mean 0.32 0.48 47 44 0.30


346 8 Second Moments of Rain

3. The comparison of the two-dimensional spectral and correlation


functions of the entire GATE1 data and the 12-hour sample
shows that data obtained from only one satellite is insufficient
for making accurate climatological generalizations concerning
the second moments of rain rate.
4. The diurnal cycle of rain is one of the causes for the possible
bias in the mean rain statistics obtained with limited satellite
data, and it must be taken into account in any comparison
of ground and satellite observations. The diurnal cycle of the
GATE1 rain data carries a relatively simple sinus wave shape,
with minimum values occurring during the night and maximum
values during the afternoon. The range of variation (3.4 to
5.4 mm/hr) is relatively small, compared with the standard
deviation (about 7 mm/hr) of observations.
5. The complicated character of the anisotropy of the spatial rain
fields precludes the possibility of transforming them into an
isotropic form by the standard procedures of scaling and ro-
tating coordinates. This is one of the reasons that any rain
model that can accurately describe both large-scale spatial
and temporal motions and the diurnal cycle—and also dis-
tinguish between the WE and NS transports—will necessarily
be extremely complicated. No simple (one or two parameter)
stochastic rain model that can describe the above-mentioned
features of rain exists at present.
6. The numerical approximation of the diffusion equation that
generates the fluctuation field by multivariate AR process has
led to the solution of the corresponding inverse problem and ob-
tained the mean advection velocity of the observed rain fields
(for example, about 12 m/s for the PRE-STORM precipita-
tion).
7. Based on the estimates of the advection velocity and of the
spatial correlation function, the methodology of mutual verifi-
cation of rain and wind satellite observations can be developed.
8. Comparing the diverse statistical estimates of this section with
the very simple (white-noise-type) model for the precipitation
climate time series in Chapter 6, one can understand the out-
put of large-scale temporal averaging of meteorological obser-
vations.
8.3 Final Remarks 347

Therefore, the results obtained can be used for a comparative clima-


tological study of rain in the statistical inference procedures about
parameters of the distribution of observations from different sources
of measurements and for solving some theoretical and engineering
problems.
References

Anderson, T.W., 1971: Statistical Analysis of Time Series. Wiley,


New York.
Arkell, R., and M. Hudlow, 1977: GATE International Meteorologi-
cal Radar Atlas. U.S. Dopt. of Commerce, NOAA Publication,
Washington D.C.
Bayley G.V., and J.W. Hamrnersley, 1946: The "Effective" Number
of Independent Observations in an Autocorrelated Time Series.
J. Roy. Stat. Soc., 8, 29-40.
Bell, T.L., and N. Reid, 1993: Detecting the Diurnal Cycle of Rain-
fall Using Satellite Observations. ,/. Applied Meteorology,
32, 311-322.
Berezin, I.S., and N.P. Zhidkov, 1965: Numerical Calculations. Per-
gamon, Oxford.
Bider, M., M. Schiiepp, and H. Rudolif, 1959: Die Reduktion der
200 jahrigen Easier Temperaturreihe. Arch. Meteorol., Geophis.
u. Bioklimatol. Ser. B, Bd 9, H 3-4.
Bider, M., and M. Schiiepp, 1961: Luftdrnckreihen der Ictzten zwei
Jahrhunderte von Basel iind Genf. Arch. Meteorol., Geophis.
u. Bioklimatol. Ser. B, Bd 11, H 1.
Blackman, R.B., and J.W. Tukey, 1.958: The Measurement of Power
Spectra from the Point of View of Communications Engineer-
ing. Dover, New York.
Bolshev, L.N., and N.B. Srnirnov, 1965: Statistical Tables. Nauka,
Moscow. (In Russian.)
Box, G.E., and G.M. Jenkins, 1976: Time Series Analysis: Fore-
casting and Control. Holden-Day, San Francisco.
Budyko, M.I., 1974: Climate Change.. Hydrometeoizdat, 1974. (In
Russian.)

348
References 349

Cahalan, R.F., D.A. Short, and G.R. North, 1982: Cloud Fluctua-
tion Statistics. Monthly Weather Review, 110, 26-43.
Chiu, L.S., G.R. North, A. Short, and A. McConnel, 1990: Rain
Estimation from Satellites: Effect of Finite Field of View. J.
Geophys. Res., 95, 2177-2185.
Christakos, G., 1992: Random Field Models in Earth Sciences. Aca-
demic Press, San Diego.
Cooley, J.W., and J.W. Tukey, 1965: An Algorithm for the Mach-
ine Calculation of Complex Fourier Series. Mathem. Comp.
19 (90).
Craddock, J.M., and B.C. Wales-Smith, 1977: Monthly Rainfall
Totals Representing the East Midlands for the Years 1726-
1975. Meteorol. Mag. 106 (1257).
Crane, R.K., 1990: Space-Time Structure of Rain Rate Fields. J.
Geophys. Res., 95, 2011-2020.
Cressie, N., 1991: Statistics for Spatial Data. John Wiley, New
York.
Cubasch, U., K. Hasseltnann, H. Hock, E. Maier-Reimer, U. Miko-
lajewicz, B.D. Santer, and R. Sausen, 1992: Time-Dependent
Greenhouse Warming Computations with a Coupled Ocean-
Atmosphere Model. Climate Dynamics, 8: 55-69.
Daley, R., 1991: Atmospheric Data Analysis. Cambridge Universi-
ty Press.
Eikhoff, P.E. (editor), 1983: Modern Methods of Systems' Identi-
fications. Mir, Moscow. (In Russian.)
Einstein, A., 1986: Method of Determining of the Observed Sta-
tistical Values Related to the Random Fluctuations. Izvestiya
Academy of Science, Physics of Atmosphere and Ocean,
22 (1), 99-100. (In Russian.)
Epstein, E.S., 1985: Statistical Inference and Prediction in Clima-
tology: a Bayesian Approach. American Meteorological Soci-
ety, Boston.
Essenwanger, 0., 1989: Applied Statistics in Atmospheric Sciences.
Elsevier, Amsterdam.
Flon, G., 1977: History and Climate Intransitivity. In: Physical
Foundation of Climate Theory and its Modeling. Hydrome-
teoizdat, Leningrad. 114-124. (In Russian.)
Gandin, L.S., 1965: Objective Analysis of Meteorological Fields.
Program for Scientific Translations. Jerusalem, Israel.
Gandin, L.S., R.L. Kagan, 1976: Statistical Methods of Meteoro-
logical Data Interpretation. Hydrometeoizdat, Leningrad. (In
Russian.)
350 References

Graves, C.E, J.B. Valdes, S.S.P. Shen, and G.R. North, 1993: Eval-
uation of Sampling Errors of Precipitation from Spaceborne
and Ground Sensors. J. Applied Meteorology, 32 (2), 374-385.
Gruza, G.V. (editor), 1987: Climate Monitoring Data. Goskomhy-
drometizdat, Moscow. (In Russian.)
Guyon, X., 1982: Parameter Estimation for a Stationary Process
on a d-Dimensional Lattice. Biometrika, 69, 95-105.
Hamming, R.W., 1973: Numerical Methods for Scientists and En-
gineers. McGraw-Hill, New York.
Hannan, E.J., 1970: Multiple Time Series. Wiley, New York.
Hassclmann, K., 1976: Stochastic Climate Models. Tellus, 28 (6),
473-485.
Hasselmann, K., 1988: PIPs and POPs: The Reduction of Complex
Dynamical Systems using Principal Interaction and Oscillation
Patterns. ,J. Geophys. Res., 93, 11015 11021.
Hays, J.D., J. Imbrie, and N.J. Shackleton, 1976: Variation in the
Earth's Orbit: Pacemaker of Ice Ages. Science, 194 (4270).
Hudlow, M.D., 1977: Precipitation Climatology for the Three Phases
of GATE. Preprints Second Conf. Hydrometeorology, Toronto,
Amer. Meteor. Soc., 290-297.
Hudlow, M.D., 1979: Mean Rainfall Pattern for the Three Phases
of GATE. J. Appl. Meteor., 18, 1656-1668.
Jenkins, J., and D. Watts, 1968: Spectral Analysis and its Applica-
tions. Holden-Day, San Francisco.
Jones, P.O., S.C.B. Raper, U.S. Bradly, H.F. Diaz, P.M. Kelly, and
T.M.L. Wigley, 1986: Northern Hemisphere Surface Air Tem-
perature Variations, 1851-1984. J. dim. Appl. MeteoroL, 25,
161 179.
Jones, R.H., and A.V. Vecchia, 1993: Fitting Continuous ARMA
Models to Unequally Spaced Spatial Data. J. American Sta-
tistical Association, 88 (423), 947-954.
Kagan, ILL., 1979: Averaging of the. Meteorological Fields. Hy-
drorneteoizdat, Leningrad. (In Russian.)
Kashyap, ILL., and A. Rao, 1976: Dynamic Stochastic Models from
Kmpirical Data. Academic Press, New York.
Katz, R.W., 1992: The Role of Statistics in the Validation of Gen-
eral Circulation Models. Climate Research, 2, 34-45.
Kendall, M., and A. Stuart, 1963: The Advance Theory of Statistics:
Distribution Theory. Hafner, New York.
Kendall, M., and A. Stuart, 1967: The Advance Theory of Statistics:
Inference and Relationship. Hafner, New York.
Lanczos, C., 1956: Applied Analysis. Prentice-Hall, Englewood
Cliffs, N.J.
References 351

Landsberg, H.E., J.M. Mitchell, and ILL. Gruntcher, 1959: Power


Spectrum Analysis of Climatological Data for Woodstock Col-
lege, Maryland. Month. Weath. Rev., 87 (8).
Lebrijn, A., 1954: The Climate of the Netherlands During the Last
Two and a Half Centuries. Koninklijk Nederlandsch Meteorol-
ogisch Institut, 102 (49), 94-99.
Lepekhina, N.A., and E.I. Fedorchenko, 1972: Temporal Structure
of Temperature Time Series. Trudy Main Geophysical Obser-
vatory, 286. (In Russian.)
Loeve, M., 1960: Probability Theory. Van Nostrand, Princeton,
N.J.
Lorenz, E.N., 1977: Climate Predictability. In: Physical Founda-
tion of Climate Theory and its Modeling. Hydrometeoizdat,
Leningrad. 137-141. (In Russian.)
Madden, R.A., 1977: Estimates of the Autocorrelation and Spec-
tra of Seasonal Mean Temperatures over North America.
Mon. Wea. Rev., 105 (1), 9-18.
Manley, G., 1974: Central England Temperatures: Monthly Means
1659 to 1973, Quart. J. Roy. Meteor. Soc., 100 (425), 389-
405.
Matrosova, L.E., and I. I. Polyak, 1990: Presentation of the Upper
Atmosphere Temperature in Stochastic Models. Izv. Acad.
Sci., U.S.S.R., Physics of Atmosphere and Ocean, 26 (9), 925-
934.
McConnell, A., and G.R. North, 1987: Sampling Errors in Satellite
Estimates of Tropical Rain. J. Geophys. Res. 92 (D8), 9567-
9570.
Meitin, J., and J. Cunning, 1985: The Oklahoma-Kansas Prelim-
inary Regional Experiment for Storm-Central. Vol. I: Daily
Operations Summary. NOAA Technical Memorandum ERL
ESG-20.
Mitchell, J.M., 1976: An Overview of Climatic Variability and Its
Causal Mechanisms. Quart. Res., 6 (4).
Monin, A.S., 1969: Weather Forecasts as a Physical Problem. Nauka,
Moscow. (In Russian.)
Monin, A.S., 1972: The fJarth's Rotation and Climate. Hydrome-
teoizdat, Leningrad. (In Russian.)
Monin, A.S., 1977: The Earth History. Nauka, Moscow. (In Rus-
sian.)
Monin A.S., and I.L. Vulis, 1971: On the Spectra of Long-Period
Oscillations of Geophysical Parameters. Tellus, 23 (4-5).
Monin, A.S., and A.M. Yaglom, 1973: Statistical Fluid Mechanics.
352 References

MIT Press, Cambridge.


Nakamoto, S., J.B. Valdes, and G.R. North, 1990: Frequency-
Wavenumber Spectrum for GATE Phase I Rain Fields. J. Appl.
Meteor., 29, (9), 842-850.
North, G.R., and S. Nakamoto, 1989: Formalism for Comparing
Rain Estimation Designs. J. Atmos. Ocean. Tech., 6, 985-
992.
North, G.R, and K.Y. Kim, 1991: Surface Temperature Fluctua-
tions in Stochastic Climate Models. J. Geophysical Research,
96, (D10), 18573-18580.
Oort,A.IL, and Rasimisson, 197 L: Atmospheric Circulation Statis-
tics. NOAA Professional Paper No. 5, U.S. Govt. Printing
Office, Washington, B.C.
Oort, A.If., 1978: Adequacy of the Radiosonde Networks for Global
Circulation Study Tested Through Numerical Model Output.
Monthly Weather Review, 106, 107-115.
Oort, A.H., 1983: Global Atmospheric Circulation Statistics, 195S-
1973.NOAA Professional Paper No. 14, U.S. Govt. Printing
Office, Washington, B.C.
Palmen, E., and C.W. Newton, 1969: Atmospheric Circulation Sys-
tems. Academic Press, New York.
Panofsky, H.A., and G.W. Brier, 1958: Some Applications of Statis-
tics to Meteorology. The Pennsylvania State University.
Parzen, E., 1966: Time Series Analysis for Models of Signal Plus
White Noise. In Spectral Analysis of Time Series, Proceedings.
J.Wiley and Sons, New York. pp. 233-258.
Parzen, E. (editor), 1984: Time Series Analysis of Irregularly Ob-
served Data. Proceedings of a symposium held at Texas A&M
University, College Station, Texas, February 10—13, 1983. Spri-
nger-Verlag, New York.
Patterson, V.L., M.D. Hudlow, P.J. Pytlowany, F.P. Richards, and
J.D. Hoff, 1979: GATE Radar Rainfall Processing System,
NOAA Tech. Memo. ED1S 26, Washington, D.C.
Polyak, 1.1., 1975: Numerical Methods for Data Analysis. Hydrom-
eteoizdat, Leningrad. (In Russian.)
Polyak, 1.1., 1979: Methods for Random Processes and Fields Anal-
ysis. Hydrometeoizdat, Leningrad. (In Russian).
Polyak, 1.1., 1989: Multivariate Stochastic Models of Climate. Hy-
drometeoizdat, Leningrad. (In Russian).
Polyak, I.I., 1992: Multivariate Stochastic Climate Models. 12th
Conference on Probability and Statistics in the Atmospheric
Sciences. Toronto, pp.273-276.
References 353

Polyak, 1.1., Z.I. Pivovarova, and L.V. Sokolova, 1979: Long-Period


Variations of the Solar Radiation. Trudy of the Main Geophys-
ical Observatory. No. 428.
Polyak, I.I., G.R. North, and J. B. Valdes, 1994: Multivariate Space-
Time Analysis of PRE-STORM Precipitation. J. of Applied
Meteorology. 33 (9), 1079-1087.
Polyak, I.I., and G.R. North , 1995: The Second Moment Climatol-
ogy of the GATE Rain Rate Data. Bulletin of the American
Meteorological Society, 76 (4), 535-550.
Polyak, 1.1., 1996: Observed versus Simulated Second-Moment Cli-
mate Statistics in the GCM Verification Problems. J. Atmo-
spheric Sciences, 53 (5), 677-694.
Preisendorfer, R.W., and C.D. Mobley, 1988: Principal Component
Analysis in Meteorology and Oceanography, Elsevier, New York.
Raibman, N., 1983: Methods of Non-Linear and Minimax Identi-
fication. In: Modern Methods of System Identification. Mir,
Moscow. (In Russian.)
Rao, R., 1973: Linear Statistical Inference and Its Applications.
Wiley, New York.
Roeckner E., 1989: The Hamburg Version of the ECMWF Model
(ECUAM). G.J. Boer, editor. CAS/JSC Working Group on
Numerical Experimentation, Res. Act. Atm. and Ocean Mod-
eling, Rep. 13, WMO/TD-322, pp. 7.1-7.4.
Santer, B.D. and T.M.L. Wigley, 1990: Regional Validation of Means,
Variances, and Spatial Patterns in General Circulation Model
Control Runs. J. Geophysical Res., 95, 829-850.
Santer, B.D., W. Bruggemann, U. Cubasch, K. Hasselmann, H.
Hock, E. Maier-Reimer, and U. Mikolajewicz, 1994: Signal-to-
Noise Analysis of Time-Dependent Greenhouse Warming Ex-
periments. Climate Dynamics, 9, 267-285.
Schmugge, T., 1977: Remote Sensing of Surface Soil Moisture. Pre-
prints Second Conf. Hydrometeorology, Toronto, Amer. Me-
teor. Soc., 304-310,
Simpson, J., R. Adler, and G.R. North, 1988: On Some Aspects
of a Proposed Tropical Rainfall Measuring Mission (TRMM).
Bull. Amer. Meter. Soc., 69, 278-295.
Thiebaux, H.J., 1994: Statistical Data Analysis for Ocean and At-
mospheric Sciences. Academic Press, New York.
Tong, H., 1990: Non-Linear Time Series. A Dynamical System
Approach. Oxford Science Publications, Oxford.
TutubaUn, V.N., 1972: The Probability Theory. Moscow University.
(In Russian.)
354 References

Understanding Climate Change. A Program for Action., 1975: Na-


tional Academy of Sciences. Washington, D.C.
van der Waerden, B.L., 1960: Mathematical Statistics. Moscow
Publishing House of Foreign Literature. (In Russian.)
Wigley, T.M.L., and B. D. Santer, 1990: Statistical Comparison
of the Spatial Fields in Model Validation, Perturbation, and
Predictability Experiments. J. Geophy. Res., 95, 851-865.
Wilks, D.S., 1995: Statistical Methods in the Atmospheric Sciences.
Academic Press, New York.
World Weather Records, 1975: U.S. Department of Commerce, NOAA,
National Climatic Center, Asheville, N.C.
Yaglom, A.M., 1986: Correlation Theory of Stationary and Related
Random Functions. Springer-Veiiag, New York.
Index

advection velocity 308 climate system identification 234


GATE rain rate 3:50 climatological fields 147
PRE-STORM data 344 coefficients of polynomial 7
temperature 288 coherency function 118, 139-146
algorithm for rrniltivariate AR mod- estimator 121, 139-146
eling 218 Cooley-Tukey methodology 115
amplitude modulated time series 131 correlated observations 61
anisotropy correlation function
climatological fields 160 latitude-temporal rain rate field
rain rate field 308 316
AR model and diffusion process 285 latitude-temporal temperature fields
AR process 165 271
mnltivariate 208 point estimates 27, 32
AR(1) process 167 PRE-STORM rainfall fields 334
rnultivariate 212 correlation matrix 62
AR(2) process 165 correlation window 113
multivariate 223 two-dimensional 135
ARMA process 163 cospectrum 118
ARMA(1,1) process 183 covariaiice
atmospheric pressure 248 function of point estimates 32
autocovariance function 111 matrix 61
estimator 112 covariances of point estimates 26
autocorrelation function 111 cross-covariance function 118, 139-
atmosperic pressure 253 146
precipitation 255 two-dimensional 132
temperature 251 cross-periodogram 119, 139-146
autodispersion function 108 two-dimensional 133
autoregressive parameters 163 cross-spectral density 118, 139-146
two-dimensional 132
Bernoulli polynomials 10 cross statistical analysis 115
Bessel function 161 cumulative spectral function 116
Blackman-Tukey methodology 115
derivative estimation 44, 49
central England temperature 193 derivatives of observed function 14
circulation statistics 225 differentiation of fields 54
climate change 263 diffusion 287

355
356 Index

diffusion equation coefficients estima- harmonic filter 39


tion 288 harmonic structure 28
diffusion of GATE rain rate 329 harmonizable processes 125
diffusion of PRE-STORM observa- historical records 239
tions 344
diffusion of temperature 285 identification 178
digital filter 26 ARMA processes 162
discrete Fourier expansion139-146 climate system 234
two-dimensional 133 polynomial degree 20
dispersion function 107 instrumental variable 100
diurnal cycle of rain rate 318 inverse normal matrix 11
double frequency spectral density 125
Kolmogorov-Smirnov criterion 117
equivalence of filters 42
Laplace equation 160, 310
fast Fourier transform (FFT) 115
latitude-temporal
filters
rain rate field 315
correlated observations 84
temperature field 291
differentiating 44, 54
least squares method 4, 61
finite differences 91
polynomial approximations 6
harmonic 37
linear regression 99
multi-dimensional 56
observed and simulated temper-
parameters 26
ature 289
regressive 28
linear trend 70
smoothing 24
air temperature 196, 240, 245,
two-dimensional 51
atmospheric pressure 243
width 26
precipitation 243, 247
window 26
finite differences 92
MA process 179
Markov process 96
matrix correlation function 209
white noise 93
GATE data 324
Fourier coefficients 21
observed and simulated temper-
Fourier set 21
ature 280-283
Fourier transforms 111, 139 146
P.RE-STORM data 334
two-dimensional 140
matrix covariance function 209
frequency characteristics 28, 37
mean estimator 70, 75, 76
GATE data 304 menorah equation 18
GATE1 data 304 meridional circulation 288
GATE2 data 320 models
Gauss-Markov Theory 3 atmospheric circulation statistics
GCM diagnosis and verification 267 225
geopotential 147 GATE1 rain rate data 323
ground truth problem 31.1 observed and simulated temper-
atures 281
Hamburg GCM zonal temperature PRE-STORM precipitation 336
268 temperature time series 225—233
Index 357

moving average process with stationary increments


harmonic analysis 124 198
parameters 163
procedure 26 quadratic form 62
processes 179 quadrature spectrum 118
multidimensional field 138 q-variate AR process 210
filter 58 qr-variate correlation function 209
multiple correlation coefficient 165 g-variate white noise 210
multivariate AR processes 208
AR(1) process 209 regressive filters 29
AR(2) process 221
sample
nonergodic stationary process 127 cospectrum 119
nonlinear covariance function of two-dimensional
random processes 107 field 133
model 205 cross-covariance function 117
nonoptimal estimation 74 two-dimensional 136
nonstationary process 122 quadrature spectrum 119
normal equations 4, 7, 9, 62 second-order derivative
matrix 4, 11 estimation 47
normalized standard deviation of fore- of field 54
cast error 164, 167-183 separable statistical dependence 83
normalized variances signal-plus-white-noise processes 82,
forecast error 164 190
point estimates 5, 22, 63 signal ratio 84, 191
spectral density 111 climate time series 193
numerical differentiation of field 54 surface monthly mean tempera-
tures 196
Parseval identity 113, 139-146 smoothing
permissible correlations 100, 104 correlated observations 87
AR process 168 two-dimensional 51
ARMA(1,1) process 184 window 30
instrumental variable 101 solar radiation budget 196
linear regression 99 spatial averaging 79
MA process 179 spatial correlation function of GATE!
multivariate AR(1) process 213 data 306
periodogram 112, 139 146 spatial spectrum of GATEl data 306
historical records 247, 251 spatial-temporal covariance 80
PRE-STORM rainfall fields 334 spectra
phase spectrum 118, J21 altitude-temporal field 158
estimator 121, 139 446 historical records 259
point estimates 5, 22, 62 latitude-temporal climate fields
polynomial coefficients 14 150
precipitation trend over territory of latitude-temporal temperature fields
Russia 245 291
PRE-STORM data 332 monthly-annual field 154
358 Index

PRE-STORM rainfall fields 334 three Fourier transformation method


zonal temperature time series 271 115
spectral analysis of 500 nib surface time series with missing data 129
geopotential field 147 Toeplitz matrix 75
spectral density 111, 132 Tukey filter 41
spectral window 113
two-dimensional 135 univariate AR processes 165
spectrum of latitude-temporal rain univariate ARMA processes 163
rate field 315 univariate modeling of historical records
standardized autodispersion function 259
108
stationary process 111 validation of GCM 266
statistics and climate change 263 variance of forecast error 164
streamflow time series 196 variance of point estimates 6, 15, 62,
Student statistic 6 89
sum of signal and white noise 80
surface air temperature 267 window width 26
surface air temperature spectra and
Yule-Worker
autocorrelation functions 251
equations 165
symmetrical grid points 7
symmetrical indexing 7 multivariate equations 211

You might also like