Practical Data Analysis in Chemistry [1st edition] 9780080548838, 9780444530547, 0444530541

The majority of modern instruments are computerised and provide incredible amounts of data. Methods that take advantage

289 38 2MB

English Pages 1-326 [341] Year 2007

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
Content:
Preface
Pages ix-xi
Marcel Maeder

Symbols
Pages xiii-xiv

1 Introduction Original Research Article
Pages 1-6

2 Matrix algebra Original Research Article
Pages 7-28

3 Physical/chemical models Original Research Article
Pages 29-100

4 Model-based analyses Original Research Article
Pages 101-212

5 Model-free analyses Original Research Article
Pages 213-316

List of matlab files
Pages 317-320

List of excel sheets
Page 321

Index
Pages 322-326

Recommend Papers

Practical Data Analysis in Chemistry [1st edition]
 9780080548838, 9780444530547, 0444530541

  • Commentary
  • 49768
  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

DATA HANDLING IN SCIENCE AND TECHNOLOGY — VOLUME 26

Practical Data Analysis in Chemistry

DATA HANDLING IN SCIENCE AND TECHNOLOGY Advisory Editors: S. Rutan and B. Walczak

Other volumes in this series: Volume 1 Volume 2 Volume 3 Volume 4 Volume 5 Volume 6

Volume 7 Volume 8 Volume 9 Volume 10 Volume 11 Volume 12 Volume 13 Volume 14 Volume 15 Volume 16 Volume 17 Volume 18 Volume 19 Volume 20A

Volume 20B Volume 21 Volume 22 Volume 23 Volume 24

Volume 25

Microprocessor Programming and Applications for Scientists and Engineers, by R.R. Smardzewski Chemometrics: A Textbook, by D.L. Massart, B.G.M. Vandeginste, S.N. Deming, Y. Michotte and L. Kaufman Experimental Design: A Chemometric Approach, by S.N. Deming and S.L. Morgan Advanced Scientific Computing in BASIC with Applications in Chemistry, Biology and Pharmacology, by P. Valkó and S. Vajda PCs for Chemists, edited by J. Zupan Scientific Computing and Automation (Europe) 1990, Preceedings of the Scientific Computing and Automation (Europe) Conference, 12–15 June, 1990, Maastricht, The Netherlands, edited by E.J. Karjalainen Receptor Modeling for Air Quality Management, edited by P.K. Hopke Design and Optimization in Organic Synthesis, by R. Carlson Multivariate Pattern Recognition in Chemometrics, illustrated by case studies, edited by R.G. Brereton Sampling of Heterogeneous and Dynamic Material Systems: Theories of Heterogeneity, Sampling and Homogenizing, by P.M. Gy Experimental Design: A Chemometric Approach (Second, Revised and Expanded Edition) by S.N. Deming and S.L. Morgan Methods for Experimental Design: Principles and Applications for Physicists and Chemists, by J.L Goupy Intelligent Software for Chemical Analysis, edited by L.M.C. Buydens and P.J. Schoenmakers The Data Analysis Handbook, by I.E. Frank and R. Todeschini Adaption of Simulated Annealing to Chemical Optimization Problems, edited by J. Kalivas Multivariate Analysis of Data in Sensory Science, edited by T. Næs and E. Risvik Data Analysis for Hyphenated Techniques, by E.J. Karjalainen and U.P. Karjalainen Signal Treatment and Signal Analysis in NMR, edited by D.N. Rutledge Robustness of Analytical Chemical Methods and Pharmaceutical Technological Products, edited by M.W.B. Hendriks, J.H. de Boer, and A.K. Smilde Handbook of Chemometrics and Qualimetrics: Part A, by D.L. Massart, B.G.M. Vandeginste, L.M.C. Buydens, S. de Jong, P.J. Lewi, and J. Smeyers-Verbeke Handbook of Chemometrics and Qualimetrics: Part B, by B.G.M. Vandeginste, D.L. Massart, L.M.C. Buydens, S. de Jong, P.J. Lewi, and J. Smeyers-Verbeke Data Analysis and Signal Processing in Chromatography, by A. Felinger Wavelets in Chemistry, edited by B. Walczak Nature-inspired Methods in Chemometrics: Genetic Algorithms and Artificial Neural Networks, edited by R. Leardi Handbook of Chemometrics and Qualimetrics, by D.L. Massart, B.M.G. Vandeginste, L.M.C. Buydens, S. de Jong, P.J. Lewi, and J. Smeyers-Verbeke Statistical Design — Chemometrics, by R.E. Bruns, I.S. Scarminio and B. de Barros Neto

DATA HANDLING IN SCIENCE AND TECHNOLOGY — VOLUME 26 Advisory Editors: S. Rutan and B. Walczak

Practical Data Analysis in Chemistry MARCEL MAEDER School of Environmental and Life Sciences The University of Newcastle Callaghan, NSW 2308, Australia

YORCK-MICHAEL NEUHOLD School of Environmental and Life Sciences The University of Newcastle Callaghan, Australia

Amsterdam – Boston – Heidelberg – London – New York – Oxford Paris – San Diego – San Francisco – Singapore – Sydney – Tokyo

Elsevier Radarweg 29, PO Box 211, 1000 AE Amsterdam, The Netherlands Linacre House, Jordan Hill, Oxford OX2 8DP, UK

First edition 2007 Copyright © 2007 Elsevier B.V. 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 written permission of the publisher Permissions may be sought directly from Elsevier’s Science & Technology Rights Department in Oxford, UK: phone (+44) (0) 1865 843830; fax (+44) (0) 1865 853333; email: [email protected]. Alternatively you can submit your request online by visiting the Elsevier web site at http://elsevier.com/locate/permissions, and selecting Obtaining permission to use Elsevier material Notice No responsibility is assumed by the publisher for any injury and/or damage to persons or property as a matter of products liability, negligence or otherwise, or from any use or operation of any methods, products, instructions or ideas contained in the material herein. Because of rapid advances in the medical sciences, in particular, independent verification of diagnoses and drug dosages should be made Library of Congress Cataloging-in-Publication Data A catalog record for this book is available from the Library of Congress British Library Cataloguing in Publication Data A catalogue record for this book is available from the British Library ISBN: 978-0-444-53054-7 ISSN: 0922-3487

For information on all Elsevier publications visit our website at books.elsevier.com

Printed and bound in The Netherlands 07 08 09 10 11 10 9 8 7 6 5 4 3 2 1

Contents PREFACE

IX

SYMBOLS

XIII



1 INTRODUCTION

1

2 MATRIX ALGEBRA

7

2.1 Matrices, Vectors, Scalars

2.1.1 Elementary Matrix Operations

Transposition

Addition and Subtraction

Multiplication

2.1.2 Special Matrices

Square Matrix

Symmetric Matrix

Diagonal Matrix

Identity Matrix

Inverse Matrix

Orthogonal and Orthonormal Matrices

8

10

10

12

16

21

21

22

22

23

24

25

2.2 Solving Systems of Linear Equations

26

3 PHYSICAL/CHEMICAL MODELS

29

3.1 Beer-Lambert's Law

33

3.2 Chromatography / Gaussian Curves

36

3.3 Titrations, Equilibria, the Law of Mass Action

3.3.1 A Simple Case: Fe3+ + SCN­

3.3.2 The General Case, Definitions

A Chemical Example, Cu2+, Ethylenediamine, Protons

3.3.3 Solving Complex Equilibria

The Newton­Raphson Algorithm

Example: General 3­Component Titration

Example: pH Titration of Acetic Acid

Equilibria in Excel

Complex Equilibria Including Activity Coefficients

Special Case: Explicit Calculation for Polyprotic Acids

3.3.4 Solving Non­Linear Equations

One Equation, One Parameter

40

40

43

45

48

48

56

58

60

62

64

69

69

vi

Contents Systems of Non­Linear Equations

3.4 Kinetics, Mechanisms, Rate Laws

3.4.1 The Rate Law

3.4.2 Rate Laws with Explicit Solutions

3.4.3 Complex Mechanisms that Require Numerical Integration

The Euler Method

Fourth Order Runge­Kutta Method in Excel

3.4.4 Interesting Kinetic Examples

Autocatalysis

0th Order Reaction

The Steady­State Approximation

Lotka­Volterra / Predator­Prey Systems

The Belousov­Zhabotinsky (BZ) Reaction

Chaos, the Lorenz Attractor

4 MODEL-BASED ANALYSES

71

76

77

77

80

80

82

86

87

89

91

92

95

97

101

4.1 Background to Least-Squares Methods

4.1.1 The Residuals and the Sum of Squares

Linear Example: Straight Line

Non­Linear Example: Exponential Decay

102

103

103

105

4.2 Linear Regression

4.2.1 Straight Line Fit ­ Classical Derivation

4.2.2 Matrix Notation

4.2.3 Generalised Matrix Notation

4.2.4 The Normal Equations

The Pseudo­Inverse

Linear Dependence, Rank of a Matrix

Numerical Difficulties

4.2.5 Errors in the Fitted Parameters

4.2.6 Excel Linest

4.2.7 Applications of Linear Least­Squares Fitting

Linearisation of Non­Linear Problems

Polynomials, the Savitzky­Golay Digital Filter

Smoothing of Noisy Data

Calculation of the Derivative of a Curve

Polynomial Interpolation

4.2.8 Linear Regression with Multivariate Data

Applications

Computation of Component Spectra, Known Concentrations

Computation of Component Concentrations, Known Spectra

The Pseudo­Inverse in Excel

109

109

113

114

115

117

119

120

121

125

127

127

130

131

135

138

139

143

144

145

146

4.3 Non-Linear Regression

4.3.1 The Newton­Gauss­Levenberg/Marquardt Algorithm

A First, Minimal Algorithm

148

148

149

Contents Termination Criterion, Numerical Derivatives

The Levenberg/Marquardt Extension

Standard Errors of the Parameters

Multivariate Data, Separation of the Linear and Non­Linear Parameters

Constraint: Positive Component Spectra

Structures, Fixing Parameters

Known Spectra, Uncoloured Species

Reduced Eigenvector Space

Global Analysis

4.3.2 Non­White Noise, f2­Fitting

Linear f2­Fitting

Non­Linear f2­Fitting

4.3.3 Finding the Correct Model

4.4 General Optimisation

4.4.1 The Newton­Gauss Algorithm

4.4.2 The Simplex Algorithm

4.4.3 Optimisation in Excel, the Solver

f2­Fitting in Excel

5 MODEL-FREE ANALYSES

vii 153

155

161

162

168

169

175

180

183

189

190

195

197

198

198

204

207

211

213

5.1 Factor Analysis, FA

5.1.1 The Singular Value Decomposition, SVD

5.1.2 The Rank of a Matrix

Magnitude of the Singular Values

The Structure of the Eigenvectors

The Structure of the Residuals

The Standard Deviation of the Residuals

5.1.3 Geometrical Interpretations

Two Components

Reduction in the Number of Dimensions

Lawton­Sylvestre

Three and More Components

Mean Centring, Closure

HELP Plots

Noise Reduction

213

214

217

219

221

222

223

224

224

228

231

235

239

241

243

5.2 Target Factor Analyses, TFA

5.2.1 Projection Matrices

5.2.2 Iterative Target Transform Factor Analysis, ITTFA

5.2.3 Target Transform Search/Fit

Parameter Fitting via Target Testing

246

250

251

253

257

5.3 Evolving Factor Analyses, EFA

5.3.1 Evolving Factor Analysis, Classical EFA

5.3.2 Fixed­Size Window EFA, FSW­EFA

5.3.3 Secondary Analyses Based on Window Information

259

260

268

271

viii

Contents

Iterative Refinement of the Concentration Profiles Explicit Computation of the Concentration Profiles

271 276

5.4 Alternating Least-Squares, ALS 5.4.1 Initial Guesses for Concentrations or Spectra 5.4.2 Alternating Least-Squares and Constraints 5.4.3 Rotational Ambiguity

280 281 282 288

5.5 Resolving Factor Analysis, RFA

290

5.6 Principle Component Regression and Partial Least Squares, PCR and PLS 5.6.1 Principal Component Regression, PCR Mean-Centring, Normalisation PCR Calibration PCR Prediction Cross Validation 5.6.2 Partial Least Squares, PLS PLS calibration PLS Prediction / Cross Validation 5.6.3 Comparing PCR and PLS

295 296 297 298 300 303 306 308 309 310

FURTHER READING

313

LIST OF MATLAB FILES

317

LIST OF EXCEL SHEETS

321

INDEX

322

Preface The word 'practical' in the title describes a characteristic feature of this book. However, it could easily be misunderstood. It is not a book that is meant to be taken into the laboratory to remain on the lab bench next to instruments and test tubes. The book is practical insofar as every bit of theory applicable to data analysis is exemplified in a short program in Matlab or in an Excel spreadsheet. The philosophy of the book is that the reader can study the programs, play with them and observe what happens. They are short and concise and thus invite and encourage meddling and improving by the reader. Suitable data are generated for each example in short routines. This ensures the reader has a clear understanding of the structure of the data and thus will have a better chance of comprehending the analysis. In fact, the programs, rather than complex equations, are often used to elucidate the principles of the analysis. The programs are written modular. The reader can replace the artificial data, generated by a function, with real data from the lab. There is extensive use of graphical output. While the plots are minimal they efficiently illustrate the results of the analyses. In order to keep the programs concise, no effort was made to build comfortable user-interfaces. In Chapter 2, we give a brief introduction to matrix algebra and its implementation in Matlab and Excel. The next three chapters form the core of the book. We distinguish two types of data analysis: model-based and model-free analyses. For both, appropriate data have to be generated for subsequent analysis. In Chapter 3, we supply the theory required for the modelling of chemical processes. Many of the example data sets used for both kinds of analyses are taken from kinetics and equilibrium processes. This reflects the background of both authors. In fact, this part of the book serves as a solid introduction to the simulation of equilibrium processes such as titrations and the simulation of complex kinetic processes. The example routines are easily adapted to the processes investigated by the reader. They are very general and there is essentially no limit to the complexity of the processes that can be simulated. Chapter 4 is an introduction to linear and non-linear least-squares fitting. The theory is developed and exemplified in several stages, each demonstrated with typical applications. The chapter culminates with the development of a very general Newton-Gauss-Levenberg/Marquardt algorithm. Chapter 5 comprises a collection of several methods for model-free data analyses. It starts with classical Factor Analysis, employing many

x

Preface

geometrical visualisations, covers popular methods, such as EFA and ALS and concludes with a brief introduction into the calibration based methods PCR and PLS. A fair amount of effort has been put into writing short and concise but still readable code. A few highlights: lolipop.m, a very short function of 7 lines of code that performs polynomial interpolation of any degree and complexity. It can be used for interpolation and, of course with caution, for extrapolation. NewtonRaphson.m, a 30 line code function that solves chemical equilibrium problems of any degree of complexity. nglm3.m, a function with 40 lines of code for the general non-linear leastsquares fitting based on the Newton-Gauss algorithm. It incorporates a very efficient handling of parameters. A package of compact PCR and PLS programs that includes cross validation. While many aspects of data analysis are introduced, starting from very basic facts, the book is not primarily written for the beginner. Its main audience is expected to come from post-graduate students, research and industrial chemists with sufficient interest in data analysis to warrant the development of their own software rather than relying on other people's packages that all too often are rather black boxes. Statistics plays a crucial role in any data analysis, and accordingly, the statistical aspects are mentioned and appropriate equations/code are supplied. E.g. examples are given for the least-squares analysis of data with white noise as well as f2-analyses for data with non-uniformly distributed noise. However, the statistical background for the appropriate choice of the two methods and more importantly, the effects of wrong assumptions about the noise structure are not included. Many of our students, colleagues and friends deserve to be acknowledged. Most important are the students. They have repeatedly forced us to think and re-think the concepts of data analysis. The principle is straightforward: it is not possible to explain anything properly without having understood it in depth. Most important were those students who have been involved in chemometrics projects over the years: Andrew Whitson, Arnaldo Cumbana, Caroline Mason, Eric Wilkes, Graeme Puxty, Jeff Harmer, Kirsten Molloy, Maryam Vosough, Monica Rossignoli, Nichola McCann, Pascal Bugnon, Peter Lye, Porn Jandanklang, Raylene Dyson, Rod Williams, and Sarah Norman. Important are also all the colleagues who helped educate us and have been part in some or many aspects of data analysis, they include: Alan Williams, André Merbach, Andreas Zuberbühler, Anna de Juan, Arne Zilian, Bernhard Jung, Bill Tolman, Charlie Meyer, Christoph Borchers, Dom Swinkels, Ed Constable, Elmars Krausz, Geoff Lawrance, Hans Brintzinger, Harald Gampp, Helmut Mäcke, Ira Brinn, Jean Clerc, Jim Ferguson, Ken Karlin, Konrad Hungerbühler, Liselotte Siegfried, Manuel Martínez, Martin

Preface

xi

Schumacher, Paul Gemperline, Peter King, Peter Comba, Robert Binstead, Romá Tauler, Sigrid Mönkeberg, Silvio Fallab, Susan Kaderli, Thomas Kaden, Tom Callcott. Special thanks to our colleagues at the Department of Chemistry, the University of Newcastle, for taking over MM's teaching and more importantly, his share of administration while on sabbatical leave, sweating over this book. Thanks also to Jenny Helman and Gudrun Ludescher for the incredible effort of proof reading a text without understanding the content. They deserve a warm thank-you from us and also from every reader.

Marcel Maeder Yorck-Michael (Bobby) Neuhold Newcastle, Australia September 2006

"Alles für die Wissenschaft", Gerda Maeder

This page intentionally left blank

Symbols y, Y vector (mオ1, nsオ1) or matrix (mオn, nsオnl) of data (e.g. single or multi wavelength absorbance data) C matrix (mオnc, nsオnc) of component concentrations a, A vector (npオ1, ncオ1) or matrix (ncオnl) of linear parameters (e.g. molar absorptivities of pure species spectra) r, R vector (mオ1, nsオ1) or matrix (mオn or nsオnl) of residuals F

design matrix for linear regression (mオnp or nsオnp)

J

Jacobian matrix of derivatives with respect to parameters

U

column matrix of all eigenvectors of YYt

S diagonal matrix of all corresponding singular values in decreasing order of their magnitude V

row matrix of all eigenvectors of YtY

¯ U

column matrix of significant eigenvectors (nsオne)

¯ S

diagonal matrix of significant singular values (neオne)

¯ V

row matrix of significant eigenvectors (neオnl)

¯ Y

factor analytically reproduced data matrix (nsオnl)

Yred

data matrix of absorbances in reduced eigenspace (nsオne)

Ared

matrix of molar absorptivities in reduced eigenspace (ncオne)

Yglob

vertically concatenated data matrices Y1…Ynm

Cglob

vertically concatenated concentration matrices C1…Cnm

p vector of parameters (npオ1) T transformation matrix (ncオnc), score matrix in PLS (nsオne) P

loading matrix in PLS (neオnl)

W

matrix of loading weights in PLS (neオnl)

vprog

prognostic vector in PCR or PLS (nlオ1)

m, ns

# of spectra (e.g. # of rows in Y and C)

n, nl

# of lambdas (e.g. # of columns in Y or A)

nc

# of components or species (e.g. # of columns in C or rows in A)

xiv

ne

nm nd

Symbols

¯ and V ¯ t or diagonal # of factors (e.g. # of columns in U ¯ elements in S, # of factors used for PCR or PLS prediction) # of measurements (e.g. # of submatrices in Yglob and Cglob) polynomial degree

df degree of freedom np nu

# of parameters # of unknown spectra

mp Marquardt parameter A, B, C, … names of chemical species [A] concentration of species A [A]tot

total concentration of component A

[A]0

initial concentration of species A

K equilibrium constant exyz

formation constant of species XxYyZz

k rate constant oj

j-th wavelength or j-th eigenvalue

ssq

sum of squared residuals

f2

sum of squared weighted residuals

vr

standard deviation of the residuals

vy

standard deviation of the noise in y or Y

vp i

standard deviation of the parameter pi

1 Introduction As the title Practical Data Analysis in Chemistry indicates, there are different facets to the book: The book is about data analysis, the data are taken from chemistry, and the emphasis is on practical considerations. Data Analysis in Chemistry is an ambitious title; of course we cannot cover all aspects of data analysis in chemistry. A substantial fraction of the examples investigated in the different chapters is based on data from absorption spectroscopy. Absorption spectroscopy is a very powerful and very readily available technique; there are very few laboratories that do not have a spectrophotometer. Further, Beer-Lambert's law establishes a very neat and simple relationship between signal and concentration of species in solution. While most of the examples discussed in this book deal with spectra measured in the visible wavelength region, this is not important. BeerLambert's law is valid at any wavelength and covers UV, as well as NIR and IR spectroscopy. Identical laws govern CD spectroscopy and thus the methods can be adapted immediately. The only difference is that CD signals can be negative and thus, those methods that rely on positive molar absorptivities, need to be modified. Also, light emission spectroscopy often obeys laws that are very similar to Beer-Lambert's law. Other examples of data types used in the book include potentiometric data (pH) and data from monovariate chromatography detectors, such as flame ionisation or refractive index. The crucial feature is that there is a clearly defined relationship between signal and concentration. All the numerical methods are developed to a level that allows the analysis of complete absorption spectra, e.g. complete absorption spectra are measured as a function of time in a kinetic investigation. More traditional singlewavelength measurements are just a special case of spectra measured at one wavelength only. This multivariate ability is also a significant aspect of the book. There are few commercial and publicly available programs that include the analysis of multivariate data. The collection of examples is extensive and includes relatively simple data analysis tasks such as polynomial fits; they are used to develop the principles of data analysis. Some chemical processes will be discussed extensively; they include kinetics, equilibrium investigations and chromatography. Kinetics and equilibrium investigations are often reasonably complex processes, delivering complicated data sets and thus require fairly complex modelling and fitting algorithms. These processes serve as examples for the advanced analysis methods.

2

Chapter 1

There are many types of data in chemistry that are not specifically covered in this book. For example, we do not discuss NMR data. NMR spectra of solutions that do not include fast equilibria (fast on the NMR time scale) can be treated essentially in the same way as absorption spectra. If fast equilibria are involved, e.g. protonation equilibria, other methods need to be applied. We do not discuss the highly specialised data analysis problems arising from single crystal X-ray diffraction measurements. Further, we do not investigate any kind of molecular modelling or molecular dynamics methods. While these methods use a lot of computing time and power, they are more concerned with data generation than with data analysis. Also, we do not cover several typical chemometrics types of analyses, such as cluster analysis, experimental design, pattern recognition, classification, neural networks, wavelet transforms, qualimetrics etc. This explains our decision not to include the word 'chemometrics' in the title. What is practical about this book? The book is not meant to be taken into the lab. It is practical in a different way: all methods and equations that are developed, are translated immediately into a short computer program that performs the particular analysis under investigation. We decided not to supply a collection of data files from real experiments that could be read by the analysis programs for further processing. Instead, we provide short files that generate 'measurements'. The main advantage of this practice is that the reader will be able to analyse and understand the structure of the data. The results of the analyses can be compared with the input, e.g. resulting rate constants in a kinetic fit can be compared with the rate constants used to generate the data. The practice also invites the reader to 'play' with the data, investigating the influence of noise level, noise structure. The reader can observe the effects of changing the parameters used to generate the data, such as rate or equilibrium constants, absorption spectra for the reacting species, general conditions such as initial concentrations, etc. The data generation or modelling functions are a powerful educational tool. The extensive collection of example programs is a unique feature of this book. They are meant to be an invitation for the reader to be used, to be incorporated into the readers' packages and also to be fiddled with and improved. We have put considerable effort into writing good code, but no doubt, there is room for improvement. Matlab is a matrix oriented language that is just about perfect for most data analysis tasks. Those readers who already know Matlab will agree with that statement. Those who have not used Matlab so far, will be amazed by the ease with which rather sophisticated programs can be developed. This strength of Matlab is a weak point in Excel. While Excel does include matrix operations, they are clumsy and probably for this reason, not well known and used. An additional shortcoming of Excel is the lack of functions for Factor Analysis or the Singular Value Decomposition. Nevertheless, Excel is very powerful and allows the analysis of fairly complex data. The book is structured in four main chapters.

Introduction

3

Chapter 2, Matrix Algebra, gives a very brief introduction into matrix algebra. Most tasks in numerical data analysis are advantageously formulated in elegant and efficient matrix notation. Matlab is a matrix based language and thus ideally suited for the development of programs dealing with numerical analyses. This point cannot be over-stressed. The few short programs presented in this chapter may also serve as a very rudimentary introduction into Matlab. Readers not familiar with Matlab but otherwise proficient in an alternative language will be surprised at the almost complete lack of for … end loops. We also introduce matrix operations in Excel, assuming that the other, more common aspects of Excel are known to the reader. While there is a reasonable collection of matrix operations available in Excel, their usage is rather cumbersome. We believe that many readers will appreciate the short introduction into this aspect of Excel. Of course parts of, or the whole chapter can be skipped by those readers who are already proficient in the basics of matrix algebra and the implementation in Matlab and Excel. Chapter 3, Physical/Chemical Models, starts with a review of Beer-Lambert's law and very importantly demonstrates its compatibility with matrix notation. After a short discussion on chromatographic concentration profiles (these are used heavily in Chapter 5, Model-Free Analyses), we start the development of a toolbox for the computational analysis of equilibrium problems. All these computations are based on the law of mass action. While a few simple equilibrium systems can be solved by analytical expressions, all other systems require modelling by iterative procedures. We explore the Newton-Raphson method and develop it into an incredibly powerful algorithm that resolves equilibrium systems of any complexity. We subsequently incorporate the algorithm into programs that model potentiometric pH-titrations and spectrophotometric titrations. At this stage, the collection of routines can serve as an educational tool; later, in the subsequent Chapter 4, Model-Based Analyses, it will be incorporated into a general non-linear least-squares fitting program for the analysis of equilibrium processes and the determination of equilibrium constants. Those equilibrium processes that can be resolved explicitly are straightforwardly modelled in Excel. While it is possible to solve equilibrium problems of essentially any complexity in Excel, it is virtually impossible to develop a reasonable spreadsheet for the modelling of a complex titration. Iterative methods are generally difficult to implement in Excel. The Newton-Raphson algorithm is further developed into a fairly generally applicable tool for the solving of sets of non-linear equations. The equivalent to the law of mass action in equilibria are the sets of differential equations in kinetics. They are defined by the chemical reaction scheme. Again, there are explicit solutions for very simple models but most other models lead to sets of differential equations that need to be integrated numerically. Matlab supplies an extensive collection of functions for

4

Chapter 1

numerical integration dealing with just about any conceivable case, in particular the so-called 'stiff problems'. It would go well beyond the limits and scope of this book to develop such algorithms. We do, however, explain the principles of numerical integration and also develop an Excel spreadsheet with a 4th order Runge-Kutta algorithm. We demonstrate the use of Matlab's numerical integration routines (ODEsolvers) and apply them to a representative collection of interesting mechanisms of increasing complexity, such as an autocatalytic reaction, predator-prey kinetics, oscillating reactions and chaotic systems. This section demonstrates the educational usefulness of data modelling. The collection of kinetic modelling programs will be adapted in the subsequent chapter for the non-linear least-squares analysis of kinetic data and the determination of rate constants. Chapter 4, Model-Based Analyses, is essentially an introduction into leastsquares fitting. It is crucial to clearly distinguish between linear and nonlinear least-squares fitting: linear problems have explicit solutions while non-linear problems need to be solved iteratively. Linear regression forms the base for just about 'everything' and thus requires particular consideration. For non-linear regression there are several iterative methods available. The simplex algorithm is a popular method. Its concept is simple but convergence and execution times are slow. In this chapter we present the Newton-Gauss algorithm enhanced by the Levenberg/Marquardt method. The method is developed using a representative collection of worked examples that illustrate the different aspects of data fitting. We end up with a collection of programs that can fit any reaction mechanism in kinetics and any system in equilibrium studies. The most advanced features include multivariate data, inclusion of known spectra, efficient handling of variable and non-variable parameters and global analysis of series of multivariate measurements. Chapter 5, Model-Free Analyses. Model-based data fitting analyses rely crucially on the choice of the correct model; model-free analyses allow insight into the data without prior chemical knowledge about the process. Model-free analysis is based on restrictions imposed on the results of the analysis. The restrictions that are demanded by the physics of the measurement rather then by the scientist. Typical restrictions of this kind are that concentrations and molar absorptivities have to be positive. Only multivariate (e.g. multi-wavelength) data are amenable to model-free analyses. While this is a restriction, it is not a serious one. The goal of the analysis is to decompose the matrix of data into a product of two physically meaningful matrices, usually into a matrix containing the concentration profiles of the components taking part in the chemical process, and a matrix that contains their absorption spectra (Beer-Lambert's law). If there are no model-based equations that quantitatively describe the data, model-free analyses are the only method of analysis. Otherwise, the results of model-

Introduction

5

free analyses can guide the researcher in the choice of the correct model for a subsequent model-based analysis. An important group of methods relies on the inherent order of the data, typically time in kinetics or chromatography. These methods are often based on Evolving Factor Analysis and its derivatives. Another well known family of model-free methods is based on the Alternating Least-Squares algorithm that solely relies on restrictions such as positive spectra and concentrations. There is a rich collection of publications describing novel methods for ModelFree Analyses. The selection presented here does not cover the complete range; it attempts to select the more useful and interesting methods. Such a selection is always influenced by personal preferences and thus can be biased. Excel does not provide functions for the factor analysis of matrices. Further, Excel does not support iterative processes. Consequently, there are no Excel examples in Chapter 5, Model-Free Analyses. There are vast numbers of free add-ins available on the internet, e.g. for the Singular Value Decomposition. Alternatively, it is possible to write Visual Basic programs for the task and link them to Excel. We strongly believe that such algorithms are much better written in Matlab and decided not to include such options in our Excel collection. This chapter ends with a short description of the important methods, Principal Component Regression (PCR) and Partial Least-Squares (PLS). Attention is drawn to the similarity of the two methods. Both methods aim at predicting properties of samples based on spectroscopic information. The required information is extracted from a calibration set of samples with known spectrum and property. In any book, there are relevant issues that are not covered. The most obvious in this book is probably a lack of in-depth statistical analysis of the results of model-based and model-free analyses. Data fitting does produce standard deviations for the fitted parameters, but translation into confidence limits is much more difficult for reasonably complex models. Also, the effects of the separation of linear and non-linear parameters are, to our knowledge, not well investigated. Very little is known about errors and confidence limits in the area of model-free analysis. We tried to keep the programs short; they perform only the essential numerical operations, followed by minimal data output. Usually there are one or two graphs of output and occasionally a few lines of numerical output. The graphs are designed to be simple but instructive. They do not make use of the richness of the Matlab graphics routines. We did not include any graphical user-interfaces (GUIs) in order to avoid difficulties with different versions of Matlab. Matlab versions starting from 5.3 should be compatible with all programs in this book.

6

Chapter 1

Iterative refining processes invariably depend on the quality of initial guesses. If these are too far from the optimum, the process can diverge and collapse, sometimes seriously. The code provided for all iterative processes is minimal and works for most reasonably well behaved problems; however, the routines are not fool-proof. In the case of divergence and collapse we recommend the user investigates the appropriateness of the initial guesses supplied to the function. A few words regarding programming style are appropriate. Any computer program is a compromise between readability, length of code and speed of execution. Matlab, in particular, offers a substantial range of powerful commands that allows the composition of extremely compact code. The reader will find a few instances where commands perform very complex operations at the expense of being almost incomprehensible. In many instances explanations will be given in the form of comments or in adjoining text but there may be several occasions when the novice will struggle to understand the line of code. The process of preparing programs for a digital computer is especially attractive, not only because it can be economically and scientifically rewarding, but also because it can be an aesthetic experience much like composing poetry or music. Donald Knuth - The Art of Computer Programming.

2 Matrix Algebra In this chapter, we present the basic matrix mathematics that is required for understanding the methods introduced later in the book. In line with the philosophy that all concepts are immediately implemented in Matlab and/or Excel, this will be done here as well. This way, Chapter 2 not only revises the basic mathematics, it also serves as a very short introduction to the Matlab and Excel languages. It is not meant to be a manual on Matlab or Excel; the reader will need to refer to more specialised texts and proper manuals. Several more advanced features of both languages are not covered at this introductory stage but will be explained as they emerge in later chapters. Generally, most chemists possess some knowledge in one or more classical programming languages such as Basic or Fortran. While this is certainly helpful, it also produces some 'bad habits'. Matlab is a matrix-based language; it incorporates an extensive function library for matrices, vectors and data arrays in general and so is particularly well designed for the numerical analysis of multivariate data. Even though classical loop-based programming is possible for matrix operations, in Matlab there is usually a much shorter and faster way of performing the same task. Matlab programs are very readable since matrix equations are almost written as in 'real life'. Very important properties of Matlab are the direct availability of a vast number of functions that directly allow high level graphical output and the fact that Matlab works on interpreter basis, i.e. compilation is done in the background and all variables and data can be accessed at the prompt. This makes Matlab one of the most used development tools in engineering and the current standard in chemometrics. Most programs provided in this book have been developed and tested in the standard version of Matlab 6.1 and do not require any additional toolboxes (e.g. the optimisation toolbox). Our philosophy is to keep the algorithms as simple as feasible. Additionally, we avoid Matlab's capabilities in programming a graphical user interface (GUI). That way, backwards compatibility of our programs to Matlab 5.3 as well as upwards to Matlab 7.x is very likely, yet not always guaranteed. As an integral component of Microsoft Office, the spreadsheet program Excel is installed on many personal computers. Thus, a widespread basic expertise can be assumed. Although initially designed for business calculations and graphics, Excel is also extremely useful for scientific purposes. Its matrix capabilities, as well as the optimisation add-in 'solver', are not widely known but can often be applied in order to quickly resolve quite complex multivariate problems. We have used Excel 2002 but any other version will do equally well. As mentioned before, this chapter has two goals, (a) to refresh some basic matrix mathematics and (b) to familiarise the reader with the essentials of both Matlab and Excel, particularly with respect to multivariate data

8

Chapter 2

analysis of chemical problems. In order to minimise abstractness we provide many examples. It is helpful to distinguish matrices, vectors, scalars and indices by typographic conventions. Matrices are denoted in boldface capital characters (A), vectors in boldface lowercase (a) and scalars in lowercase italic characters (s). For indices, lower case characters are used (i). The symbol 't' indicates matrix and vector transposition (At, at). All chemical applications discussed later in this book will deal exclusively with real numbers. Thus, we introduce matrix algebra for real numbers only and do not include matrices formed by complex numbers.

2.1 Matrices, Vectors, Scalars A matrix is a rectangular array of numbers. e.g.

A

ᆰa1,1 ᆱC ᆱ ᆱai ,1 ᆱ ᆱC ᆱam,1 ᆲ@

B a1, j E B ai , j B am , j

B a1,n ᄎ@ ᄏ C ᄏ@ B a j , n ᄏ@ ᄏ E C ᄏ@ ᄏ@ B am , n ᄐ@

(2.1)

The size of a matrix is defined by its number of rows, m, and number of columns, n. We refer to the dimensions by mオn. In Matlab, the appropriate notation is [m,n]=size(A). For any matrix A, ai,j is the element in row i (i=1…m) and column j (j=1…n). Vectors and scalars can be seen as special cases of matrices where one or both dimensions have collapsed to 1. Thus, a row vector represents a 1オn matrix, a column vector an mオ1 matrix and a scalar a 1オ1 matrix. Matlab is based on the philosophy that Everything is a Matrix. In order to visualise matrices, column and row vectors, it is convenient to use rectangles, vertical and horizontal lines, as outlined in Figure 2-1.

Figure 2-1. A Matrix, a column vector and a row vector Sometimes it is helpful to specifically distinguish between row and column vectors. In such instances, we borrow Matlab's colon (:) notation. A vector x

Matrix Algebra

9

is represented by x1,: if it is a row vector (the row dimension is 1) and by x:,1 if it is a column vector (the column dimension is 1). Furthermore, every row of a matrix A can be seen as a row vector or sub matrix of A with the dimensions 1オn, while every column of A represents a column vector or sub matrix of the dimensions mオ1. Thus, the second row of matrix A can be referred to as the row vector a2,:, the third column of A as the column vector a:,3, etc…. With this notation it is generally possible to denote any sub matrix of A. For example, A2:4,3:6 is a matrix of dimensions 3オ4 comprised of the elements of A that are within the rectangle defined by rows 2 to 4 and columns 3 to 6. Let's see how this is done in Matlab: A=[1 2 3 4 5 6; ... % 1st row of A

2 3 4 5 6 7; ... % 2nd

3 4 5 6 7 8; ... % 3rd

4 5 6 7 8 9]

% 4th

a_r2c3=A(2,3)



% extract element of A in row 2 column 3

a_r2=A(2,:)





% extract 2nd row of A

a_c3=A(:,3)





% extract 3rd column of A

A_sub=A(2:4,3:6)

% extract 2nd to 4th row and 3rd to 6th column

A =

1

2

3

4

a_r2c3 =

4

a_r2 =

2

a_c3 =

3

4

5

6

A_sub =

4

5

6

2

3

4

5









3

4

5

6









4

5

6

7









5

6

7

8









6

7

8

9

3

4

5

6

7

5

6

7

6

7

8

7

8

9

A few remarks are in order: ク

all matrix entries are enclosed by two square brackets, the elements within each row are separated by blanks (or equivalently by commas), the end of a row is indicated by a semicolon



three dots at the end of a line tell Matlab there is continuing input in the next line



the percent (%) character introduces a comment

10

Chapter 2



for references to individual elements, rows, columns or sub-matrices, the corresponding row and column indices (or colon operators) are separated by a comma and put between parentheses



there is an equivalent command for the last line that creates the identical sub-matrix (A2:4,3:6) but refers to all rows and columns individually: A_sub=A([2 3 4],[3 4 5 6]). The sub-matrix is created according to the two vectors comprising the row and column indices. This can be handy if rows and/or columns that are not in a sequence are to be combined .

2.1.1 Elementary Matrix Operations

Transposition The transposed matrix At is defined as the interchange of rows and columns of A. This can also be seen as the reflection of all elements of A at its main diagonal (along ai,j, i=j ), according to transposition ai , j ソソソソソッa j , i

(2.2)

Thus, the row dimension of A becomes the column dimension of At, the column dimension of A will be the row dimension of At. j i

i

A

j

Att

Figure 2-2. A matrix A and its transposed At

A few lines in Matlab illustrate this on an example. A=[1 2 3; ... 4 5 6]

% definition of matrix A

At=A'

% matrix transposition

dim_A=size(A)

dim_At=size(At)

% retrieving the dimensions of A % retrieving the dimensions of At

A =

1

2

3

4

5

6

Matrix Algebra

11

At =

1

4

2

5

3

6

dim_A =

2

3

dim_At =

3

2

Note that Matlab uses the quote (') as the transposition operator. In an Excel spreadsheet, the TRANSPOSE function can be applied to an array of data. For this, we need to become familiar with the two most important rules to perform matrix operations in Excel: ク

The result has to be pre-selected, i.e. the dimensions of the result have to be known.

ク@ The SHIFT+CTRL keys must be held while pressing the ENTER key to confirm the operation. Figure 2-3 shows an example spreadsheet. Cells A3:C4 contain the elements of the 2オ3 matrix A. In order to perform the matrix transposition cells E3:F5, which will contain the result, At, have to be pre-selected. Next =TRANSPOSE(A3:C4) is typed on the Excel command line followed by the SHIFT+CTRL+ENTER key combination.

Figure 2-3. Matrix transposition in Excel Note the curly braces that have appeared in the command line and indicate a matrix operation applied to a block of cells. The curly braces must not be typed in explicitly. As for most Excel functions, there is always an alternative way via the main menu's Insert-Function feature and an upcoming graphical user interface that leads you through the process of selecting the correct function and the source cells corresponding to the matrix A. This takes a bit longer but has the advantage that you do not have to recall the exact syntax. However, the target cells, i.e. the elements corresponding to the transposed matrix At still

12

Chapter 2

have to be pre-selected. When the TRANSPOSE function is selected from the Insert-Function menu, a window similar to the one shown in Figure 2-4 appears.

Figure 2-4. Matrix transposition via Excels interactive graphical user interface With the correct source cells in the input line and pushing the OK button while holding the SHIFT+CTRL keys, the equivalent result as in Figure 2-3 is obtained. Importantly, once an array of cells has been declared a matrix by applying the SHIFT+CTRL+ENTER or the SHIFT+CTRL+OK combination, it is no longer possible to alter or delete its individual cells. Always, the whole array has to be pre-selected and thus only the complete array can be modified. As you have probably already noticed, Excel uses alphabetical letters for the column index and numbers for the row index in order to specify a cell on the spreadsheet. The column index appears before the row index, e.g. cell A3 refers to column A row 3. This is contrary to the general convention as introduced earlier and, if desired, you can change Excel's default settings (Tools-Options-General) to accommodate a row-column notation. The notation A3 will then be altered into R3C1 (row 3, column 1). Addition and Subtraction Addition or subtraction of matrices is done element-wise and thus is straightforward. Obviously, the dimensions of the matrices to be added have to match.

AイB

a1,1 イ@b1,1 B a1, j イ@b1, j ᆰ@ ᆱ@ C E C ᆱ ᆱ@ ai , j イ@bi , j ai ,1 イ@bi ,1 ᆱ@ C C ᆱ ᆱ@ am ,1 イ@bm ,1 B am , j イ@bm , j ᆲ@

B a1,n イ@b1,n ᄎ@ ᄏ@ C ᄏ@ ai ,n イ@bi ,n ᄏ@ ᄏ@ E C ᄏ@ B am ,n イ@bm ,n ᄏᄐ@

(2.3)

Matrix Algebra

13

Matrix addition and subtraction are commutative

A イ@B

イ@ B.A

(2.4)

and associative.

( A イ@B) イ@C

A イ@( B イ C )

(2.5)

In Matlab the standard mathematical operators for addition (+) and subtraction (-) can be used directly with matrices. As with transposition, Matlab automatically calls the appropriate functions to perform the operations. A=[1 2 3; ... 4 5 6]

% definition of A

B=[0.1 0.2 0.3; ... 0.4 0.5 0.6]

% definition of B

Y=A+B

% matrix addition

A =

1

2

3

4

5

6

B =

0.1000

0.4000

Y =

1.1000

4.4000



0.2000

0.3000

0.5000

0.6000

2.2000

3.3000

5.5000

6.6000

Note that Matlab directly adds/subtracts a single scalar element-wise to/from matrices. Suppose you want to subtract element b1,2 (a scalar) from all elements of matrix A. A valid Matlab command would be A=[1 2 3; 4 5 6];

B=[0.1 0.2 0.3; 0.4 0.5 0.6];

Y=A-B(1,2)

% subtracting element b1,2 from all elements of A

Y =

0.8000

1.8000

2.8000

3.8000

4.8000

5.8000

In Excel, mathematical operations of one or more cells can be dragged to other cells. Since a cell represents one element of an array or matrix, the effect will be an element-wise matrix calculation. Thus, addition and subtraction of matrices are straightforward. An example:

14

Chapter 2

Figure 2-5. Matrix addition in Excel Cells A3:C4 and E3:G4 comprise the elements of array A and B respectively. First, in cell I3, the addition is done for one pair of elements, A3+E3 (a1,1+b1,1). Then, the calculation is repeated by dragging cell I3 to the remaining cells of the rectangle I3:K4. It is worthwhile mentioning that matrix addition can alternatively be performed by pre-selection of all cells of the prospective Y, assigning the array addition according to A3:C4+B3:C4 and applying the SHIFT+CTRL+ENTER key combination. Usually, this has no particular advantage.

Figure 2-6. Matrix addition by array command Note the difference in the command line between Figure 2-5 and Figure 2-6. The capability of dragging results from one cell to others is a very useful property of Excel and becomes even more powerful in combination with the dollar operator ($) correctly applied within the cell reference. Referring to the previous Matlab example, if the scalar element b1,2 (cell F3) is to be subtracted from matrix A (A3:C4) in Excel, putting the dollar operator ($) in front of the column and row reference of the source cell containing the scalar b1,2 ($F$3), prevents "dragging-over" of the source cell F3 in both column and row direction.

Figure 2-7. Subtracting element b1,2 from all elements of A

Matrix Algebra

15

Similarly, it is possible to add/subtract one column, say b:,j (or one row bi,:) of B to/from all columns (or rows) of A. Then, the $ symbol is put before the column (row) index only. In the example below, the third column of B (b:,3), containing the elements of cells G3:G4, is subtracted from all columns of A and a matrix Y of the same dimensions as A is formed.

Figure 2-8. Subtracting the third column of B from all columns of A In the same way the '$' symbol can be used for other element-wise operations in Excel. The power and importance of the $ operator cannot be overrated and we apply it in several additional examples later. In Matlab the plus (+) and minus (-) operators cannot be directly applied to equations that involve vectors or matrices of different dimensions. In order to perform the same operation as in the former Excel example, column vector b:,3 must be replicated three times to match the dimensions of A. For this, the Matlab command repmat can be used. A=[1 2 3; 4 5 6];

B=[0.1 0.2 0.3; 0.4 0.5 0.6];

Y=A-repmat(B(:,3),1,3) % replicating the third column of B 3オ and

% subtracting the result from A

Y =

0.7000

1.7000

2.7000

3.4000

4.4000

5.4000

Matlab employs B(:,3) as the notation for the third column of B, b:,3. By using repmat(B(:,3),1,3) a matrix is created consisting of an 1-by-3 (horizontal) tiling of copies of B(:,3). Naturally, this function can also be used to create a vertical tiling of copies of row vectors, e.g. if row vector b2,: is to be added/subtracted to/from all rows of A. An appropriate function call would then be repmat(B(2,:),2,1). We refer to the Matlab manuals for further details on this function. The repmat command in the above application can be replaced by a more conventional loop: A=[1 2 3; 4 5 6];

16

Chapter 2

B=[0.1 0.2 0.3; 0.4 0.5 0.6];

for i=1:3

Y(:,i)=A(:,i)-B(:,3); end

It is a matter of opinion whether the loop or the repmat option is preferable. One is shorter and the other is easier to comprehend; one is Matlab specific and the other is more general. In this book, we tend to use the shorter, Matlab-style version, at least as long as the readability is not severely compromised. Multiplication

A matrix product Y=CA is defined in the following way

yi , j

nc

ᆭ@c

ci ,: オ a:, j

i ,k

ak , j

(2.6)

k 1

It is only defined if the number of columns of matrix C matches the number of rows of matrix A. The column dimension nc of C (i.e. the length of ci,:) has to be the same as the row dimension of A (i.e. the length of a:,j). j i

j i

A C

=

Y

Figure 2-9. Matrix multiplication Let ci,: be the i-th row vector of C and a:,j be the j-th column vector of A, then each element yi,j of Y is calculated as the scalar product ci,:オa:,j. In other words, the element in the i-th row and j-th column of Y is the sum over the element-wise products of the i-th row of matrix C and the j-th column of matrix A. Thus, if the dimensions of C are mオnc and the dimensions of A are ncオn, Y has the dimensions mオn. Figure 2-10 illustrates the multiplication Y=CA on a simple example. The factor matrices C (4オ2) and A (2オ3) can be arranged in such a way that the rows of C align with the rows of Y and the columns of A align with the columns of Y.

17

Matrix Algebra

C ᆰ1 ᆱᆱ@ 3 ᆱᆱ@ ᆱᆱ@ 5 ᆱᆱ@ 7 ᆲᆲ@

2ᄎ 4ᄏᄏᄏ@ ᄏ@ 6ᄏᄏ@ ᄏᄏ@ 8 ᄐᄐ@

ᆰ9 ᆱᆱ@ 19 19 ᆱᆱ@ ᆱᆱ@ 29 29 ᆱᆱ@ 39 39 ᆲᆲ@

12 15 ᄎᄎ@Y 26 26 33 33 ᄏᄏ@ ᄏᄏ@ 40 51 40 51 ᄏᄏ@ ᄏᄏ@ 54 54 69 69 ᄐᄐ@

1 2 2 ᆰᆰ@1 ᆱᆱ@4 5 ᆲᆲ@4 5

3 3 ᄎᄎ@A 6 6 ᄏᄐᄏ@ ᄐ@

Figure 2-10. Matrix multiplication Y = C A That way, the dimensions of Y (4オ3) become immediately obvious. One particular element, e.g. y2,3, calculates as the scalar product of the second row of C with the third column of A according to c2,:オa:,3=3オ3+4オ6=33. In Matlab the asterisk operator (-) is used for the matrix product. If the corresponding dimensions match all individual scalar products, ci,:オa:,j, are evaluated to form Y. C=[1 2; ... 3 4; ... 5 6; ... 7 8]

A=[1 2 3; ... 4 5 6]

Y=C*A

% the matrix product

C =

1

3

5

7



2



4



6



8

A =

1

2

3

4

5

6

Y =

9

19

29

39



12

26

40

54



15

33

51

69

Matlab automatically determines the correct dimensions of Y. In Excel, the cells comprising the prospective result Y have to be pre-selected as we have already seen for matrix transposition. For this, we need to predict the dimensions of Y from the row dimension of C and column dimension of A. Also, there is no direct operator for matrix multiplication in Excel. The function MMULT in conjunction with the SHIFT+CTRL+ENTER key

18

Chapter 2

combination needs to be applied in order to perform the operation. MMULT is called with two arguments, the cell range B3:C6, containing the elements of C, and the cell range E8:G9, containing the elements of A. The correct syntax can be taken from the Excel command line in Figure 2-11.

Figure 2-11. Matrix multiplication in Excel Vectors have earlier been introduced as matrices with one dimension being reduced to one, i.e. they are comprised of one column or one row only. When the rule for matrix multiplication given by equation (2.6) is formally applied to vectors and the appropriate dimensions match, there are two immediate consequences: (1) The product of a row vector and a column vector of same length results in a scalar (the scalar product). e.g.

ᆰ4ᄎ 1 2 3 ^ ` ᆱᆱ5 ᄏᄏ ᆱᆲ6 ᄏᄐ

@ @@1 オ 4 . 2 オ 5 . 3 オ 6 @ 33 @

(2) The product of a column vector with m rows and a row vector with n columns results in a matrix with m rows and n columns. This is the so-called outer product.

e.g.

ᆰ1 ᄎ ᆱ 2 ᄏ ᆱ ᄏ ᆱ3 ᄏ ᆱ ᄏ ᆲ4ᄐ

@ @ ^@@5 6 7`@ @ @

6 7ᄎ @ ᆰ5 ᆱ10 12 14 ᄏ @ ᆱ ᄏ ᆱ1 @ 18 21ᄏ@ 5 ᆱ ᄏ @ 0 24 28 ᄐ @ ᆲ2





There are a few useful rules for dealing with matrix products. The list is not complete and for more details, we refer to a textbook on linear algebra.

Matrix Algebra

19

If a matrix is to be multiplied by a scalar x, the multiplication is performed on every element of the matrix. Obviously, this operation is commutative. xA

ᆰ x a1,1 A x a1,n ᄎ@ ᆱ@ ᄏ@ E C ᄏ Ax ᆱ C ᆱ x am,1 B x am , n ᄏ ᆲ@ ᄐ@

(2.7)

The multiplication of matrices, however, is generally not commutative; i.e. the order of the factors must not be changed. CA コ AC

(2.8)

The multiplication of a matrix with more than one scalar (e.g. x, y) is associative and commutative (x A ) y

x (A y ) xy A

(2.9)

and the product with a sum of scalars is distributive. A (x . y ) A x .@A y

(2.10)

When three or more matrices are to be multiplied, the operation is also associative

( C A ) D

C ( A D )

(2.11)

and with the sum of two matrices involved in the product, it is also distributive.

( A .@B) D

AD . BD

(2.12)

The transpose of the product of matrices is the product of the transposed individual matrices in reversed order. (C A D) t

Dt A t C t

(2.13)

As stated earlier, Matlab's philosophy is to read everything as a matrix. Consequently, the basic operators for multiplication, right division, left division, power (*, /, \, ^) automatically perform corresponding matrix operations (^ will be introduced shortly in the context of square matrices, / and \ will be discussed later, in the context of linear regression and the calculation of a pseudo inverse, see The Pseudo-Inverse, p.117). Element-wise operations can, however, be enforced. If this is desired, the dot (.) needs to be placed before each operator (.*, ./, .\, .^).

20

Chapter 2

+ A

.*

B

=

C

./ … Figure 2-12. Element-wise matrix operations

Note that the operators for addition (+) and subtraction (-) need not be preceded by a dot. These two operations are always done element-wise. Some examples: A=[1 2 3; 4 5 6];

B=[0.1 0.2 0.3; 0.4 0.5 0.6];

X=A.*B







% element-wise multiplication Y1=A./B





% element-wise right division Y2=A.\B





% element-wise left division

Z=A.^B







% element-wise raising to the power

X =

0.1000

0.4000

1.6000

2.5000

Y1 =

10

10

10

10

10

10

Y2 =

0.1000

0.1000

0.1000

0.1000

Z =

1.0000

1.1487

1.7411

2.2361



0.9000



3.6000



0.1000

0.1000

1.3904

2.9302

Element-wise right division (./) leads to the inverse result of element-wise left division (.\) and the operator '.^' raises all elements to the corresponding power. For element-wise operations, the dimensions of the matrices always have to match. In contrast to Matlab, where the defaults are the matrix operators, in Excel the default is the element-wise operation. In fact, all basic operations (e.g. +, -, *, /, ^) and functions (e.g. EXP, LN, LOG) work element-wise in Excel. All matrix functions such as TRANSPOSE, MMULT, and MINVERSE require a

Matrix Algebra

21

pre-selection of a target cell block and the SHIFT+CTRL+ENTER key combination to perform the calculation.

2.1.2 Special Matrices There are several matrices that have special properties and thus require closer attention. The following list is not complete but it is sufficient for many applications. Square Matrix

Matrices that have the same number of rows and columns are called square matrices. From what we have learnt so far there are two immediate consequences with respect to matrix multiplication. (1) The product of any matrix A with its transpose At or vice versa results in square matrices that are symmetric (see below); but recall that AAt コ@AtA. (2) Any square matrix can be multiplied with itself repeatedly and the resulting matrix is also square. This is identical to raising the power of the matrix. A few examples in Matlab: A=[1 2 3; 4 5 6];

Y1=A*A'

% multiplication if A with its transpose Y2=A'*A

% or vice versa

Z1=Y2*Y2

% multiplication of Y2 with itself Z1=Y2^2

% or, equivalently, raising the power of Y2

Y1 =

14

32

Y2 =

17

22

27

Z1 =



32

77

22

27

29

36

36

45

1502



1984



2466

1984



2621



3258

2466



3258



4050

Z1 =

1502



1984



2466

1984



2621



3258

2466



3258



4050

22

Chapter 2

Symmetric Matrix

Symmetric matrices are square matrices that are identical to their transpose. They are invariant to an inflection at their main diagonal, i.e. invariant to the interchange of row and column index. In the former Matlab example both the 2オ2 matrix Y1 and the 3オ3 matrix Y2 are symmetric. Diagonal Matrix

A square matrix comprised of zeros except for the main diagonal elements, is called a diagonal matrix. Naturally, a diagonal matrix is symmetric. The Matlab command diag(x) forms a diagonal matrix D with the entries of vector x as its diagonal elements. Reversely, the command diag(D) extracts the diagonal elements of D into a vector. x=[1 2 3]; D=diag(x)

x=diag(D)

% forming a diagonal matrix % extracting diagonal element into a vector

D =

1

0

0

0

2

0

0

0

3

x =

1

2

3

Diagonal matrices are handy when individual rows or columns of a matrix are to be multiplied by different scalar factors s1…sn. One typical example is the normalisation of B so that the square root of the sum of all squared elements in, for example, each row of B becomes one, i.e. unity length of each row vector. B=[0.1 0.2 0.3; 0.4 0.5 0.6]; s=1./sqrt(sum(B.^2,2)) % vector of row-wise normalisation coeff.

B_n=diag(s)*B

% row-wise normalisation

Note that the Matlab command sum(B.^2,2) performs a row-wise addition of all squared elements of B. For a column-wise summation sum(B.^2,1) or just sum(B.^2) could be used. s =

2.6726

1.1396

B_n =

0.2673

0.5345

0.8018

0.4558

0.5698

0.6838

Matrix Algebra

23

Another common task is to normalise in such a way that the maximum value in each e.g. column of B, becomes one. B=[0.1 0.2 0.3; 0.4 0.5 0.6]; s=1./max(B)

% vector of column-wise normalisation coeff.

B_n=B*diag(s)

% column-wise normalisation

Note that the command max(B,1) or simply max(B) finds the column-wise maxima, max(B,2) the row-wise maxima of B. s =

2.5000

2.0000

1.6667

B_n =

0.2500

0.4000

0.5000

1.0000

1.0000

1.0000

Note that the order of the multiplication with the square matrix diag(s) is different in the two examples. In the first example the rows are normalised, in the second, the columns. Identity Matrix

A diagonal matrix with only ones as diagonal elements is called identity matrix, commonly abbreviated as I. It is the neutral element with respect to matrix multiplication; i.e. left or right multiplication of a matrix A with an identity matrix I of appropriate dimensions results in A itself. In Matlab the command eye(n) can be used to build an identity matrix of dimensions nオn. A=[1 2 3; 4 5 6]; I1=eye(2)

Y1=I1*A



I2=eye(3)

Y2=A*I2















% 2オ2 identity matrix

% left multiplication

% 3オ3 identity matrix

% right multiplication

I1 =

1

0

Y1 =

1

4

I2 =

1

0

0

Y2 =

1

4



0



1

2

3

5

6

0

0

1

0

0

1

2

3

5

6

24

Chapter 2

Inverse Matrix

The inverse X-1 of a matrix X is defined in such a way that X X 01

X 01 X

I

(2.14)

The left and right product of a square matrix with its inverse results in an identity matrix. In Matlab the command inv(X) or equivalently X^(-1) is used for matrix inversion. Only square matrices can be inverted. X=[1 2; 3 4]; X_inv=inv(X) % matrix inversion

I=X*X_inv

X_inv =

-2.0000

1.5000

I =

1.0000

0.0000



1.0000

-0.5000





0

1.0000

Singular matrices cannot be inverted. They have linearly dependent rows or columns. In Matlab or any other computer language, singularity can simply be an issue due to the numerical precision. Consider the following example: X=[1 2; 1+1e-16 2]; rank_X=rank(X) % rank of X

X_inv=inv(X) % matrix inversion

rank_X =

1

Warning: Matrix is singular to working precision. X_inv =

Inf Inf

Inf Inf

Within the Matlab's numerical precision X is singular, i.e. the two rows (and columns) are identical, and this represents the simplest form of linear dependence. In this context, it is convenient to introduce the rank of a matrix as the number of linearly independent rows (and columns). If the rank of a square matrix is less than its dimensions then the matrix is call rank-deficient and singular. In the latter example, rank(X)=1, and less than the dimensions of X. Thus, matrix inversion is impossible due to singularity, while, in the former example, matrix X must have had full rank. Matlab provides the function rank in order to test for the rank of a matrix. For more information on this topic see Chapter 2.2, Solving Systems of Linear Equations, the Matlab manuals or any textbook on linear algebra. In Excel, matrix inversion can be performed similarly to matrix transposition (see earlier). Figure 2-13 gives an example. Cells D3:E4, defining the target matrix, have to be pre-selected and now the MINVERSE function is applied to the source cells A3:B4. Finally, the SHIFT+CTRL+ENTER key combination is used to confirm the matrix operation.

Matrix Algebra

25

Figure 2-13. Matrix inversion in Excel

Orthogonal and Orthonormal Matrices

Matrices with exclusively orthogonal column (or row) vectors are called orthogonal matrices. For any two columns x:,i and x:,j of a matrix X to be orthogonal the necessary condition states that their scalar product is zero. t x :,i x :, j

0

for all i コ@j

(2.15)

In our three dimensional world we can perceive three vectors to be orthogonal. However, in a higher dimensional space the set of equations defined by (2.15) must suffice. If columns (or rows) of X are normalised to the square root of the sum of their squared elements (i.e. to unity length), the matrix is called orthonormal. Recall that earlier this kind of normalisation was solved most elegantly by right (left) multiplication with a diagonal matrix comprising the appropriate normalisation coefficients. See the section introducing diagonal matrices for more details. Alternatively, Matlab's built-in function norm can be used to determine normalisation coefficients and perform the same task. An example for column-wise normalisation of a matrix X with orthogonal columns is given below. It is worthwhile to compare X with equation (2.15); the subspace command can be used to determine the angle between the vectors (in rad) and reconfirm orthogonality. X=[1 4; ... 2 3; ... 5 -2]















% orthogonal column matrix

angle=subspace(X(:,1),X(:,2));

% angle (in rad, pi/2=90°) angle=rad2deg(angle)







% angle (in grad)

Xn=[]; for i=1:2

Xn(:,i)=X(:,i)./norm(X(:,i)); % col.-wise normalisation end

Xn

26

Chapter 2

X =

1

4

2

3

5

-2

angle = 1.5708e+000

angle = 90

Xn =

1.8257e-001 7.4278e-001

3.6515e-001 5.5709e-001

9.1287e-001 -3.7139e-001

Note that this kind of normalisation, via the norm function, can only be performed column- (or row-) wise via a loop as seen in the Matlab box above. Calling norm with one matrix argument determines a different kind of normalisation coefficients. We refer to the Matlab help and function references for more detail. Orthonormal matrices have very special properties. If a matrix X is comprised of orthonormal rows then X X t

I

(2.16)

If matrix X is comprised of orthonormal columns then Xt X

I

(2.17)

with the appropriate dimensions of the identity matrices. If matrix X is square and has orthonormal rows, its columns are also orthonormal. The inverse is then equal to its transpose X -1

X t

(2.18)

and consequently X Xt

Xt X

I

(2.19)

2.2 Solving Systems of Linear Equations Matrix multiplication and inversion provide very useful means of representing and solving systems of linear equations. Consider the following matrix equation: y

cA

(2.20)

27

Matrix Algebra

where y

^@y1 y2 y3 ` , c

y

^@c1 c 2 c 3 ` , A

=

ᆰ@ a1,1 a1,2 a1,3 ᄎ@ ᄏ@ ᆱ@ a ᆱ@ 2,1 a 2,2 a 2,3 ᄏ@. ᆱa 3,1 a 3,2 a 3,3 ᄏ ᄐ@ ᆲ@

c

A

Figure 2-14. System of linear equations in their matrix form According to the rule for matrix multiplication introduced earlier, each element of y is calculated as the scalar product between c and the corresponding column of A. These linear operations are represented exactly by the following system of inhomogeneous linear equations: y1 c1a1,1 . c 2 a 2,1 . c 3 a 3,1 y2 c1a1,2 . c 2 a 2,2 . c 3 a 3,2

(2.21)

y3 c1a1,3 . c 2 a 2,3 . c 3 a3,3 Let's assume the elements c1, c2 and c3 of vector c are the unknowns. Thus, the system is comprised of three equations with three unknowns. Such systems of n equations with n unknowns have exactly one solution if none of the individual equations can be expressed by linear combinations of the remaining ones, i.e. if they are linearly independent. Then, the coefficient matrix A is of full rank and non-singular and its inverse, A-1, exists such that right multiplication of equation (2.20) with A-1 allows the determination of the unknowns. c

c

= =

y A 01

y

(2.22)

A-1

Figure 2-15. Solving Systems of linear equations A typical example arises from Beer-Lambert's law. In spectrophotometry, it describes the linear relationship between the concentration of a chemical species and the measured absorbance at a particular wavelength. The corresponding coefficients are called molar absorptivities. They are specific for each species and wavelength. We refer to Chapter 3.1, Beer-Lambert's Law, for a more detailed introduction of Beer-Lambert's law. Consider a mixture of three species of unknown concentrations c1, c2 and c3 for which the absorbances y1, y2 and y3 have been measured at three

28

Chapter 2

different wavelengths. Suppose that the molar absorptivity coefficients for the individual three species have been determined independently at all three wavelengths beforehand and are known. These absorptivities are collected in matrix A such that each individual row contains the values for one specific species at the three wavelengths. This case is exactly covered by equations (2.20) and (2.21). If the wavelengths have been chosen reasonably, A will be invertible and the individual concentrations c1, c2 and c3 can be determined from equation (2.22). A small Matlab routine could be as follows: y=[0.8 0.6 0.9];

A=[90 30 80; ... 20 70 50; ...

10 50 40]; c=y*inv(A)

% absorbances

% molar absorptivities % concentrations

c =

0.0079

0.0033

0.0026

It is important to stress that for this to work, the independently known matrix A of absorptivity coefficients needs to be square, i.e. it has previously been determined at as many wavelengths as there are chemical species. Often complete spectra are available with information at many more wavelengths. It would, of course, not be reasonable to simply ignore this additional information. However, if the number of wavelengths exceeds the number of chemical species, the corresponding system of equations will be over determined, i.e. there are more equations than unknowns. Consequently, A will no longer be a square matrix and equation (2.22) does not apply since the inverse is only defined for square matrices. In Chapter 4.2, we introduce a technique called linear regression that copes exactly with these cases in order to find the best possible solution.

3 Physical/Chemical Models Any textbook on Physical Chemistry is full of mathematical equations that quantitatively describe chemical and physical processes. Often these equations are explicit, often they are not. Some are simple equations and some are complex. Explicit equations are relatively straightforward to deal with Ɇ all that is required is to translate the equation into computer code, be it Matlab or Excel or any other language. As an example, let us consider the ideal gas law pV

nRT

(3.1)

where p is the pressure, V the volume, n the number of moles, R the gas constant and T the temperature in Kelvin. The equation can be rearranged to allow the calculation of the pressure as a function of volume and temperature of a gas sample of a given number of moles. The results can be plotted in a mesh plot of pressure vs. volume and temperature. We create a range of temperatures and volumes. The command logspace(­ 1.2,0,50) creates a vector of 50 logarithmically spaced values between 10­ 1.2 and 100=1; the command meshgrid produces the matrices V and T that contain all the pressures and temperatures required for the grid of values needed for the plot. MatlabFile 3-1. Gas_Laws.m %Gas_Laws

n=1;





% number of moles

R=8.206e-2;

% [L atm mol-1 K-1]

volume=logspace(-1.2,0,50)'; temp=200:10:300; [V,T]=meshgrid(volume,temp);

% Ideal gas pressure=n*R*T./V;

mesh(log10(volume),temp,pressure); view(20,30) xlabel('log(volume)');ylabel('temp');zlabel('pressure');

30

Chapter 3

Figure 3-1. The pressure of an ideal gas as a function of log(volume) and temperature. Ideal gases do not exist and for real gases, several approximate equations have been developed to describe the pressure as a function of volume and temperature. The first useful equation is due to van der Waals p

nRT ᄃn ᄋ 0a ᄄ ᄌ V 0@nb ᄅV ᄍ

2

@ @

(3.2)

The van der Waals coefficients a and b are determined experimentally for each gas. For CO2 they are a=3.610 atmL2mol-2 and b=0.0429 Lmol-1. The approximation is not perfect and negative volumes are computed at certain ranges of pressure and temperature. These values are replaced by zero. MatlabFile 3-2. Gas_Laws.m …continued %Gas_Laws ... continued

%Van der Waals corrections for CO2

a=3.61;



% [L^2 atm mol-2] b=.0429;



% [L mol-1]

pressure_vdW=n*R*T./(V-n*b)-a*n^2./(V.*V); pressure_vdW(pressure_vdW0 >0 Calculate Jacobian Calculate Jacobian

Calculate shift vector and add to component concentrations to component concentrations

Figure 3-11. A flow diagram for the Newton-Raphson algorithm.

53

Physical/Chemical Models MatlabFile 3-9. NewtonRaphson.m function c_spec=NewtonRaphson(Model, beta, c_tot, c,i)

ncomp=length(c_tot);

nspec=length(beta);

c_tot(c_tot==0)=1e-15;

% number of components % number of species % numerical difficulties if c_tot=0

it=0; while it C

d=dsolve('Da=-k1*a','Db=k1*a-2*k2*b^2','Dc=k2*b^2', ... 'a(0)=a_0','b(0)=0','c(0)=0'); pretty(simplify(d.c))

As it turns out, the explicit solution is very complex including several Bessel functions. We leave it to the reader to explore the output. As a matter of fact, most mechanisms do not have explicit solutions and require numerical integration. The few mechanisms discussed so far are exceptions rather than the rule. Fortunately, numerical integration is always possible and next we demonstrate how this can be achieved.

3.4.3 Complex Mechanisms that Require Numerical Integration Numerical integration of sets of differential equations is a well developed field of numerical analysis, e.g. most engineering problems involve differential equations. Here, we only give a very brief introduction to numerical integration. We start with the Euler method, proceed to the much more useful Runge-Kutta algorithm and finally demonstrate the use of the routines that are part of the Matlab package. The Euler Method The simplest method for the numerical integration of a system of differential equations is named after Euler. Not surprisingly, it can be seen as an adaptation of the truncated Taylor series expansion, equation (3.81), which is the standard tool for non-linear problems. We have already encountered it in Solving Complex Equilibria (p.48), and we employ it again for non-linear least-squares fitting in Chapter 4.3, Non-Linear Regression. Please note: the Euler method should not be used. It is very slow even if only a modest accuracy is required. But because of its simplicity it is ideally suited to demonstrate the general principles of the numerical integration of ordinary differential equations. As usual, we start with a simple example; consider the reversible reaction: k+ zzzz x 2 A yzzz z B k­

(3.79)

81

Physical/Chemical Models

While there is an analytical solution for this mechanism, the formula for the calculation of the concentration profiles for A and B is fairly complex, involving the tan and atan functions (according to Matlab’s symbolic toolbox). We use it to demonstrate the basic ideas of numerical integration. The Euler method can be represented graphically, see Figure 3-29.

[A]00

slope [ A% ]0 slope

[A]11

slope [ A% ]1 slope slope [ A% ]2 slope

[A]22 [A]33

t0 0

t1 1

t2 2

t3

Figure 3-29. Euler method for numerical integration.

At any time t, knowing the concentrations [ A ]t and [B ]t , the derivatives of the % ] and [B% ] , can be calculated by applying the concentrations of A and B, [ A t

t

rate law.

[ A% ]t = -2k+[ A ]t 2 + 2 k- [B ]t % ] = [B t

1 % [ A ]t = k+[A ]t 2 - k- [B ]t 2

(3.80)

Figure 3-29 only deals with the concentration [A], the principle for the treatment of concentration [B] is identical. Starting at time t0 the initial concentrations are [A]0 and [B]0. The derivatives [ A% ]0 and [B% ]0 are calculated according to equation (3.80). This allows the computation of new concentrations, [A]1 and [B]1 after a short time interval Gt = t1–t0. [A ]1 = [A ]0 + Ʀt [A% ]0 [B ]1 = [B ]0 + Ʀt [B% ]0

(3.81)

These new concentrations at time t1 in turn allow the determination of new derivatives and thus another set of concentrations [A]2 and [B]2, after the second time interval t2–t1. As shown in Figure 3-29, this procedure is simply repeated until the desired final reaction time is reached. The main disadvantage of the Euler method is that the calculated approximation for the concentrations are systematically wrong for each step.

82

Chapter 3

In the example, the tangent [A% ] always overestimates the change of the concentration for any time interval. The longer it is, the larger the deviation. Thus to maintain a good accuracy, step sizes have to be very small; but then computation times are very long and additionally other numerical problems start to interfere. Fortunately, there are many better methods available. Amongst them algorithms of the Runge-Kutta type are frequently used in chemical kinetics. In contrast to our preferred standard mode in this book, we do not develop a Matlab function for the task of numerical integration of the differential equations pertinent to chemical kinetics. While it would be fairly easy to develop basic functions that work reliably and efficiently with most mechanisms, it was decided not to include such functions since Matlab, in its basic edition, supplies a good suite of fully fledged ODE solvers. ODE solvers play a very important role in many applications outside chemistry and thus high level routines are readily available. An important aspect for fast computation is the automatic adjustment of the step-size, depending on the required accuracy. Also, it is important to differentiate between stiff and non-stiff problems. Proper discussion of the difference between the two is clearly outside the scope of this book, however, we indicate the stiffness of problems in a series of examples discussed later. So, instead of developing our own ODE solver in Matlab, we will learn how to use the routines supplied by Matlab. This will be done in a quite extensive series of examples. The situation is different for those readers who do not have access to Matlab and rely completely on Excel. In the following, we explain how a fourth order Runge-Kutta method can be incorporated into a spreadsheet and used to solve non-stiff ODE's. Fourth Order Runge-Kutta Method in Excel The fourth order Runge-Kutta method is the workhorse for the numerical integration of ODEs. Elaborate routines with automatic step-size control are available in Matlab. Here, we develop an Excel spreadsheet for the numerical integration of the k .@ zzzz x reaction mechanism 2A yzzz z B based on the 4th order Runge-Kutta k 0@

method, see Figure 3-30. The 4th order Runge-Kutta method requires four evaluations of concentrations and derivatives per step. This appears to be a serious disadvantage, but as it turns out, significantly larger step sizes can be taken for an acceptable accuracy and overall the computation times are very much shorter.

Physical/Chemical Models

83

ExcelSheet 3-6. Chapter2.xls-RungeKutta

=B5+E5/6*(F5+2*J5+2*N5+R5)

=B5+E5/6*(F5+2*J5+2*N5+R5)

=A6­A5

­A5 =A6

=B5+E5/2*J5

=B5+E5/2*J5

=­2*$B$1*B5^2+2*$B$2*C5

=­ 2*$B$1*B5^2+2*$B$2*C5

2*$B$1*L5^2+2*$B$2*M5

=­2*$B$1*L5^2+2*$B$2*M5

=B5+E5/2*F5

=B5+E5/2*F5

=B5+E5*N5

=B5+E5*N5

= ­2*$B$1*H5^2+2*$B$2*I5 =­2*$B$1*H5^2+2*$B$2*I5

=­2*$B$1*P5^2+2*$B$2*Q5

=­2*$B$1*P5^2+2*$B$2*Q5

Figure 3-30. Excel spreadsheet for the numerical integration of the k. zzzz x rate law for the reaction 2A yzzz z B using 4th order Runge-Kutta k 0@

equations. The 4th order Runge-Kutta method is reasonably complex. Without developing the equations, we demonstrate their application by applying them stepwise in an Excel spreadsheet that is written specifically for the reaction k. zzzz x 2A yzzz z B . k 0@

We start at time zero where the initial concentrations are [A]0=1 and [B]0=0 (cells B5 and C5). The time interval Gt=1 is calculated in cell E5. 1. Calculate the derivatives of the concentrations at time point 0: 2 [ A% ]0 = -2k+ [ A ]0 + 2k- [ B ] 0 2 % [B ] = k [A ] - k [B ] 0

+

0

-

0

(cell F5) (cell G5)

In the Excel language, for [ A% ]0 , this translates into =-2*$B$1*B5^2+2*$B$2*C5, as indicated in Figure 3-30. Note, in the figure only the cell formulas for the computations of component A are given. 2. Calculate approximate concentrations at the intermediate time point Gt/2, based on the concentrations and derivatives at time 0.

Gt % [A ]0 2 Gt % [B ]1 = [B ]0 + [B ]0 2

[A ]1 = [A ]0 +

(cell H5) (cell I 5)

Again, the Excel formula for component A is given in Figure 3-30.

84

Chapter 3

3. Calculate the derivatives at the intermediate time point: 2

[ A% ]1 = -2 k+ [ A ]1 + 2k- [ B ] 1 [B% ] = k [A - ]2 k [B ] 1

+

1

-



(cell J5) (cell K5)

1

4. Calculate another set of concentrations at the intermediate time point, now based on the concentrations at time 0 and the derivatives at the intermediate time point:

Gt % [A ]1 2 Gt [B ]2 = [B ]0 + [B% ]1 2

[A ]2 = [A ]0 +

(cell L5) (cell M5)

5. Compute a new set of derivatives at the intermediate time point, based on the concentrations just calculated:

[ A% ]2 = -2 k + [ A ]22 + 2 k- [B ] 2 [B% ] = k [A ] 2- k [B ] 2

+

2

-



(cell N5) (cell O5)

2

6. Next, the concentrations at the new time point 1, after the complete time interval, are computed, based on the concentrations at time point 0 and these new derivatives at the intermediate time point:

[A ]3 = [A ]0 + Gt [A% ]2 [B ] = [B ] + G@ t[B% ] 3

0

2



(cell P5)



(cell Q5)

7. Computation of the derivatives at time point 1: 2 [ A% ]3 = -2 k+ [ A ]3 + 2 k- [B ] 3 [B% ] = k [A ] 2- k [B ] 3

+

3

-

3



(cell R5) (cell S5)

8. Finally the new concentrations after the full time interval are computed as: Gt % % ] .@2[ A % ] . [A % ] (cell B6) [ A ]0 +2[ A [ A ]new = [ A ]0 + 1 3 2 6 Gt % % ] . 2[B% ] .@ % [B ]0 . 2[B (cell C6) [B ]new = [B ]0 + 1 2 [B ]3 6 These concentrations are put as the next elements into the columns B and C.







,@

85

Physical/Chemical Models

These final equations can be compared with the equivalent in the Euler approach, equation (3.81). In the Runge-Kutta method a weighted average of 4 different approximations for the derivatives is used instead of the derivative at the beginning of the time interval. Figure 3-31 displays the resulting concentration profiles for species A and B. As it is a reversible reaction, an equilibrium is established after a certain time.

1

concentration

0.8

0.6

0.4

0.2

0

0

2

4

6

8

10



time

k. zzzz x Figure 3-31. Concentration profiles for a reaction 2A yzzz z B (ソ@ k 0@

[A], ス [B]) as modelled in Excel using a 4th order Runge-Kutta for numerical integration. While the above spreadsheet looks moderately complex, it nevertheless allows the accurate numerical integration of a real chemical reaction in a very short time. We might get the impression that "this is it". Far from it, there are several important aspects that we can only point out very briefly here. As mentioned before, the spreadsheet has been set up specifically for the numerical integration of the differential equations for the reaction k. zzzz x 2A yzzz z B . It is not convenient to adapt this spreadsheet for the k 0@

computation of a different reaction scheme. Most equations need to be rewritten. Such manual changes are error prone, and ensuing wrong results can easily go unnoticed. It is of course possible to set up a more elaborate spreadsheet that more readily allows adaptation for different reaction mechanisms. (The web-site http://www.cse.csiro.au/poptools/index.htm offers a large downloadable selection of tools for Excel, amongst them a package for the numerical integration of ODE's)

86

Chapter 3

For fast computation the determination of the best step-size (interval) is crucial; steps that are too small result in correct concentrations at the expense of long computation times. On the other hand, steps that are too long save computation time but result in poor approximations. The best intervals lead to the fastest computation of concentration profiles within some pre-defined error limits. This of course requires knowledge about the required accuracy. The ideal step-size is not constant during the reaction and so needs to be adjusted continuously. If more complex mechanisms and thus systems of differential equations are to be integrated, adaptive step size control is absolutely essential. The Runge-Kutta algorithm cannot handle so-called stiff problems. Computation times are astronomical and thus the algorithm is useless, for that class of ordinary differential equations, specialised 'stiff solvers' have been developed. In our context, a system of ODEs sometimes becomes stiff if it comprises very fast and also very slow steps and/or very high and very low concentrations. As a typical example we model an oscillating reaction in The Belousov-Zhabotinsky (BZ) Reaction (p.95). It is well outside the scope of this chapter to expand on the intricacies of modern numerical integration routines. Matlab provides an excellent selection of routines for any situation. For further reading we refer to the relevant literature and the Matlab manuals. Thus, rather than trying to explain how they work, we demonstrate how they are used.

3.4.4 Interesting Kinetic Examples Fast and accurate ODE solvers are very complex algorithms. In particular the design of adaptive step size and the analysis of stiff problems require sophisticated algorithms. The development of such algorithms is beyond the scope of this book. Matlab supplies a good collection of routines that cater for all the needs of the kineticist dealing with any reasonably complex mechanism. In contrast to the other parts of this book, we do not develop a generally applicable algorithm. Instead, we demonstrate in a representative collection of interesting and exemplary mechanisms, how to use the Matlab solvers. In this chapter, we concentrate on the simulation of chemical kinetics, i.e. based on a given chemical mechanism and the relevant rate constants, the concentration profiles (the matrix C) of all reaction species is computed. The next chapter incorporates these functions into a general fitting routine that can be used to fit the optimal rate constants for a given mechanism to a particular measurement. We start with simple chemical examples, later we examine a few interesting and surprising non-chemical examples.

Physical/Chemical Models

87

Autocatalysis Processes are called autocatalytic if the products of a reaction accelerate their own formation. Autocatalytic reactions get faster as the reaction proceeds, sometimes dramatically, sometimes slowly and steadily. Exponential growth is a very basic non-chemical example. Of course the acceleration cannot be permanent; the reaction will slow down and eventually come to an end once the starting materials have been used up. Only economists believe in sustainable growth. An extreme example of an autocatalytic reaction is an explosion. In this case, it is not directly a chemical product that accelerates the reaction, it is the heat generated by the reaction. The more heat produced, the faster the reaction; the faster the reaction, the more heat, etc. There are many mechanisms that display autocatalytic behaviour. A minimal and very basic autocatalytic reaction scheme is presented below: k1 A ソソソ ッ B 2 A + B ソソソ ッ 2 B

k

(3.82)

Starting with component A there is a relatively slow first order reaction to form the product B. The second reaction is of the order two, it opens another path for the formation of component B. As it is a second order reaction, the higher the concentration of B, the faster is the decomposition of A to form more B. The system of differential equations for this reaction scheme is given below (3.83): [A% ] = -k1[A ] - k2 [A ][B ] 2[A ][B ] [B% ] = k1 [A ] + k

(3.83)

The organisation of the Matlab ODE solvers requires some explanation. For this example, the core is a function, ode_autocat.m, that returns the derivatives of the concentrations at any particular time or better, for any set of concentration of the reacting species. Essentially it is the Matlab code for equation (3.83). MatlabFile 3-20. ode_autocat.m function c_dot=ode_autocat(t,c,flag,k)

% A --> B

% A + B --> 2 B

c_dot(1,1)=-k(1)*c(1)-k(2)*c(1)*c(2);

c_dot(2,1)= k(1)*c(1)+k(2)*c(1)*c(2);

% A_dot

% B_tot

c_dot is a column vector of the two derivatives [A% ] and [B% ]. The vectors k and c contain the rate constants and the actual concentrations; t is the time at which the derivatives are computed; it is not used within this particular function. The flag is not used here either.

88

Chapter 3

This function is called numerous times from the Matlab ODE solver. In the example it is the ode45 which is the standard Runge-Kutta algorithm. ode45 requires as parameters the file name of the inner function, ode_autocat.m, the vector of initial concentrations, c0, the rate constants, k, and the total amount of time for which the reaction should be modelled (20 time units in the example). The solver returns the vector t at which the concentrations were calculated and the concentrations themselves, the matrix C. Note that due to the adaptive step size control, the concentrations are computed at times t which are not predefined. MatlabFile 3-21. autocat.m % autocat

% A --> B

% A + B --> 2 B

c0=[1;0];

% initial conc of A and B

k=[1e-6;2];

% rate constants k1 and k2

[t,C]=ode45('ode_autocat',20,c0,[],k); % call ode-solver

plot(t,C)

% plotting C vs t

xlabel('time');ylabel('conc.');

1.2

1

conc.

0.8

0.6

0.4

0.2

0

­0.2

0

5

10

time

15

20

Figure 2-32. Concentration profiles for the autocatalytic reaction k2 k1 A ソソソ ッ@B ; A + B ソソソ ッ 2 B .

Figure 2-32 shows the calculated corresponding concentration profiles using the rate constants k1=10-6s-1 and k2=2M-1s-1 for initial concentrations [A]0=1M and [B]0=0M. After an induction period of some 5 time units the reaction accelerates dramatically. At around 10 time units, when the

89

Physical/Chemical Models

component A is almost used up, the reaction decelerates and quickly comes to the end. We are using the solvers here in their very basic version. Many additional parameters can be controlled, such as maximal step size or required accuracy. We refer to the original documentation for more information about these topics. In the above program, autocat.m, the 20 represents the total time. The ODE solver calculates the optimal step size automatically and returns the time vector t with the concentrations C. The ODE solver can also be forced to return concentrations at specific times by passing the complete vector of times instead of only the total time. 0th Order Reaction In strict terms, 0th order reactions do not really exist. They are always macroscopically observed reactions where the rate of the reaction is independent of the concentrations of the reactants for a certain time period. Formally, the ODE for a basic 0th order reaction is defined below: [A% ] = -k [A ]0 = -k

(3.84)

A simple mechanism that mimics a 0th order reaction is the catalytic transformation of A to C. A reacts with the catalyst Cat to form an intermediate activated complex B. B in turn reacts further to form the product C releasing the catalyst that continues reacting with A. A + Cat ソソソ k1 ッ B 2 B ソソソ ッ C + Cat

k

(3.85)

The total concentration of catalyst is much smaller than the concentrations of the reactants or products. Note, that in real systems, the reactions are reversible and usually there are more intermediates, but for the present purpose, this minimal reaction mechanism is sufficient. The system of ODEs: [ A% ] 0k1[ A ][ Cat ] % ] 0k1[A ][Cat [Cat ] . k2 [B ] % [B ] ] 0@ k2 [B ] k1[A ][Cat % [C ] k [ B ] 2

MatlabFile 3-22. ode_zero_order.m function c_dot=ode_zero_order(t,c,flag,k)

% 0th order kinetics

% A + Cat --> B

% B --> C + Cat

c_dot(1,1)=-k(1)*c(1)*c(2);





% A_dot

(3.86)

90

Chapter 3

c_dot(2,1)=-k(1)*c(1)*c(2)+k(2)*c(3); % Cat_dot

c_dot(3,1)= k(1)*c(1)*c(2)-k(2)*c(3); % B_dot

c_dot(4,1)= k(2)*c(3);









% C_dot

The production of C is governed by the amount of intermediate B which is constant over an extended period of time. As long as there is an excess of A with respect to the catalyst, essentially all of the catalyst exists as complex B and thus this concentration is constant. The crucial differential equation is the last one; it is a 0th order reaction as long as [B] is constant. The kinetic profiles displayed in Figure 3-32 have been integrated numerically with Matlab's stiff solver ode15s using the rate constants k1=1000 M-1s-1, k2=100 s-1 for the initial concentrations [A]0=1 M, [Cat]0=10­ 4 M and [B] =[C] =0 M. For this model the standard Runge-Kutta routine is 0 0 far too slow and thus useless. MatlabFile 3-23. zero_order.m % zero_order

% A + Cat --> B

% B --> C + Cat

c0=[1;1e-4;0;0];

% initial conc of A, Cat, B and C k=[1000;100];

% rate constants k1 and k2

[t,C] = ode15s('ode_zero_order',200,c0,[],k);

% call ode-solver

figure(1); plot(t,C)

% plotting C vs t

xlabel('time');ylabel('conc.');

1

0.9

0.8

0.7

conc.

0.6

0.5

0.4

0.3

0.2

0.1

0

0

50

100

time

150

200

Figure 3-32. Concentration profiles for the reaction k2 A + Cat ソソソ@ k1 ッ B , B ソソソ ッ@C + Cat . The reaction is th approximately 0 order for about 100 s.

91

Physical/Chemical Models

The Steady-State Approximation Traditionally, reaction mechanisms of the kind above have been analysed based on the steady-state approximation. The differential equations for this mechanism cannot be integrated analytically. Numerical integration was not readily available and thus approximations were the only options available to the researcher. The concentrations of the catalyst and of the intermediate, activated complex B are always only very low and even more so their % ] and [B% ] . In the steady-state approach these two derivatives derivatives [Cat are set to 0. [B% ]

% ] k1[A ][Cat 0 [Cat ] 0 k2[B ]

0

(3.87)

This equation allows the computation of the concentration [B] k [ A ][Cat ] [B ] 1 k2

(3.88)

and the conservation of mass for the catalyst which either exists as Cat or as B [Cat ] [Cat ]0 0@[B ]

(3.89)

Introduction of (3.89) into (3.88) and a few rearrangements result in k [ A ][C at ]0 [B ] 1 k2 . k1[ A ]

(3.90)

For most of the time, up to 100 sec, k2 2 wolves % wolf --> dead wolf

c_dot(1,1)=k(1)*c(1)-k(2)*c(1)*c(2);

% sheep_dot c_dot(2,1)=k(2)*c(1)*c(2)-k(3)*c(2);

% wolf_dot

The kinetic population profiles displayed in Figure 3-34 have been obtained by numerical integration using Matlab's Runge-Kutta solver ode45 with the rate constants k1=2, k2=5, k3=6 for the initial populations [sheep]0=2, [wolf]0=2. For simplicity, we ignore the units. In ode_Lotka_Volterra.m the function that generates the differential equations is given. It is repeatedly called by the ODE-solver ode45. MatlabFile 3-26. Lotka_Volterra.m % Lotka_Volterra

% sheep --> 2 sheep % wolf + sheep --> 2 wolves % wolf --> dead wolf

c0=[2;2];

% initial 'conc' of sheep and wolves

k=[2;5;6];

% rate constants k1, k2 and k3

[t,C] = ode45('ode_lotka_volterra',10,c0,[],k); % call ode-solver figure(1); plot(t,C) % plotting C vs t

xlabel('time');ylabel('conc.') legend('sheep','wolves');

Surprisingly, the dynamic of such a population is completely cyclic. All properties of the cycle depend on the initial populations and the ‘rate constants’. This behaviour is best seen in a plot of the wolf vs. the sheep 'concentration'. For any set of initial 'concentrations' and ‘rate constants’, this cyclic behaviour is maintained.

94

Chapter 3

4 sheep wolves

3.5 3

conc.

2.5 2 1.5 1 0.5 0 0

2

4

6

8

10

time

Figure 3-34. Lotka-Volterra’s predator and prey ‘kinetics’. MatlabFile 3-27. Lotka_Volterra.m …continued %Lotka_Volterra ...continued  figure(2); plot(C(:,1),C(:,2)) xlabel('[sheep]');ylabel('[wolf]') 

2.5

2

[wolf]

1.5

1

0.5

0 0

1

2 [sheep]

3

4

Figure 3-35. The concentration of wolves plotted versus the concentration of sheep in the Lotka-Volterra predator-prey kinetics.

Physical/Chemical Models

95

Note the imperfect coincidence of the line. This effect is due to small numerical errors; increasing the accuracy of the solver reduces these differences. The Belousov-Zhabotinsky (BZ) Reaction Chemical mechanisms for real oscillating reactions are very complex and presently not understood in every detail. Nevertheless, there are approximate mechanisms which correctly model several crucial aspects of real oscillating reactions. In these simplified systems, often not all physical laws are strictly obeyed, e.g. the law of conservation of mass. The Belousov-Zhabotinsky (BZ) reaction involves the oxidation of an organic species such as malonic acid (MA) by an acidified aqueous bromate solution in the presence of a metal ion catalyst such as the Ce(III)/Ce(IV) couple. At excess [MA] the stoichiometry of the net reaction is catalyst 2BrO 30@. 3CH 2 (COOH )2 .@2H .@ ソソソソッ 2BrCH COOH ( @)2 .@3CO2 . 4H 2O (3.94)

A short induction period is typically followed by an oscillatory phase, visible by the alternating colour of the aqueous solution due to the different oxidation states of the metal catalyst. Addition of a coloured redox indicator, such as the Fe(II)/(III)(phen)3 couple, results in more dramatic colour changes. Typically, several hundred oscillations with a periodicity of approximately one minute, gradually die out within a couple of hours and the system slowly drifts towards its equilibrium state. In order to understand the BZ system Field, Körös and Noyes developed the so-called FKN mechanism. From this, Field and Noyes later derived the Oregonator model, an especially convenient kinetic model to match individual experimental observations and predict experimental conditions under which oscillations might arise. k1 BrO 30 . Br 0@ソソソ ッ@HBrO 2 . HOBr 2 BrO30 .@HBrO 2 ソソソ ッ@2HBrO 2 . 2M ox

k

k3 ッ@2HOBr HBrO 2 . Br 0@ソソソ

2HBrO 2 ソソソ ッ@BrO30 k4

(3.95)

. HOBr



MA . M ox ソソソ ッ 12 Br k5

Mox represents the metal ion catalyst in its oxidised form (Ce(IV)). It is important to note that this model is based on an experimentally determined empirical rate law and does clearly not comprise stoichiometrically correct elementary processes. The five reactions in the model provide the means to kinetically describe the four essential stages of the BZ reaction: ク@ formation of HBrO2

96

Chapter 3

ク@ autocatalytic formation of HBrO2 ク@ consumption of HBrO2 ク@ oxidation of malonic acid (MA) For the calculation of the kinetic profiles displayed in Figure 3-36, we used the rate constants k1=1.28M-1s-1, k2=33.6M-1s-1, k3=2.4オ106M-1s-1, k4=3オ103M-1s-1, k5=1M-1s-1 which result in approximate concentration profiles in acidic solution. The initial concentrations are [BrO3-]0=0.063M, [Ce(IV)]0=0.002M (=[Mox]0) and [MA]0=0.275M. The code is fairly complex and thus its development can be error prone. MatlabFile 3-28. ode_BZ.m function c_dot = ode_BZ(t,c,flag,k)

% BZ

%

% BrO3 + Br

--> HBrO2 + HOBr

% BrO3 + HBrO2 --> 2 HBrO2 + 2 Mox

% HBrO2 + Br --> 2 HOBr

% 2 HBrO2

--> BrO3 + HOBr

% MA + Mox

--> 0.5 Br

c_dot(1,1)=-k(1)*c(1)*c(2)-k(2)*c(1)*c(3)+k(4)*c(3).^2;

%BrO3_dot

c_dot(2,1)=-k(1)*c(1)*c(2)-k(3)*c(3)*c(2)+0.5*k(5)*c(6)*c(5); %Br_dot c_dot(3,1) =k(1)*c(1)*c(2)+k(2)*c(1)*c(3)-k(3)*c(3)*c(2) -2*k(4)*c(3).^2;

















%HBrO2_dot

c_dot(4,1)= k(1)*c(1)*c(2)+2*k(3)*c(3)*c(2)+k(4)*c(3).^2; %HOBr_dot c_dot(5,1)= 2*k(2)*c(1)*c(3)-k(5)*c(6)*c(5);







%Mox_dot

c_dot(6,1)=-k(5)*c(6)*c(5);



















%MA_dot

MatlabFile 3-29. BZ.m % BZ

% BrO3 + Br

--> HBrO2 + HOBr

% BrO3 + HBrO2 --> 2 HBrO2 + 2Mox

% HBrO2 + Br --> 2 HOBr

% 2 HBrO2

--> BrO3 + HOBr

% MA + Mox

--> 0.5 Br

BrO3_0=0.063; Mox_0=0.002; MA_0=0.275; k=[1.28;33.6;2.4e6;3e3;1]; options=odeset('RelTol',1e-6,'AbsTol',1e-10); [t,C] = ode15s('ode_BZ',1000,[BrO3_0 0 0 0 Mox_0 MA_0],options,k); plot(t,log10(C(:,[1 4 6])));axis([0 1000 -8 0]);

% BrO3,HOBr,MA hold;plot(t,log10(C(:,[2 3 5])),'linewidth',2);hold; % Br,HBrO2,Mox xlabel('time');ylabel('log(conc)') legend('BrO3','HOBr','MA','Br','HBrO2','Mox')

97

Physical/Chemical Models

0

BrO3

HOBr

MA

Br

HBrO2

Mox

­1

log(conc)

­2

­3

­4

­5

­6

­7

­8

0

200

400

600

800

1000

time

Figure 3-36. The BZ reaction as represented by the Oregonator model. The species Br 0 , HBrO2 and M ox display regular oscillations while the species BrO30 , HOBr and MA change their concentrations slowly and more steadily. One important note for this system: we had to increase the default accuracy of the integration (RelTol and AbsTol) and also use the stiff solver ode15s. We leave it to the reader to experience the Runge-Kutta solver ode45 or the default accuracy. Chaos, the Lorenz Attractor Our chemical experiences suggest that differential equations seem to be something stable, and by that we mean that, if there is a small change in one of the conditions, either initial concentrations or rate constants, we expect small changes in the outcomes as well. The classical example for a stable system is our solar system of planets orbiting the sun. Their trajectories are defined by their masses and initial location and velocity, all of which are the initial parameters of a relatively simple system of differential equations. As we all know, the system is very stable and we can predict the trajectories with an incredible precision, e.g. the eclipses and even the returns of comets. For a long time, humanity believed that the whole universe behaves in a similarly predictable way, of course much more complex but still essentially predictable. Descartes was the first to formally propose such a point of view.

98

Chapter 3

In the 1960's the meteorologist Edward Lorenz worked on systems of differential equations describing weather patters, and found something utterly different. The smallest modification in the initial conditions can have a dramatic effect, resulting in a completely different outcome after a certain time. Such behaviour is called chaotic. The sets of differential equations initially were rather complex but later he developed a simpler set which shows the same effect. A% B%

k 1(B - A)

C%

AB - k3C

k2 A - B - AC



(3.96)



MatlabFile 3-30. ode_Lorenz.m function c_dot=ode_Lorenz(t,c,flag,k)

% A_dot = k1(B-A) % B_dot = k2A-B-AC

% C_dot = AB-k3C

c_dot(1,1)= k(1)*(c(2)-c(1));





% A_dot

c_dot(2,1)= k(2)*c(1)-c(2)-c(1)*c(3);

% B_tot

c_dot(3,1)= c(1)*c(2)-k(3)*c(3);



% C_tot

Naturally, A, B and C as well as the constants ki have a completely different meaning than the ones we are used to from chemical kinetics; they are not species with a certain concentration. Chaotic behaviour is restricted to certain ranges of initial values and parameters. It is up to the reader to play with these options. The short program Lorenz.m calculates the 'concentrations' for A, B and C for the initial conditions. c0=[1;1;20]. Figure 3-37 displays the trajectories in a fashion that is not common in chemical kinetics. It is a plot of the time evolution of the values of A vs. B vs. C (see also Figure 3-35). Most readers will recognise the characteristic butterfly shape of the trajectory. The important aspect is that, in contrast to Figure 3-35, the trajectory is different each time. This time, it is not the effect of numerical errors but an essential aspect of the outcome. Even if the starting values for A, B and C are away from the ‘butterfly’, the trajectory moves quickly into it; it is attracted by it and thus the name, Lorenz attractor. MatlabFile 3-31. Lorenz.m % Lorenz

c0=[1;1;20];

% initial conc of A, B, C

k=[10;30;3];

% parameters k1, k2, k3

[t,C]=ode45('ode_Lorenz',30,c0,[],k); % ode-solver

figure(1); plot3(C(:,1),C(:,2),C(:,3)); grid; xlabel('A');ylabel('B');zlabel('C');

99

Physical/Chemical Models

50 40

C

30 20 10 0 40 20

20 10

0

0

-20 B

-10 -40

-20

A

Figure 3-37. The trajectory for the Lorenz attractor. And now the chaotic aspect. Lets start with very similar initial conditions of c0=[1;1;20.00001], and store the result in the matrix C1. A plot of A only for the two calculations as a function of time is most revealing. MatlabFile 3-32. Lorenz.m …continued % Lorenz ...continued  c0=[1,1,20.00001]; [t1,C1]=ode45('ode_lorenz',30,c0,[],k);  % ode­solver  figure(2); plot(t,C(:,1),t1,C1(:,1)); xlabel('time');ylabel('A'); axis([10 20 ­20 20]) 

20 15 10

A

5 0 -5 -10 -15 -20 10

12

14

16

18

20

time

Figure 3-38. Two trajectories with very slightly different initial conditions. They are indistinguishable for a relatively long time and then suddenly move apart.

100

Chapter 3

For the first 14 time units the two traces are virtually indistinguishable and then, rather suddenly, they move apart. Each trajectory still stays within the original 'butterfly' but follows a completely different path.

4 Model-Based Analyses Very rarely are measurements themselves of much use or of great interest. The statement "the absorption of the solution increased from 0.6 to 0.9 in ten minutes", is of much less use than the statement, "the reaction has a half-life of 900 sec". The goal of model-based analysis methods presented in this chapter is to facilitate the above 'translation' from original data to useful chemical information. The result of a model-based analysis is a set of values for the parameters that quantitatively describe the measurement, ideally within the limits of experimental noise. The most important prerequisite is the model, the physical-chemical, or other, description of the process under investigation. An example helps clarify the statement. The measurement is a series of absorption spectra of a reaction solution; the spectra are recorded as a function of time. The model is a second order reaction A+B→C. The parameter of interest is the rate constant of the reaction. The purpose of this chapter is to develop a collection of methods that allow the determination of the 'best' set of parameters for a particular given model and one or a collection of measurements. In other words we fit the parameter(s) to the measurement(s). It cannot be over-stressed that the task of finding the 'best' model for the measurement is a much more difficult undertaking. A crucial difference between finding the optimal parameters for a given model and finding the optimal model, lies in the fact that the parameters of a model form a continuous space, while models are discrete entities. Model-based parameter fitting relies on the continuous relationship between the quality of the fit and the parameters. There are no equivalent continuous transitions from one trial model to the next and thus, all the powerful fitting algorithms are useless. A lot of chemical intuition, experience, knowledge etc. is involved in the process of establishing the correct model. It is not the goal of this chapter to offer much help on this subject. The usual procedure is to chose a selection of reasonable models and fit them all, and subsequently make a decision on the 'best' or 'correct' one by analysing the individual results of these analyses. Some data fitting algorithms provide statistical information that allows an estimation of the quality of the fit and thus about the suitability of the model. The tools we created in Chapter 3, Physical/Chemical Models, form the core of the fitting algorithms of this chapter. The model defines a mathematical function, either explicitly (e.g. first order kinetics) or implicitly (e.g. complex equilibria), which in turn is quantitatively described by one or several parameters. In many instances the function is based on such a physical model, e.g. the law of mass action. In other instances an empirical function is chosen because it is convenient (e.g. polynomials of any degree) or because it is a reasonable approximation (e.g. Gaussian functions and their linear combinations are used to represent spectral peaks).

102

Chapter 4

A crucial point, not mentioned so far, is the question about the meaning of the expression 'best' parameters. Intuitively it seems to be clear; they are the parameters for which the calculated data match the measured data as closely as possible. Almost invariably the sum of the squares of the differences between the measured data and the calculated model function is minimised and is the measure for the quality of the fit.

4.1 Background to Least-Squares Methods There are several reasons why the sum of squares, i.e. the sum of squared differences between the measured and modelled data, is used to define the quality of a fit and thus is minimised as a function of the parameters. It is instructive to consider alternatives to the sum of squares. (a) Minimal sum of differences - is not an option, as positive and negative differences cancel each other out. Huge deviations in both directions can result in zero sums. (b) This suggests minimising the sum of the absolute values of differences. (c) Another possibility is to take the sum of the differences to the power of 4, 6, …. The higher the power the more weight is applied to the relatively large differences. (d) The ultimate is to minimize the maximal difference, which is identical to taking a very high power of the differences prior to summation. It is beyond the scope of this book to discuss the statistical theories behind the very common least-squares fitting. We refer the reader to the list of reference books in Further Reading for more details. Here we give a glimpse into the reasoning. Given a set of measurements, it is obvious that parameters producing function values that are very different from the data, are less likely correct than parameters producing function values very similar to the measurement. The statistical argument goes the other way. Given a set of parameters, what is the probability that the particular measurement could have occurred? Assuming normally distributed errors in the measurements, one can start the mathematical formalism and come to the expected end: those parameters that result in the minimal sum of squares are the ones most likely to produce the actual measurement. Leastsquares fitting delivers a maximum likelihood estimation of the parameters. We need to stress again: this is the case only if the measurement errors are independent, normally distributed and of constant standard deviation. Often these rather stringent statistical requirements are not met. However, ignoring this fact and applying the method of least-squares fitting anyway, usually does not result in a disaster. All alternatives to the least-squares measure are computationally more demanding. Note that non-uniformly distributed noise is not a problem and can easily be incorporated into the fit. Weighting according to known standard deviations of the noise results in χ2 fitting, see Non-White Noise, χ2-Fitting (p.189).

103

Model-Based Analyses

4.1.1 The Residuals and the Sum of Squares The measured data are approximated by an appropriate function. For each measured data pair (xi, yi) there is a calculated value ycalci for that particular xi. The value ycalci is computed as a function of the parameters and xi. The difference between the measurement yi and its calculated value ycalci is defined as the residual ri. This is represented in Figure 4-1.

+ + yi

ri

ycalci

+

+

+

+

+ xi

Figure 4-1. The residual ri is the difference between the measured yi and the calculated ycalci ri = yi − ycalci = yi − f (xi , parameters)

(4.1)

The residuals are a function of the parameters. Note that they are also a function of the model and the data, but we take these as given and ignore this for the time being. The sum of squares, ssq, is the sum of all the squares of the individual residuals and thus is also a function of the parameters: ssq = ∑ ri2 = ∑ (yi − ycalci )2 = f ( parametrs ) i

(4.2)

i

It is instructive to represent the situation for two typical examples. Linear Example: Straight Line For a straight line, the function for ycalc describing a vector of measurements is: ycalci = intercept + slope × xi

(4.3)

Thus ssq is a function of the parameters intercept and slope and therefore can be displayed in a 3-dimensional plot, see Figure 4-3.

104

Chapter 4

First some noisy data are generated, they are scattered around a straight line: MatlabFile 4-1. Data_mxb.m function [x,y]=Data_mxb  x=(1:10)'; y=20+6*x; randn('seed',2);              % initialise random number generator y=y+5*randn(size(y));         % adding normally distributed noise  MatlabFile 4-2. Main_mxb.m % Main_mxb  [x,y]=Data_mxb;  plot(x,y,'+',[1 10],[26 80]) axis([0 11 0 100]) xlabel('x');ylabel('y'); 

100 90 80 70

y

60 50 40 30 20 10 0 0

2

4

6

8

10

x

Figure 4-2. Noisy data scattered around the underlying straight line We can calculate ssq for a range of values for slope and intercept and plot the result in a mesh-plot, see Figure 4-3. MatlabFile 4-3. Main_mxb.m …continued % Main_mxb ...continued  intercepts=­20:5:60; slopes=0:12; for i=1:length(intercepts); for j=1:length(slopes); 

105

Model-Based Analyses

SSQ(i,j)=sum((y­(intercepts(i)+slopes(j).*x)).^2); end  end  mesh(slopes,intercepts,SSQ+5e4); colormap([0 0 0]); hold on; contour(slopes,intercepts,SSQ,50); xlabel('slopes'); ylabel('intercepts'); zlabel('ssq+5e4'); hold off; 

x 10

4

ssq+5e4

15

10

5

0 60 40

15 10

20 0 intercepts

5 -20

0

slopes

Figure 4-3. ssq vs. slope and intercept. The minimum of ssq is near the true values slope=6 and intercept=20 that were used to generate the data (see Data_mxb.m). ssq is continuously increasing for parameters moving away from their optimal values. Analysing that behaviour more closely, we can observe that the valley is parabolic in all directions. In other words, any vertical plane cutting through the surface results in a parabola. In particular, this is also the case for vertical planes parallel to the axes, i.e. ssq versus only one parameter is also a parabola. This is a property of so-called linear parameters. Non-Linear Example: Exponential Decay In order to further explore the properties of the landscape formed by the sum of squares as a function of parameters, we concentrate on a slightly more

106

Chapter 4

complex function. As an example, we use the exponential decay of the intensity of the radiation of a sample of a radioisotope. I = I 0 e −k t

(4.4)

The two parameters defining this function are the rate constant k and the initial intensity I0. First we create and plot a noisy measurement: MatlabFile 4-4. Data_Decay.m function [t,y]=Data_Decay  t=[0:50]'; k=0.05; I_0=100; randn('seed',0); y=I_0*exp(­k*t); y=y+10*randn(size(y));  MatlabFile 4-5. Main_Decay_2d.m % Main_Decay_2d  [t,y]=Data_Decay; plot(t,y,'x'); xlabel('time'); ylabel('intensity'); 

120 100

intensity

80 60 40 20 0 -20 0

10

20

30

40

50

time

Figure 4-4. Exponential decay of radiation intensity. And now we repeat what we have done earlier with the straight line fit, i.e. calculating and plotting the sum of squares, ssq, as a function of a range of parameters.

Model-Based Analyses

107

MatlabFile 4-6. Main_Decay_ssq.m % Main_Decay_ssq  [t,y]=Data_Decay;  I_0=0:10:200; k=0:.01:.2; for i=1:length(I_0); for j=1:length(k); SSQ(i,j)=sum((y­(I_0(i)*exp(­k(j).*t))).^2); end  end  mesh(k,I_0,SSQ+1e6); colormap([0 0 0]); hold on; contour(k,I_0,SSQ,100); xlabel('k'); ylabel('I_0'); zlabel('ssq+1e6'); hold off; view(50,20); 

Figure 4-5. ssq vs. initial intensity I0 and k. Comparing Figure 4-5 with the corresponding plot from the straight line fit in Figure 4-3, an important difference is that the landscape is no longer parabolic. There is a flat region and a very steep increase at the back corner. Nevertheless, the contour lines clearly indicate that there is a minimum near the correct position.

108

Chapter 4

More careful examination of this shape reveals two important facts. (a) Plots of ssq as a function of k at fixed I0 are not parabolas, while plots of ssq vs. I0 at fixed k are parabolas. This indicates that I0 is a linear parameter and k is not. (b) Close to the minimum, the landscape becomes almost parabolic, see Figure 4-6. We will see later in Chapter 4.3, Non-Linear Regression, that the fitting of non-linear parameters involves linearisation. The almost parabolic landscape close to the minimum indicates that the linearisation is a good approximation.

Figure 4-6. Close-up of ssq vs. I0 and k near the minimum. The surface is approximately parabolical. The exact location of the minimum cannot be computed explicitly for non­ linear parameters. Starting from a set of initial guesses for the parameters, the location of the minimum has to be approached iteratively. Some methods are robust, i.e. they converge reliably, even if started far from the minimum. These methods, however, tend to be slow in localising the exact position of the minimum (e.g. in Chapter 4.4.2, The Simplex Algorithm). Several alternative algorithms require the computation of the first and sometimes the second derivatives, either of the sum of squares or of the residuals with respect to the parameters (e.g. The Newton-Gauss Algorithm, see Chapters 4.3.1 and 4.4.1). If the initial guesses for the parameters are poor, convergence is often not reliable; generally, these methods are not as robust as the simplex algorithm when started far from the minimum. However, if the initial guesses for the parameters are reasonable, the progress of the iterative process is very fast.

Model-Based Analyses

109

4.2 Linear Regression This chapter on linear regression is central to the whole book, in fact linear regression is central for almost all numerical computations. This might sound surprising, as linear regression is significantly simpler than non­ linear regression or many other algorithms. The justification for the statement lies in the fact that most non-linear problems are linearised and solved iteratively, but each iterative linearisation step is solved by a linear regression calculation. Generally, the linearisation is based on a Taylor series expansion, truncated after the second term. We have already encountered the Taylor expansion in the Chapter The Newton-Raphson Algorithm (p.48). We meet it again in Chapter 4.3.1, The Newton-GaussLevenberg/Marquardt Algorithm. We can conclude that linear regression calculations are very, very common. They are continuously performed deep inside the non-linear problem solving routines. As it turns out, linear regression is, with a few exceptions, the most complex computation undertaken in any program. For this reason we specifically discuss numerical problems that may occur in certain situations. Matlab recognises the importance of linear regression calculations and introduced a very elegant and useful notation: the / forward- and \ back­ slash operators, see p.117-118. Note that the term 'Linear Regression' is somewhat misleading. It is not restricted to the task of just fitting a straight line to some data. While this task is an example of linear regression, the expression covers much more. However, to start with, we return to the task of the straight line fit.

4.2.1 Straight Line Fit - Classical Derivation It makes sense to start with the well known task of finding the best straight line through a set of (x,y)-data pairs. We can refer back to Figure 4-3 which displays the sum of squares, ssq, as a function of the two parameters defining a straight line, the slope and the intercept. The task is to find the position of the minimum, the values for slope and intercept that result in the least sum of squares. Earlier, we promised an explicit solution for the determination of linear parameters. We first change the original notation introduced in equation (4.3): ycalci = intercept + slope × xi = a1 + a 2 xi

(4.5)

It is more efficient to use a1 and a2 as parameters rather than intercept and slope. More importantly, in 4.2.3, Generalised Matrix Notation we will be able to extend the vector containing the a-values to any higher dimension.

110

Chapter 4

The sum of squares, ssq, can be written as

ssq = ∑ ri2 = ∑ (yi − ycalci )2 = ∑ (yi − (a1 + a2 xi ))2 m

m

m

i =1

i =1

i =1

(4.6)

where m denotes the number of elements in the data vector y. At the minimum, the derivatives of ssq with respect to a1 and to a2 are both zero. ∂ssq ∂ssq = =0 ∂a0 ∂a1

(4.7)

It is a matter of substituting (4.6) into (4.7) and a bit of straight algebra to arrive at: ∂ssq = ∂a1

∂ ∑ (yi − (a1 + a 2xi ))2 m

i =1

∂a1

= ∑ −2(yi − a1 − a2 xi ) = −2∑ yi + 2ma1 + 2a 2 ∑ xi = 0 m

m

m

i =1

i =1

i =1

(4.8)

and

∂ssq = ∂a 2

∂ ∑ (yi − (a1 + a 2xi ))2 m

i =1

∂a 2

= ∑ −2xi (yi − a1 − a 2 xi ) = −2∑ xi yi + 2a1 ∑ xi + 2a 2 ∑ xi2 = 0 m

m

m

m

i =1

i =1

i =1

i =1

This represents a system of 2 equations with 2 unknowns, a1 and a2. After division

by

2

and

introducing

(e.g. ∑ xi yi = Σxy ), we can write:

a

short

notation

for

the

sums

m

i =1

a1m + a2Σx = Σy

a1Σx + a2Σx 2 = Σxy

(4.9)

and this in turn is written as a matrix equation (see Chapter 2.2) ⎡m ⎢ ⎣ Σx

with the solution

Σx ⎤ ⎡a1 ⎤ ⎡ Σy ⎤ ⎥⎢ ⎥ = ⎢ ⎥ Σx 2 ⎦ ⎣a2 ⎦ ⎣ Σxy ⎦

(4.10)

111

Model-Based Analyses

⎡a1 ⎤ ⎡ m ⎢ ⎥=⎢ ⎣a2 ⎦ ⎣ Σx

Σx ⎤ ⎥ Σx 2 ⎦

−1

⎡ Σy ⎤ ⎢Σxy ⎥ ⎣ ⎦

(4.11)

The inverse of the 2-by-2 matrix can be calculated as ⎡m ⎢ ⎣ Σx

Σx ⎤ ⎥ Σx 2 ⎦

−1

=

⎡ Σx 2 1 2 ⎢ m Σx − (Σx ) ⎣ −Σx 2

−Σx ⎤ ⎥ m ⎦

(4.12)

Inserting (4.12) into the matrix product of (4.11) results in a set of two explicit equations for the two parameters:

Σx 2Σy − Σx Σxy m Σx 2 − (Σx )2 −Σx Σy + m Σxy a2 = m Σx 2 − (Σx )2 a1 =

(4.13)

Rather than writing a short program in Matlab for this result, we demonstrate how to perform the task of a straight line fit in Excel. Excel actually provides several ways of performing the job of fitting the best line through a set of data pairs. The most convenient is probably the Add Trendline … tool which delivers the result in a few clicks. Right-click on one of the points of the data series and a context menu appears as shown in Figure 4-7. ExcelSheet 4-1. Chapter3.xls-trendline

Figure 4-7. Using Trendline for a straight line fit. Select Add Trendline … to get the graphical input selection menu for the trendline as shown in Figure 4-8.

112

Chapter 4

Figure 4-8. The Add Trendline menu. Keep the default Linear Regression under the type tab and on the Options tab select Display equation on chart. This results in 14

y = 1.2667x + 1.6667

12 10 8 6 4 2 0 0

5

10

Figure 4-9. Fitted trendline with equation. One difficulty with the Trendline is that the equation only appears graphically. The values for slope and intercept have to be copied manually into the spreadsheet if they are to be used in later calculations. In Chapter 4.2.6, Excel Linest, we discuss the LINEST function of Excel which is much more versatile while still covering the best line fit. LINEST delivers the results into the spreadsheet where they can be used for further calculations. Additionally, LINEST supplies a statistical analysis.

Model-Based Analyses

113

4.2.2 Matrix Notation A useful first step towards the fitting of more complex linear functions, is to translate the equations into a matrix oriented notation. Equation (4.5) is actually a system of m equations, where m is the number of (x,y)-data pairs. ycalc1 = a1 + a 2 x1

ycalc 2 = a1 + a 2 x 2

#

ycalci = a1 + a 2 xi

(4.14)

#

ycalcm = a1 + a 2 xm

This system of m equations can be written as one matrix equation.

⎡ ycalc1 ⎤ ⎡1 x ⎤ 1 ⎢ ⎥ ⎢ ⎥ ⎢ ycalc 2 ⎥ ⎢1 x 2 ⎥ ⎢ ⎥ ⎢# # ⎥ a ⎡ 1⎤ ⎢# ⎥=⎢ ⎥ ⎢ ycalci ⎥ ⎢1 xi ⎥ ⎢⎣a2 ⎦⎥ ⎢ ⎥ ⎢ ⎥ ⎢# ⎥ ⎢# # ⎥ ⎢ ⎥ ⎢1 xm ⎦⎥ ⎣ycalcm ⎦ ⎣

or

y calc = F(x ) a

(4.15)

(4.16)

ycalc is a column vector containing the m individual elements ycalci, F(x), or shorter just F, is an m by 2 matrix; the first column is formed by ones and the second column is composed of the elements xi. The vector a contains the parameters a1 and a2.

Similarly, the vector of residuals r, as introduced in equation (4.1), can be defined in a matrix equation: y = y calc + r

r = y − y calc = y − F a

(4.17)

And the sum of squares can be written as

ssq = ∑ ri2 = r t r m

i =1

(4.18)

The task is to find that vector a for which ssq is minimal. Now we are in a better position to generalise to more complex linear functions. The prototype of linear least-squares fitting is the fitting of a

114

Chapter 4

polynomial of a higher degree. Remember, a straight line is a polynomial of degree one and hence is only a special case.

4.2.3 Generalised Matrix Notation Equations (4.15) or (4.16) represent the fit of a straight line to a set of data pairs. These equations can be written in an expanded form:

y calc = F a = f:,1 a1 + f:,2 a2

(4.19)

Recall the colon (:) notation as introduced in Chapter 2.1, Matrices, Vectors, Scalars. The first column of F, f:,1, contains m ones, while the second column, f:,2, contains the values x1,…,xm. The j-th column of F is multiplied by its corresponding j-th element of the parameter vector a and the products are summed. Equation (4.19) and its predecessors describe the special case for a polynomial of degree one. It is straightforward to generalise by adding any number of terms or columns in F and elements in a.

y calc = F a = f:,1 a1 + f:,2 a2 + ... + f:, j a j + ... + f:,np anp =

∑ f:, j a j np

j =1

(4.20)

The prototype application is the fitting of the np linear parameters, a1,…,anp defining a higher order polynomial of degree np-1. The generalisation of equation (4.5) reads as:

ycalci = a1 + a2xi + a3 xi2 + ... + a j xij −1 + ... + anp xinp −1 =

∑ a j xij −1 np

j =1

(4.21)

In matrix notation, the equivalent of equation (4.15) can be written in the following way:

y calc

⎡1 x1 ⎢ ⎢1 x 2 ⎢# # =⎢ ⎢1 xi ⎢ ⎢# # ⎢ ⎣1 xm

⎡a ⎤ x12 " x1j −1 " x1np −1 ⎤ ⎢ 1 ⎥ ⎥ a2 ⎥ x 22 " x 2j −1 " x 2np −1 ⎥ ⎢ ⎢ a3 ⎥ ⎥ # % # " # ⎥⎢ ⎥ # ⎥ j −1 2 np −1 ⎥ ⎢ xi " xi " xi ⎢ ⎥ ⎥ aj ⎥ # # # % # ⎥⎢ ⎢ # ⎥ np −1 ⎥ xm2 " xmj −1 " xm ⎦ ⎢anp ⎥ ⎣ ⎦

(4.22)

where the j-th column f:,j contains the (j-1)-th power of the elements of the vector x. It is most important to realise that the columns of F can comprise any function of x, not just the different powers. Examples include sin(x), exp(­

115

Model-Based Analyses

3x), 1./x, tan(ln(x+1)), … there is no end to the possibilities. The matrix F is often called design matrix . We repeat, the task of linear regression is to determine those values of the vector a for which the product vector ycalc=Fa is as close as possible to the actual measurements y. Closeness of course is defined by the sum of the squared differences between y and ycalc. There are several ways to derive the equations for the computation of the optimal vector a. One option would be to generalise the procedure we used for the straight line fit in equations (4.5) - (4.13), which would be rather cumbersome. In the following we use a different approach.

4.2.4 The Normal Equations We already stated that this chapter on linear regression is central to the whole book and to numerical methods in general. Thus, it is worthwhile putting extra effort into trying to fully understand the process, its limits, its dangers, etc. It is possible to represent the principles geometrically. However, due to the restriction of the human mind to comprehend three dimensions only, these geometrical representations necessarily are restricted as well. We can only deal with two columns in the design matrix F. y r

f:,2 ycalc=Fa f:,1

Figure 4-10. The best residual vector r is orthogonal to the plane spanned by the vectors f:,1 and f:,2 .

The two vectors f:,1 and f:,2 span a plane that is represented by the grey rectangle. Note that the two vectors are not orthogonal; they do not form a normal system of axes. Any point on this plane can be written as a linear combination of the two base vectors f:,1 and f:,2 or as Fa. Thus, any point on that plane is defined by a pair of numbers and this pair forms the vector a. The pair could be called coordinates but this might be misleading, as coordinates usually are based on an orthogonal system of axes. In this graph, the linear regression problem can be understood as finding the point on the plane that is closest to the measurement vector y. The vector of measured data y usually does not lie in the plane spanned by the columns of F. It would lie there if there were no measurement errors or no noise. Be reminded: the column vectors in F, y and r are m-dimensional vectors, thus there is a high dimensional space in which y can be outside the plane F.

116

Chapter 4

Unfortunately, it is not possible for us to 'see' the grey plane imbedded in an m-dimensional space, 3 dimensions have to suffice. Some people have a good 3 dimensional imagination. They immediately 'see' that the closest point on the plane is just vertically underneath the tip of the vector y. Using more appropriate expressions: the minimal residual vector r, which is the shortest difference between y and Fa, is orthogonal, or normal, to the plane defined by F. Now the expression 'Normal Equations' starts to make sense. The residual vector r is normal to the grey plane and thus normal to both vectors f:,1 and f:,2. As outlined earlier, in Chapter Orthogonal and Orthonormal Matrices (p.25), for orthogonal (normal) vectors the scalar product is zero. Thus, the scalar product between each column of F and vector r is zero. The system of equations corresponding to this statement is: t f:,1 r=0

t f:,2 r=0

(4.23)

This set of equations can be further simplified and written as one matrix equation:

Ft r = 0

(4.24)

where 0 is now a column vector with two 0's. All that is needed now are a few matrix algebraic manipulations to arrive at the equation for the calculation of the best a. F t r = F t (y − F a) = 0 thus

F y = F Fa t

(4.25)

t

to result in a = (F t F )-1 F t y

(4.26)

This last equation is very crucial and we will spend considerable time investigating it further. Equations (4.11) and (4.26) have to be identical. Simple verification based on the rules for matrix multiplication:

117

Model-Based Analyses

⎡1 1 " F tF = ⎢ ⎣ x1 x 2 " ⎡1 1 " Fty = ⎢ ⎣ x1 x 2 " thus

⎡a ⎤ ⎡ m a = ⎢ 1⎥ = ⎢ ⎣a2 ⎦ ⎣Σx

⎡1 x1 ⎤ ⎢ 1 ⎤ ⎢1 x 2 ⎥⎥ ⎡m = ⎢ ⎥ xm ⎦ ⎢ # # ⎥ ⎣ Σx ⎢ ⎥ 1 x m⎦ ⎣ y ⎡ 1⎤ 1 ⎤ ⎢⎢ y2 ⎥⎥ ⎡ Σy ⎤ = ⎢ ⎥ ⎥ ⎢ ⎥ xm ⎦ # ⎣Σxy ⎦ ⎢ ⎥ ⎣ym ⎦ Σx ⎤ ⎥ Σx 2 ⎦

−1

Σx ⎤ ⎥ Σx 2 ⎦

(4.27)

⎡ Σy ⎤ ⎢Σxy ⎥ ⎦ ⎣

The important point is that equation (4.26) is very general and does not require any calculations of derivatives of ssq with respect to the parameters, etc.

The Pseudo-Inverse Equation (4.17) +r can be written in a casual way y ≈ Fa

(4.28)

where the ≈ represents the least-squares solution. Don't forget that it is not a proper equation as y=Fa. As the solution to equation (4.28) one might be tempted to write " a = F -1 y "

(4.29)

which, of course, is mathematically incorrect as F is not square and thus cannot be inverted. The matrix (FtF)-1Ft in equation (4.26) replaces "F-1" in equation (4.29). It is known as the Pseudo-Inverse of F, for which the notation F+ is often used. Thus we can write a = F+y

F + = (F t F )-1 F t

(4.30)

Matlab is, of course, aware of the fundamental importance of the pseudoinverse and created its own notation for it. In Matlab we could write a=inv(F'*F)*F'*y but it is numerically much more efficient to use the appropriate Matlab back-slash \ command as in a=F\y. It is to be read from the right to the left as 'y divided by F', implying, of course, the multiplication of the left pseudo-inverse of F with y as given in equation (4.30).

118

Chapter 4

The equation y≈Fa written more casually as y=Fa can be represented by the following general scheme

=

y =

F

a

Figure 4-11. Schematic representation of the matrix equation y=Fa. y and a are column vectors while F is a matrix of the appropriate dimensions. It is possible to transpose the equation represented in Figure 4-11 and one could write yt=atFt. Renaming yt, at and Ft to y1, a1 and F1 and using other dimensions as in Figure 4-11, the next figure represents a generalised 'transposed' situation.

=

y1

=

a1

F1

Figure 4-12. Schematic representation of the matrix equation y1=a1F1. The least-squares solution for a1 in the above equation is a1 = y1 F1t (F1 F1t )−1

or

a1 =

(4.31) y1 F1+

In Matlab we could write a=y*F'*(inv(F*F')); but again it is numerically much more efficient to use the appropriate Matlab forward slash / operator as in a=y/F, which again states 'y divided by F' but now reading from left to right and meaning the multiplication of y with the right pseudo-inverse of F, as stated in equation (4.31). Matrix multiplication is not commutative, (see Chapter 2.1.1, Elementary Matrix Operations) and thus the order of the factors is important and in a similar way the order is important with 'division'. There are the left and right pseudo-inverses and the / and \ slashes in Matlab represent them in a very elegant way. With beginners of Matlab, it is a very common mistake to use the wrong slash. In the best

119

Model-Based Analyses

case, an error occurs as the dimensions do not match. In the worst case the error goes unnoticed. It is very helpful to write down schematic matrix equations in order to verify the dimensions and correctness of the corresponding calculations. In Figure 4-13 and Figure 4-14 this was done to visualise equations (4.30) and (4.31) using the matrix/vector dimensions shown in Figure 4-11 and Figure 4-12.

=

=

a = (FtF)-1

Ft

F+

y =

y

Figure 4-13. Schematic representation of the matrix equations involving multiplication of y by the left pseudo-inverse F+=(FtF)-1Ft.

=

a1 =

=

y1

F1t

(F1F1t)-1 =

y1

F 1+

Figure 4-14. Schematic representation of the matrix equations involving multiplication of y1 by the right pseudo-inverse F1+=F1t(F1F1t)-1.

Linear Dependence, Rank of a Matrix For the computation of the pseudo-inverse, it is crucial that the vectors f:,j are not parallel, or more correctly, that they are linearly independent. Otherwise, the matrix FtF is singular and cannot be inverted. Matlab issues a warning. We can gain a certain level of understanding by adapting Figure 4-10:

120

Chapter 4 y r

f:,2

ycalc=Fa

f:,3 f:,1

Figure 4-15. Linear dependence, three vectors f:,1, f:,2, f:,3 lie in one plane Assume there are three columns in F that all lie in one plane, as indicated in Figure 4-15. In such a case, there is no unique set of coordinates defining the point on the plane that is nearest the measurement y. The coordinates can be given by any two of the three vectors. They cannot be calculated uniquely or, in other words, the three parameters of the vector a are not defined. In this case, when trying to perform a=F\y, Matlab would usually respond with an error message such as Warning: Matrix is singular to working precision. The number of linearly independent columns (or rows) in a matrix is called the rank of that matrix. The rank can be seen as the dimension of the space that is spanned by the columns (rows). In the example of Figure 4-15, there are three vectors but they only span a 2-dimensional plane and thus the rank is only 2. The rank of a matrix is a very important property and we will study rank analysis and its interpretation in chemical terms in great detail in Chapter 5, Model-Free Analyses.

Numerical Difficulties Figure 4-10 can be adapted further to represent another important aspect. In Figure 4-16 we see that the two vectors f:,1 and f:,2 are almost parallel, not exactly parallel, as then the rank would be one only. y f:,1

f:,2

r ycalc=Fa

Figure 4-16. Near linear dependence if the base vectors are almost parallel

Model-Based Analyses

121

Because the two base vectors are almost parallel, the plane they lie in is not well defined. Figure 4-16 attempts to represent the problem: the plane can be turned about the two vectors like the pages of a book about the spine. Consequently the projection of y and the residuals r are poorly defined as well. The figure also indicates that the problem is less serious if y is close to the vectors f:,1 and f:,2, than if it is almost orthogonal. As linear regression is a very fundamental operation, several methods have been developed in order to improve the numerical stability of the calculation. It is beyond the objective of this book to discuss these issues in any detail. We do feel, however, that the reader has to be aware of the potential problems and should be able to avoid them as much as possible. The Matlab computations invoked by the back-slash \ and forward-slash / operators do not perform the calculation as given in equations (4.26) and (4.31). Here a short extract from the Matlab HELP: If F is not square and is full, then Householder reflections are used to compute an orthogonal-triangular factorization. F*P = Q*R where P is a permutation, Q is orthogonal and R is upper triangular (see qr). The least squares solution X for the equation B=AX is computed with X = P*(R\(Q'*B)

Without attempting to fully understand this, the essence is important: a=F\y 

or a1=y1/F1 

are numerically much better than a=inv(F'*F)*F'*y 

or a1=y1*F1'*inv(F1*F1'). 

4.2.5 Errors in the Fitted Parameters As there is a difference between the measurements and the values of the calculated function, we can safely assume that the fitted parameters are not perfect. They are our best estimates for the true parameters and an obvious question is, how reliable are these fitted parameters? Are they tightly or they are loosely defined? As long as the assumption of random white noise applies, there are formulas that allow the computation of the standard deviation of the fitted parameters. While these answers should always be taken with a grain of salt, they do give an indication of how well defined the parameters are.

122

Chapter 4

We are not going to derive the formulas that allow the calculation of the standard deviations of the parameters. The reader is invited to refer to more specialised texts on statistics. We just give the formulas and also give ways of calculating the required information. In equation (4.32) the standard deviation of the parameter aj is given. σa j = σr d j , j

(4.32)

where σr is the standard deviation of the residuals σr =

ssq m − np

(4.33)

Here m is the number of points in y and np the number of fitted parameters. The difference m-np is the number of degrees of freedom, df. The elements dj,j in equation (4.32) are the diagonal elements of the inverse of the so-called curvature matrix, Curv, that contains the second derivatives of the sum of squares with respect to the parameters. The definition of the element Curvj,k is

Curv j ,k =

1 ∂ 2ssq 2 ∂a j ∂ak

(4.34)

This looks horrendous, but as will be shown in a moment it is not. In fact it is 'trivial' to compute. We start with the first derivative ∂ssq ∂ = ∂a j ∂a j

∑ ri2 m

i =1

np ⎛ ⎞ ⎜ yi − ∑ f i , j a j ⎟ ⎜ ⎟ j =1 ⎝ ⎠ np m ⎛ ⎞ = −2∑ ⎜ yi − ∑ f i , j a j ⎟ f i , j ⎟ ⎜ j =1 i =1 ⎝ ⎠

∂ =∑ i =1 ∂a j m

= −2 r t f:, j

and then the second derivative:

2

(4.35)

123

Model-Based Analyses

∂ 2ssq ∂ ∂ssq = ∂ak ∂a j ∂ak ∂a j ∂ ∂ak

= −2





∑ ⎜⎜ yi − ∑ f i , ja j ⎟⎟ f i , j m

i =1 ⎝

np



j =1

⎛ ⎞ ⎛ np ⎞ ⎜ yi f i , j − ⎜ ∑ f i , j a j ⎟ f i , j ⎟ ⎜ ⎟ ⎜ ⎟ ⎝ j =1 ⎠ ⎝ ⎠ m ⎞ ∂ ⎛ np = 2∑ ⎜ ∑ fi , ja j ⎟ f i , j ⎜ ⎟ i =1 ∂ak ⎝ j =1 ⎠ = −2∑

∂ ∂a i =1 k m

(4.36)

= 2∑ f i ,k f i , j m

i =1

= 2 f:,tk f:, j

The complete set of first and second derivatives can be written elegantly in matrix notation: ∂ssq = −2 r t F ∂a and

(4.37)

∂ ssq = 2F t F ∂a∂a 2

The elements dj,j, as required in equation (4.32), are the diagonal elements of the inverse of FtF. D = Curv −1 = (F t F)−1

(4.38)

It is possible to represent the situation graphically.

ssq

ssq

ak

aj

Figure 4-17. The parameter ak is well defined, while the parameter aj is much more loosely defined.

124

Chapter 4

If ssq increases sharply with small movement of the parameter away from the minimum, then the parameter is well defined. Otherwise the parameter is only loosely defined. Referring to Figure 4-17 the parameter ak is better defined than the parameter aj. The example below shows a short Matlab program that fits the function y=tan(x) with a polynomial of degree 3 defined by 4 linear parameters, i.e. the elements of a. The statistical analysis is problematic for this example as the residuals are obviously not normally distributed. Nonetheless, the high errors in the parameters and the large standard deviation of the residuals indicate a bad fit. MatlabFile 4-7. tan_poly.m % tan_poly  % fitting y=tan(x0) with a polynomial of degree 3  x=(0.1:0.1:1.5)'; y=tan(x);  nd=3;  np=nd+1; 

    % degree of polynomial      % number of parameters 

F(:,1)=ones(size(x));  for j=1:nd F(:,j+1)=F(:,j).*x; end 

    % design matrix 

a=F\y;  r=y­F*a;  df=length(x)­np;  ssq=r'*r;  sigma_r=sqrt(ssq/df);  D=inv(F'*F);  sigma_a=sigma_r*sqrt(diag(D)); 

             

  % linear parameters    % residuals    % degrees of freedom    % sum of squares    % sigma_r  % inverted curvature matrix    % sigma_parameters 

fprintf(1,'sigma_r: %g\n',sigma_r);  % print sigma_r for i=1:np      % print sigma_a  fprintf(1,'a(%i): %g +­ %g\n',i,a(i),sigma_a(i)); end  plot(x,y,'.',x,F*a); xlabel('x'); ylabel('y'); 

sigma_r: 1.21374 a(1): ­2.46752 +­ 1.6462 a(2): 20.9388 +­ 8.62303 a(3): ­37.9646 +­ 12.3167 a(4): 20.2432 +­ 5.0712 

125

Model-Based Analyses

16 14 12 10

y

8 6 4 2 0 -2 0

0.5

1

1.5

x

Figure 4-18. Fitting the function y=tan(x) with a polynomial of degree 3.

4.2.6 Excel Linest As indicated in Chapter 4.2.1, Straight Line Fit, the Excel function LINEST is a more general function than the TRENDLINE. In addition to allowing the fitting of any linear function, it also delivers a statistical analysis. In order to demonstrate its use, we fit the same polynomial to the same tan-function as we have done in the preceding section using Matlab and tan_poly.m. The columns A and B of the spreadsheet shown in Figure 4-19 are made up by the (x,y)-data-pairs; the array (F4:I18) is the equivalent of the matrix F. LINEST is an array function. The parameters have to be entered as demonstrated in Figure 4-20. The parameter Const has to be set to FALSE at this stage. We explain its meaning in a moment. Stats needs to be set to TRUE for the statistical analysis. As LINEST is an array function, the output area (F21:I25) needs to be selected (the area is shaded) and the command is executed by pressing Ctrl-Shift-Enter (all three keys together). The statistical analysis results are listed as explained by the text underneath the output. Note that, for unknown reasons, Excel reverses the order of the parameters, i.e. the parameter belonging to the first column is listed last. It is encouraging to see that the results of Matlab (see p.124) and Excel are identical. Excel returns a few additional statistical numbers, r2, F and ssreg in cells F23:F25. They are not vital for our present purpose and we refer to the Excel Help for details.

126

Chapter 4

ExcelSheet 4-2. Chapter3.xls-linest

=LINEST(B4:B18,F4:I18,FALSE,TRUE)

Figure 4-19. Excel spreadsheet demonstrating the use of the LINEST function.

Figure 4-20. The window for the LINEST parameter input.

Model-Based Analyses

127

Excel has an alternative way of doing the same. The column of 1's can be omitted, selecting only the three others and fitting with the LINEST function but having the Const option set to TRUE. Excel internally adds the columns of ones and delivers exactly the same results. That one special parameter is the y-intercept. Similar options exist in the trendline function.

4.2.7 Applications of Linear Least-Squares Fitting In the following we present several linear least-squares analyses. Linearisation of Non-Linear Problems In many instances non-linear functions can be linearised and in this way a non-linear, iterative fitting procedure can be reduced to an explicit linear fit. A typical example is the exponential decay of the intensity of the emission of a radioactive sample. We use the data already used for Figure 4-4, produced by the function Data_Decay.m. The equation describing the data is:

y = I 0 × e −kt

(4.39)

This equation can be rewritten in a logarithmic form ln(y ) = ln(I 0 ) - kt

(4.40)

Plotting ln(y) versus time t results in a straight line with slope -k and intercept ln(I0). These two parameters can be computed non-iteratively in a linear regression. The program could look like this: MatlabFile 4-8. Main_Decay.m % Main_Decay  [t,y]=Data_Decay; yb=y(y>0);  tb=t(y>0); 

% we have to get rid of negative values

F=ones(length(tb),2); F(:,2)=tb;  a=F\log(yb); I_0=exp(a(1)) k=­a(2)  plot(tb,log(yb),'o',tb,F*a); xlabel('time'); ylabel('ln(y)');  I_0 =  100.4029  k = 

128

Chapter 4

    0.0502 

5 4.5 4

ln(y)

3.5 3 2.5 2 1.5 1 0

10

20

30

40

50

time

Figure 4-21. The logarithmic transform of the exponential data set used in Figure 4-4. The fitted exponential curve appears as a straight line. Several observations have to be made: (a) negative y-values have to be deleted as their logarithms are not defined. In real emission experiments this is usually not a problem. Due to the non-uniform error structure of emission data, emission intensity readings are most likely not negative. We will further investigate this aspect in Non-White Noise, χ2-Fitting (p.189). In other applications, eg. spectrophotometric measurements, negative readings are to be expected if the absorbance reading is close enough to zero. (b) As is obvious from the figure, the distribution of the noise is not uniform and thus the later part of the measurement carries more weight than the earlier part in defining the parameters. (c) Thus, the fitted values are different from the best values resulting from a non-linear least-squares fit of the original (non­ linearised) data. In this example the difference is minimal and not relevant. (d) In the example, the reading of emission intensity is expected to reach zero after enough time. If the infinity reading is not zero, the formula cannot be applied directly. If the measurement is described by an equation of the form

y = I 0 × e −kt + const .

(4.41)

the logarithm of y minus the constant (which is the value at time infinity) has to be plotted versus time

Model-Based Analyses

129

ln(y - const ) = ln(I 0 ) - kt

(4.42)

This task can be problematic as the correct value for const is not necessarily accurately known. Subtracting a 'wrong' value obviously results in a flawed analysis, and this is not always easily detected. The calculations below demonstrate the problem. There is a constant offset added to the y-data, which is not subtracted according to equation (4.42). The plot in Figure 4-22 does not show any 'obvious' miss-fit, however, the calculated parameters are significantly wrong. MatlabFile 4-9. Data_Decay_Offset.m function [t,y]=Data_Decay_Offset  t=(0:50)'; k=0.05; I_0=100; randn('seed',0); y=50+I_0*exp(­k*t); y=y+10*randn(size(y));  MatlabFile 4-10. Main_Decay_Offset.m % Main_Decay_Offset  [t,y]=Data_Decay_Offset; yb=y(y>0); tb=t(y>0);  F=ones(length(tb),2); F(:,2)=tb;  a=F\log(yb); I_0=exp(a(1)) k=­a(2)  plot(tb,log(yb),'o',tb,F*a); xlabel('time'); ylabel('ln(y)'); 

I_0 =  137.7157  k =  0.0204 

It is worthwhile noting that if a smaller amount of noise is added to the ydata, the subtraction of the wrong constant offset manifests in a visible curvature of their logarithmic plot. This in turn could be misinterpreted as non-exponential behaviour.

130

Chapter 4

5.2 5 4.8

ln (y)

4.6 4.4 4.2 4 3.8 3.6 3.4 0

10

20

30

40

50

time

Figure 4-22. Logarithmic plot with incorrect offset showing no obvious deviation from linearity. Iterative processes are always time consuming and, if possible, should be avoided. Thus, it seems attractive to apply a linearisation procedure in order to avoid a direct iterative non-linear least-squares fit of the exponential. The linearisation approach might be perfectly acceptable in some instances (see the first example) but as a general recipe it is not recommended (e.g. refer to the second example). Another drawback is the non-uniform error distribution invoked by the linearisation of the exponential data. This should in fact be counteracted by appropriate weighting of the residuals as will be introduced in Non-White Noise, χ2-Fitting (p.189). Further, linearisations are only possible in a few selected cases and therefore are not of great general value. In summary: there are not many convincing reasons to do it. Polynomials, the Savitzky-Golay Digital Filter Polynomials do not play an important role in real chemical applications. Very few chemical data behave like polynomials. However, as a general data treatment tool, they are invaluable. Polynomials are used for empirical approximations of complex relationships, smoothing, differentiation and interpolation of data. Most of these applications have been introduced into chemistry by Savitzky and Golay and are known as Savitzky-Golay filters. Polynomial fitting is a linear, fast and explicit calculation, which, of course, explains the popularity.

131

Model-Based Analyses

Smoothing of Noisy Data Measured data are always corrupted by a certain level of noise. For graphical purposes, it is sometimes desirable to display, for example, a spectrum as a smooth line instead of a band of noisy data points. It is important to stress from the very beginning that data smoothing should only be a graphical aid. Data smoothing prior to any parametric or non-parametric analysis hardly ever improves the results. Data smoothing distorts the original values in ways that are difficult to control and the distortion of the results of the fitting is virtually impossible to correct. Nothing is won and a lot can be lost. The basic idea of the Savitzky-Golay digital filter is fairly straightforward. The y-value of a particular (x,y)-data pair is replaced by the value of a polynomial of a certain degree, which has been fitted to a number of neighbouring data points. The computation is repeated for all data. In Figure 4-23 the procedure is illustrated graphically: There are 100 data points in the graph. One particular point is marked by ×. It is the one for which we want to compute its smoothed equivalent. 41 neighbouring data points are marked in black points. They include 20 points to the right and 20 to the left plus the point × itself. A parabola has been fitted though these 41 points and its graph in the range of the fit is represented in the figure. The point 'o' on the parabola is the smoothed representation for the original '×'. 0.114 0.113 0.112 0.111

y

0.11 0.109 0.108 0.107 0.106 0.105 0.104 0.01

0.015

0.02

0.025 x

0.03

0.035

0.04

Figure 4-23. Savitzky-Golay filtering. A polynomial is fitted to a range of data points and the original point (×) is replaced by the value on the polynomial (o).

132

Chapter 4

Two parameters define the Savitzky-Golay filter: the number of points, n, to the right and left of the centre, which are used for the fit and the degree, nd, of the polynomial to be fitted. It is crucial to choose those two parameters carefully; they have to be appropriate for the curves to be smoothed. Many data points and a low degree polynomial result in excellent smoothing but narrow features are broadened. The extreme in this direction is to fit a polynomial of degree 0 to many data points; this is also called moving window averaging. The opposite choice, few data points and a high order polynomial, result in poor smoothing with much less distortion of narrow features. The following short program creates a series of three noisy Gaussians of decreasing width. The density of data is constant along the x-axis and thus there is a decreasing number of points defining each Gaussian. The function SavGol.m performs a Savitzky-Golay smoothing. The parameters are the x- and y-vectors, the number (n) of neighbouring left or right data points that are used for one polynomial fit (i.e. if n=5, 2n+1=11 data points are fitted) and the degree (nd) of the polynomial to be fitted. Figure 4-24 displays the data and the results of the smoothing. The top panel contains the original true curve as well as the noisy data. The second panel displays the result of the Savitzky-Golay filter using 11 data points (2×5+1) to fit a forth order polynomial. All features of the curve are reasonably well preserved, but the smoothing is much less efficient than in the lower part of the Figure where a parabola (polynomial of degree 2) was fitted to 21 (2×10+1) data points. However, in this very smooth curve, the features of the narrow Gaussians are completely lost. The higher the polynomial and the smaller the number of data points fitted, the better it follows narrow features, but the smoothing effect is diminished. The user has to find the appropriate compromise. MatlabFile 4-11. Main_SavGol.m % Main_SavGol  x=(1:150)'; y=gauss(x,50,25)+gauss(x,100,10)+gauss(x,125,2); randn('state',0) yn=y+0.1*randn(size(y));  y1=SavGol(x,yn,5,4); y2=SavGol(x,yn,10,2);  subplot(3,1,1);plot(x,y,x,yn,'.k'); axis([0 150 ­0.1 1.1]); ylabel('y'); subplot(3,1,2);plot(x,y,x,y1,'.k'); axis([0 150 ­0.1 1.1]); ylabel('y'); subplot(3,1,3);plot(x,y,x,y2,'.k'); axis([0 150 ­0.1 1.1]); xlabel('x'); ylabel('y'); 

133

Model-Based Analyses

y

1 0.5 0 0

50

100

150

50

100

150

100

150

y

1 0.5 0 0

y

1 0.5 0 0

50 x

Figure 4-24. In all panels the true data are represented by the line marker. The top panel displays the noisy (•) data; the middle panel shows the result of a 4th degree polynomial fitted through 11 noisy data points (•); and the bottom panel, the result of a 2nd degree smoothing through 21 noisy data points (•). It is worthwhile discussing the function SavGol.m in some detail. There are some interesting aspects that can illustrate a few issues of numerically reliable programming. It is tempting to write a routine such as SavGol_bad.m, to perform the Savitzky-Golay filtering, but we will show its numerical weakness. F is built up by the appropriate range of x-values and used to calculate the polynomial coefficients as a=F\y(i-n:i+n), see e.g. equation (4.31). MatlabFile 4-12. SavGol_bad.m function y1=SavGol_bad(x,y,n,nd)  % Savitzki­Golay % n:  number of points to the right or left % nd: degree of polynomial  y1=zeros(size(x)); for i=1+n:length(x)­n  x1=x(i­n:i+n); F(:,1)=ones(size(x1)); for j=1:nd F(:,j+1)=F(:,j).*x1; end  a=F\y(i­n:i+n); y1(i)=F(1+n,:)*a; end 

% the first and last n points are lost

134

Chapter 4

Using y1=SavGol_bad(x,yn,5,4); 

rather than y1=SavGol(x,yn,5,4); 

within Main_SavGol.m, the result is a large number of messages of the type Warning: Rank deficient, rank = 4  tol =   5.2917e­006. 

These messages indicate that the rank of the matrix F is not 5, as expected for the case of a fourth order polynomial, but only 4. Why is this the case? At the beginning of the loop F is

1 1 1 ⎤ ⎡1 1 ⎢ ⎥ 2 3 2 24 ⎥ ⎢1 2 2 F=⎢ # # # # # ⎥ ⎢ ⎥ ⎢1 11 112 113 114 ⎥ ⎣ ⎦

(4.43)

This is perfectly ok. However, towards the end of the loop, the matrix F looks like ⎡1 140 1402 1403 1404 ⎤ ⎢ ⎥ ⎢1 141 1412 1413 1414 ⎥ F=⎢ ⎥ # # # # ⎥ ⎢# ⎢ 2 3 4⎥ ⎣1 150 150 150 150 ⎦

(4.44)

In a strictly mathematical sense this matrix is not singular but numerically it is rank deficient and has effectively a rank of only 4. Calculation of its pseudo-inverse consequently is impossible, or at least numerically unsafe. What can we do about that? It is important to realise that the fitting of a polynomial to a series of (x,y)­ data pairs is really independent of the actual values in the x-vector. In other words, in the above example, it does not matter whether the x-values are between 140 and 150 or between 1 and 11 or between -5 and +5. What matters is the relationship between the x-values and their y-values. As the data are equidistant, any equidistant vector with the right number of values can be used to generate F. In the improved SavGol function we chose the values from -n to +n. The second important observation is that consequently we do not have to recalculate F each time and more importantly, we do not have to recalculate its pseudo-inverse F+, which is computed outside the loop for all points. The result is a routine which is much faster and numerically much sounder: MatlabFile 4-13. SavGol.m function y1=SavGol(x,y,n,nd)  % Savitzki­Golay 

Model-Based Analyses

135

y1=zeros(size(x)); x1=(­n:n)'; F(:,1)=ones(size(x1)); for j=1:nd F(:,j+1)=F(:,j).*x1; end  F_plus=inv(F'*F)*F';  for i=1+n:length(x)­n a=F_plus*y(i­n:i+n); y1(i)=F(1+n,:)*a; end 

The routine SavGol.m is very basic. It does not include the beginning and the end of the curve, and it does not allow asymmetric selection of data points for the polynomial fitting. Both these features could easily be implemented. We leave it to the reader to improve the function accordingly. In its present form it cannot be used for non-equidistant x-values. F would have to be recalculated within the loop but the vector x1 used to generate F would still have to be centred around zero. Calculation of the Derivative of a Curve The computation of the derivative would be straightforward for perfect, noise-free data. The first derivative could be well approximated by dyi yi +1 − yi ≈ dxi xi +1 − xi

(4.45)

and to calculate higher derivatives the process is repeated. For noisy data, this approach results in a disaster. The noise component is amplified with each differentiation and soon there is nothing left but amplified noise. A possible remedy is to use the Savitzky-Golay filter. An initial idea might be tempting: first smooth the data as just demonstrated, and subsequently differentiate the treated data using equation (4.45). There is a better way: the fitted polynomial can be differentiated analytically; this explicit computation is computationally more efficient, both faster and numerically safer. The data set is the same as the one used for smoothing, with the same amount of noise added. The 'true' derivatives are computed using the Savitzky-Golay algorithm on the noise-free data. This is not quite correct but suffices here. The three panels of Figure 4-25 display the results of three different computations of the derivative. The first plot in the figure shows the derivative calculated as simple differences between noisy data. There is only noise left. The two next parts are the derivatives calculated as 2nd and 4th order polynomials through 11 points. Again, a compromise has to be sought between noise reduction and the loss of narrow features. As with smoothing, large numbers of data to define the polynomials and a low order polynomial result in smooth curves that might suffer from loss of narrow features.

136

Chapter 4

MatlabFile 4-14. Main_SavGol_Deriv.m % Main_SavGol_Deriv.m  x=(1:150)'; y=gauss(x,50,25)+gauss(x,100,10)+gauss(x,125,2); randn('state',0); yn=y+0.1*randn(size(y));  yd=SavGol_deriv(x,y,2,2);  for i=2:length(y); ys(i)=yn(i)­yn(i­1); end  y1=SavGol_deriv(x,yn,5,2); y2=SavGol_deriv(x,yn,5,4); 

  % the 'true' derivative 

subplot(3,1,1);plot(x,ys,'.',x,yd,'k'); axis([0 150 ­0.6 0.6]); subplot(3,1,2);plot(x,y1,'.',x,yd,'k'); axis([0 150 ­0.6 0.6]); subplot(3,1,3);plot(x,y2,'.',x,yd,'k'); axis([0 150 ­0.6 0.6]); xlabel('x'); 

0.5 0 -0.5 0

50

100

150

50

100

150

100

150

0.5 0 -0.5 0 0.5 0 -0.5 0

50 x

Figure 4-25. The top panel displays the true derivatives and those computed as the quotient of differences; the middle and bottom panels show the result of a 2nd and 4th degree polynomial fitted through 11 data points.

137

Model-Based Analyses

The calculation of the derivative of a general polynomial is straightforward:

yi = a1 + a2 xi + a3 xi2 + a 4 xi3 + ... + and +1xind = ⎛ dy ⎞ nd -1 2 = ⎜ ⎟ = a 2 + 2a3 xi + 3a 4 xi + ... + nd and +1x i ⎝ dx ⎠i

∑ a j xij -1

nd +1 j =1

∑ nd

j =1

(4.46)

ja j +1xij -1

Defying Matlab elegance, one could write equation (4.46) as a loop, but it is certainly faster to vectorise the equation. The vectorised Matlab code (note that the polynomial degree equals the number of parameters minus one, nd=np-1) dydx(i)=F(1+n,1:nd)*([1:nd]'.*a(2:nd+1)); 

is a bit less transparent. Here the 'explanation':

∑ ja j +1xij −1 = ⎡⎣ xi0 nd

j =1

x 1i

x i2

⎛ ⎡ 1 ⎤ ⎡ a2 ⎤ ⎞ ⎜⎢ ⎥ ⎢ ⎥⎟ ⎜ ⎢ 2 ⎥ ⎢ a3 ⎥ ⎟ " x ind −1 ⎤⎦ ∗ ⎜ ⎢ 3 ⎥ . ∗ ⎢ a 4 ⎥ ⎟ ⎜⎢ ⎥ ⎢ ⎥⎟ ⎜⎢ # ⎥ ⎢ # ⎥⎟ ⎜ ⎢nd ⎥ ⎢a ⎥⎟ ⎝ ⎣ ⎦ ⎣ nd +1 ⎦ ⎠

⎛ ⎡ 1 ⎤ ⎡ a(2) ⎤ ⎞ ⎜⎢ ⎥ ⎢ ⎥⎟ ⎜ ⎢ 2 ⎥ ⎢ a(3) ⎥ ⎟ = [F (n + 1,1) F (n + 1, 2) F (n + 1, 3) " F (n + 1, nd )] ∗ ⎜ ⎢ 3 ⎥ . ∗ ⎢ a (4) ⎥ ⎟ ⎜⎢ ⎥ ⎢ ⎥⎟ # ⎜⎢ # ⎥ ⎢ ⎥⎟ ⎜ ⎢nd ⎥ ⎢a (nd + 1)⎥ ⎟ ⎦ ⎣ ⎦ ⎣ ⎝ ⎠ MatlabFile 4-15. SavGol_deriv.m function dydx=SavGol_deriv(x,y,n,nd)  % Savitzki­Golay %  % polynomial interpolation, degree nd, through 2n+1 data points  dydx=zeros(size(x)); x1=(­n:+n)'; F(:,1)=ones(size(x1)); for j=1:nd F(:,j+1)=F(:,j).*x1; end  F_plus=inv(F'*F)*F';  for i=1+n:length(x)­n a=F_plus*y(i­n:i+n); dydx(i)=F(1+n,1:nd)*([1:nd]'.*a(2:nd+1)); end 

(4.47)

138

Chapter 4

Polynomial Interpolation The Savitzky-Golay algorithm could readily be adapted for polynomial interpolation. The computations are virtually identical to smoothing. In smoothing, a polynomial is fitted to a range of (x,y)-data pairs arranged around the x-value that needs to be smoothed. For polynomial smoothing, the polynomial is evaluated for a set number of data points around the desired x-value and the computed y-value at that x is the interpolated value. Polynomial fitting is a very important tool and, as expected, Matlab provides a set of functions for the task. Instead of adapting the Savitzky-Golay routine used previously, we demonstrate the handling of the Matlab routines polyfit.m and polyval.m. The developed function is a very general polynomial interpolation routine that deals with almost anything imaginable. An obvious name would be polypol, we couldn't resist the temptation and rearranged two consonants to turn it into a lolipop. MatlabFile 4-16. Main_lolipop.m % Main_lolipop  x=(1:.5:6)';              % x/y data pairs y=gauss(x,4,3); randn('seed',0); yn=y+0.02*randn(size(y)); x1=1:0.1:5;               % x1­values for which y1 is interpolated y1=lolipop(x,yn,x1,3,6);  % interpolation  plot(x,yn,'+',x1,y1,'.k'); xlabel('x');ylabel('y'); 

1.4 1.2 1

y

0.8 0.6 0.4 0.2 0 1

2

3

4

5

x

Figure 4-26. The result of polynomial interpolation.

6

Model-Based Analyses

139

In Main_lolipop.m a Gaussian curve is calculated at 0.5 intervals. A small amount of noise is added and this demonstrates a certain level of smoothing, resulting from polynomial fitting. The function lolipop.m interpolates the value y1 at position x1, using a polynomial of degree nd through points neighbouring (x,y)-data points whose x are closest to x1. MatlabFile 4-17. lolipop.m function y1=lolipop(x,y,x1,nd,npoints)  % General Polynomial Inter/Extrapolation, degree nd, using npoints % x,y,x1,y1 vectors ­ x,y do not have to be the same length as x1,y1 % nd: degree of polynomials % npoints: number of total points to define each polynomial  for i=1:length(x1) N=sortrows([x y abs(x­x1(i))],3); % sort x,y by abs(x­x1(i))  x_npoints=N(1:npoints,1);         % npoints nearest nodes y_npoints=N(1:npoints,2);  a=polyfit(x_npoints­mean(x_npoints),y_npoints,nd); % polyn. par. y1(i)=polyval(a,x1(i)­mean(x_npoints));            % interpolate end 

The Matlab functions polyfit.m and polyval.m need to be explained. a=polyfit(x,y,nd) fits a polynomial of degree nd to the x/y data pairs, a is the vector of coefficients defining the polynomial. Note that the elements of a are arranged in the opposite order than 'our' a as defined in (4.26). The command y1(i)=polyval(a,x1(i)) evaluates the polynomial defined by a at x1(i). Using lolipop.m the routine Main_lolipop.m fits a polynomial of degree nd=3 through 6 nodes, closest to x1(i). These 6 nodes are determined by the first 3 lines of the loop, taking advantage of the sortrow function. Note that lolipop.m also allows for extrapolation but choosing x1-values outside the range of x is not recommended.

4.2.8 Linear Regression with Multivariate Data In this chapter we expand the linear regression calculation into higher dimensions, i.e. instead of a vector y of measurements and a vector a of fitted linear parameters, we deal with matrices Y of data and A of parameters. We derive the new concept by using a chemical example based on absorption data. First, consider a consecutive reaction A→B→C, with rate constants k1 and k2, where the absorption at one particular wavelength was recorded as a function of time. Let's say our task is to determine the molar absorptivities of species A, B and C at this wavelength, knowing all individual concentrations at all reaction times. Previously we used the notation F for the matrix of the 'known' function. In many chemical applications involving spectroscopic absorption

140

Chapter 4

measurements, an equivalent matrix is made up of molar concentrations of several chemical species. In these circumstances, we call the matrix C, referring to Chapter 3.1, Beer-Lambert's Law.

The above example of recording the kinetics of the reaction A→B→C at one wavelength is then best described by the matrix equation. y = Ca + r

(4.48)

The (ns×1) column vector y contains the absorption data at ns reaction times; the concentration profiles of three species A, B and C form the columns of an (ns×3) matrix C and their molar absorptivities form an (3×1) column vector a. Vector r contains the residuals between y and C×a and has the same dimensions as y. Having the measurements y and supposedly knowing C, it is, as earlier in equation (4.30), a linear least-squares calculation that computes the best a: a = C+ y

(4.49)

The next step is to imagine having measured whole absorption spectra as a function of time, e.g. by using a diode array spectrophotometer. The kinetic traces at nl different wavelengths are arranged as columns of a matrix Y and, similarly, the molar absorptivities as columns of a matrix A, thus (4.48) transforms into

Y = CA +R

(4.50)

It is most helpful to recall the 'rectangle' notation for this equation introduced in Chapter 3.1, Beer-Lambert's Law:

nl ns

Y

nc

=

C

×

nl

A

nl nc

+ ns

R

Figure 4-27. The structure of Beer-Lambert's law in matrix notation. It is important to realise that each column a:,j can independently be calculated from the appropriate column y:,j, irrespective of all the other wavelengths, using equation (4.49). The pseudo-inverse C+ is the same for all. The equivalent of (4.49) for all wavelengths can be written as A = C + Y (in Matlab A = C \ Y )

(4.51)

Equation (4.51) minimises the sum of squares, ssq, of the residuals in R, defined by the multivariate equivalent to equation (4.2):

Model-Based Analyses

ssq = ∑ ∑ ri2, j

141

ns nl

i =1 j =1

(4.52)

The complete matrix A, containing the absorption spectra of the components A, B and C, is computed in one step! Presently, we are able to compute A knowing Y and C. Computing Y knowing C and A is trivial. What about calculating the concentration matrix C, knowing Y and A? We could transpose equation (4.50): Yt = A t Ct

(4.53)

Applying the equivalent of (4.51), we can write +

and transposing back

C = Y A+

Ct = A t Y t

(4.54)

(in Matlab C = Y / A )

(4.55)

This is, of course, the matrix equivalent of equation (4.31). There is another, more casual, way of deriving equations (4.51) and (4.55). We start with

Y = CA

(4.56)

multiply the equation with At from the right Y A t = CA A t

(4.57)

and then with (AA t )-1 , again from the right

to result in

noting that

Y A t (AA t )-1 = C AA t (AA t )-1

(4.58)

C = Y A t (AA t )-1 = Y A +

(4.59)

A + = A t (AA t )-1

(4.60)

The same sequence for the calculation of C, given Y and A:

142

Chapter 4

Y = CA

C Y = Ct C A t

(C t C)-1 C t Y = (C t C)-1 C t C A

(4.61)

A = (C C) C Y t

= C+ Y

-1

t

Of course, All these derivations and equations hold for any matrix product of the kind Y=CA, irrespective of what the physical meaning of the matrices is. In addition, vectors are just special matrices and the equations also hold for vectors. The derivation of equation (4.59) and its equivalent (4.61) is not mathematically proper, but more importantly, the results are correct in the least-squares sense. They are identical to the ones derived via the normal equations, e.g. equation (4.26). Figure 4-28 represents the shapes of the matrices. A Y

=

C

C+ =

A

Y

A+ C

=

Y

Figure 4-28. Schematic representations of the dimensions of the matrices in equations (4.50), (4.51) and (4.55). Referring back to Matlab, it is very important to use the correct slash operator \ or / for the left and right pseudo inverse. Applying the wrong one will invariably result in an error message or worse, in a potentially undetected error.

143

Model-Based Analyses

Applications First we construct a kinetic measurement which we then analyse in both of the above ways. The reaction is the set of two consecutive first order reactions with rate constants k1 and k2: k2 k1 A ⎯⎯⎯ → B ⎯⎯⎯ →C

(4.62)

Integration of the appropriate differential equations for the reaction scheme is straightforward, the resulting equations for the concentrations of A, B, and C as a function of time (see Chapter 3.4.2, Rate Laws with Explicit Solutions) are: [ A ] = [ A ]0 × e −k1t

[B ] = [ A ]0

(

k1 e −k1t − e −k2t k2 − k1

[C ] = [ A ]0 − [ A ] − [B ]

)

(4.63)

The absorption spectra are modelled by Gaussians (see 3.2 Chromatography / Gaussian Curves). MatlabFile 4-18. Data_ABC.m function [t,lam,Y,C,A]=Data_ABC  % absorbance data generation for A ­> B ­> C  t   =[0:25:4000]';     % reaction times  lam =400:10:600;       % wavelengths k   =[.003; .0015];    % rate constants  A_0=1e­3;              % initial concentration of A  C(:,1)=A_0*exp(­k(1)*t);     % concentrations of A  C(:,2)=A_0*k(1)/(k(2)­k(1))*(exp(­k(1)*t)­exp(­k(2)*t)); % conc. of B C(:,3)=A_0­C(:,1)­C(:,2);       % concentrations of C  A(1,:)=1e3*gauss(lam,450,50); % molar spectrum of A A(2,:)=4e2*gauss(lam,500,50); % molar spectrum of B A(3,:)=5e2*gauss(lam,550,50); % molar spectrum of C  Y=C*A;                        % applying Beer's law to generate Y randn('seed',0);              % fixed start for random number generator Y=Y+0.01*randn(size(Y));      % standard deviation 0.01 

The short routine Main_ABC_3D.m reads in the absorbance data modelled by Data_ABC.m and plots them in Figure 4-29 against wavelength and reaction time. MatlabFile 4-19. Main_ABC_3D % Main_ABC_3D  [t,lam,Y]=Data_ABC;  mesh(lam,t,Y); xlabel('wavelength') 

144

Chapter 4

ylabel('time') zlabel('absorbance') 

Figure 4-29. Mesh-plot of the absorption matrix for a consecutive reaction A→B→C, measured at several wavelengths. We have now a set of data we can analyse in the two ways.

Computation of Component Spectra, Known Concentrations Assuming we know the two rate constants k1 and k2 that allow the computation of C. Assuming further, we only have measured spectra between time = 200 and 1200 (fast reaction with significant dead time of the instrument). The task is to determine the three absorption spectra of the pure compounds A, B and C. All three are not accessible directly in the range of available spectra because of severe overlap. MatlabFile 4-20. Main_ABC_Lin1.m % Main_ABC_Lin1  [t,lam,Y,C,A]=Data_ABC; C_p=C(9:49,:); Y_p=Y(9:49,:);  A_calc=C_p\Y_p;     % component spectra via multivariate linear regression plot(lam,A,'k:',lam,A_calc,'k­'); xlabel('wavelength');ylabel('mol. absorptivity'); 

145

Model-Based Analyses

1200 1000

mol. absorptivity

800 600 400 200 0 -200 400

450

500 wavelength

550

600

Figure 4-30. Calculated and true absorption spectra of species A, B and C (from left to right).

Computation of Component Concentrations, Known Spectra The molar absorptivity spectra of species A, B and C have been determined Main_ABC_Lin2.m calculates the corresponding independently. concentration profiles of the 3 components using the complete data set from Data_ABC.m. They are shown in Figure 4-31. MatlabFile 4-21. Main_ABC_Lin2.m % Main_ABC_Lin2  [t,lam,Y,C,A]=Data_ABC;  C_calc=Y/A;      % conc. profiles via multivariate linear regression plot(t,C,'k:',t,C_calc,'k­'); xlabel('time'); ylabel('concentration'); 

146

Chapter 4

12

x 10

-4

10

concentration

8 6 4 2 0 -2 0

1000

2000 time

3000

4000

Figure 4-31. Calculated and true concentration profiles of species A, B and C. It is probably more realistic to assume that we know neither the rate constants nor the absorption spectra for the above example. All we have is the measurement Y and the task is to determine the best set of parameters which include the rate constants k1 and k2 and the molar absorptivities, the whole matrix A. This looks like a formidable task as there are many parameters to be fitted, the two rate constants as well as all elements of A. In Multivariate Data, Separation of the Linear and Non-Linear Parameters (p.162), we start tackling this problem.

The Pseudo-Inverse in Excel We have encountered Excel's LINEST as a tool for linear regression. Unfortunately, LINEST cannot be generalised from vectors to matrices. To deal with matrices, we do not have an option but to use equations (4.59) and (4.61). It is possible to do so, but not as convenient as in Matlab. In order to keep the spreadsheet reasonably small, the dimensions are much smaller than those in the Matlab examples. It is still a consecutive reaction scheme; the spectra were recorded at 11 times and at 6 wavelengths.

147

Model-Based Analyses ExcelSheet 4-3. Chapter3.xls-pseudoinverse

=TRANSPOSE(C5:E15)

=MMULT(G5:Q7,C5:E15)

=MINVERSE(G10:I12)

=MMULT(G5:Q7,A18:F28)

=MMULT(K10:M12,H15:M17)

=MMULT(MINVERSE(MMULT(TRANSPOSE(C5:E15) ,C5:E15)),MMULT(TRANSPOSE(C5:E15),A18:F28))

Figure 4-32. Excel spreadsheet applying the equation A=(Ct C)-1 Ct Y in two ways, stepwise and in one big formula. There are two paths to reach the result. The first path is a stepwise construction of a series of intermediate matrices, they are framed in Figure 4-32. (a) Ct (b) Ct C (c) (Ct C)-1 (d) Ct Y and finally (e) (Ct C)-1 Ct Y This approach uses a lot of space on the spreadsheet, in particular the transpose of a long column is a very wide row. However, it is reasonably easy to detect potential errors in the formulas. The other path does the whole calculation in one step, the result is the shaded matrix Aone_eq. This is a very neat approach on the spreadsheet, but a very difficult equation to be entered in the matrix mode: =MMULT(MINVERSE(MMULT(TRANSPOSE(C5:E15),C5:E15)),MMULT(TRANS POSE(C5:E15),A18:F28)) Remember to preselect the rectangle of correct dimensions and use CtrlShift-Enter for matrix equations in Excel!

148

Chapter 4

4.3 Non-Linear Regression Non-linear regression calculations are extensively used in most sciences. The goals are very similar to the ones discussed in the previous chapter on Linear Regression. Now, however, the function describing the measured data is non-linear and as a consequence, instead of an explicit equation for the computation of the best parameters, we have to develop iterative procedures. Starting from initial guesses for the parameters, these are iteratively improved or 'fitted', i.e. those parameters are determined that result in the optimal 'fit', or, in other words, that result in the minimal sum of squares of the residuals. There are a multitude of methods for this task. Those that are conceptually simple usually are computationally intensive and slow, while the fast algorithms have a more complex mathematical background. We start this chapter with the Newton-Gauss-Levenberg/Marquardt algorithm, not because it is the simplest but because it is the most powerful and fastest method. We can't think of many instances where it is advantageous to use an alternative algorithm. Because of its relative complexity and tremendous usefulness, we develop the Newton-Gauss-Levenberg/Marquardt algorithm in several small steps and thus examine it in more detail than many of the other algorithms introduced in this book.

4.3.1 The Newton-Gauss-Levenberg/Marquardt Algorithm Later, in Chapter 4.4, General Optimisation, we discuss non-linear leastsquares methods where the sum of squares is minimised directly. What is meant with that statement is, that ssq is calculated for different sets of parameters p and the changes of ssq as a function of the changes in p are used to direct the parameter vector towards the minimum. In this section we demonstrate that it is possible to use the complete vector or matrix of residuals to drive the iterative refinement towards the minimum. As expected in an iterative algorithm, we start from an initial guess for the parameters. This parameter vector is subsequently improved by the addition of an appropriate parameter shift vector δp, resulting in a better, but probably still not perfect, fit. From this new parameter vector the process is repeated until the optimum is reached. As with almost any other non-linear problem that has to be solved iteratively, linearisation via a Taylor expansion with truncation after very few elements, is the solution. In Chapter 4.1 Background to Least-Squares Methods, e.g. in Figure 4-3 and Figure 4-5, we have seen that for univariate data, the vector r of residuals and thus the sum of squares ssq, is a function of the measurement y and the parameters p of the model of choice.

Model-Based Analyses

r = f (y, p)

149

(4.64)

The basic principle of the algorithm is to add a shift vector δp to the parameter vector p. The shift vector is computed with the aim of producing a new parameter vector for which ssq is minimal, or at least smaller.

The residuals r(p+δp) after the application of the shift vector, are approximated by a Taylor series expansion. With sufficient terms, any precision for the approximation can be achieved. r(p + δp) = r(p) +

1 ∂r(p) 1 ∂r 2 (p) 2 δp + δp + ... 1! ∂p 2! ∂p2

(4.65)

As done previously, in The Newton-Raphson Algorithm (p.48), we neglect all but the first two terms in the expansion. This leaves us with an approximation that is not very accurate but, since it is a linear equation, is easy to deal with. Algorithms that include additional higher terms in the Taylor expansion, often result in fewer iterations but require longer computation times due to the calculation of higher order derivatives. r(p + δp) ≅ r(p) +

∂r(p) δp = r(p) + Jδp ∂p

(4.66)

The derivative ∂r(p)/∂p is known as the Jacobian J.

The task is to compute the ‘best’ parameter shift vector δp that minimises the new residuals r(p+δp) in the least-squares sense. This is a linear regression equation with the explicit solution. δp = − J + r(p)

(4.67)

Note that equation (4.66) ( r(p + δp) ≅ r(p) + Jδp ) has the same structure as equation (4.17) ( r = y − Fa ) where the calculation of a was a = F + y .

The Taylor series expansion is always only an approximation and therefore the shift vector δp will not result in the minimum directly. However, the new parameter vector p+δp will usually be better than the preceding p. Thus, an iterative process should move towards the optimal parameters.

A First, Minimal Algorithm We are now in a position to devise a first, very crude program that should, starting from a set of initial guesses, move towards the best fit. Below, a flow diagram is given that represents the basic principle of the Newton-Gauss algorithm:

150

Chapter 4

guess parameters, p=pstart

calculate residuals, r(p) and the sum of squa squares, res, ssq

calculate Jacobian J calculate shift vector δp, and p = p + δp

Figure 4-33. First version of the Newton-Gauss algorithm The crucial part of this algorithm is the computation of J, the derivatives of the residuals with respect to the parameters. It might be best to demonstrate this by an example. An exponential curve, including some noise, is generated by the function Data_exp.m. The curve is defined by three parameters, the rate, p1, the amplitude p2 and the value at infinity time p3. y = p3 + p2 e

− p1 t

(4.68)

MatlabFile 4-22. Data_exp.m function [t,y]=Data_exp  p=[2e­2;­4;10]; t=(1:2:100)'; y=p(3)+p(2)*exp(­p(1)*t);  randn('seed',0); y=y+5e­2*randn(size(y));  MatlabFile 4-23. Main_exp_2d.m % Main_exp_2d  [t,y]=Data_exp; plot(t,y,'.'); xlabel('time'); ylabel('y'); 

The main routine Main_exp_2d.m reads in the data and plots them in Figure 4-34.

151

Model-Based Analyses

10 9.5 9

y

8.5 8 7.5 7 6.5 6 0

20

40

60

80

100

time

Figure 4-34. An exponential function The derivatives of the vector r of residuals with respect to the parameter vector p are given by the following equations. Note that the first column of the Jacobian matrix contains the derivative of the residuals with respect to the first parameter, the second with respect to the second, etc. In this example the derivatives can be computed explicitly., later we will introduce the computation of numerical derivatives. r = y − ycalc

= y − ( p3 + p2 e − p1 t )

j:,1 =

j:,2 = j:,3 =

∂r = p2 t e − p1 t ∂p1

(4.69)

∂r = −e − p1 t ∂p2 ∂r = −1 ∂p3

With that, we are in a position to write a short program, Main_NG1.m, that iterates towards the optimal set of parameters and fits the curve in Figure 4-34, provided the initial guesses are reasonable. MatlabFile 4-24. Main_NG1.m % Main_NG1  [t,y]=Data_exp;  p=[.01; ­3 ; 15 ]; % initial guesses [rate const, amp, inf] 

152

Chapter 4

for i=1:10  y_calc=p(3)+p(2)*exp(­p(1)*t); r=y­y_calc; ssq(i)=sum(r.*r);  J(:,1)=p(2)*t.*exp(­p(1)*t); J(:,2)=­exp(­p(1)*t); J(:,3)=­1;  delta_p=­J\r;  % calculate parameter shifts p=p+delta_p;  % add parameter shifts end  p subplot(1,2,1); plot(t,y,'.',t,y_calc);xlabel('time');ylabel('y'); subplot(1,2,2); plot(log(ssq),'+');xlabel('iteration');ylabel('log(ssq)');  p =  0.0195  ­3.9882  10.0252 

10

8

9.5

6

9 4 log(ssq)

y

8.5 8

2

7.5 0 7 -2

6.5 6 0

50 time

100

-4 0

5 iteration

10

Figure 4-35. Fitted exponential and iterative decrease of ssq The left panel of Figure 4-35 shows the result of the fit. The right panel displays the sum of squares for each iteration, featuring a continuous decrease. The program at this stage is very crude and needs several stages of improvements. For the next version we implement two new measures: a

Model-Based Analyses

153

proper termination criterion and numerical calculation of the derivatives for the calculation of the Jacobian.

Termination Criterion, Numerical Derivatives We start with the termination criterion. The right panel of Figure 4-35 immediately tells us that the iterations 6 to 10 are wasted. The minimal ssq has been reached at the fifth iteration and there is no further improvement. There are different ways of testing whether there is continuing improvement of the fit or whether the progress is finished and, hopefully, the best minimal ssq has been reached. In the progress of the iterations, the shifts δp, as well as the sum of squares, usually decrease continuously. Thus both could be inspected for constancy. The most common and intuitively correct test is the constancy of the sum of squares, as indicated in Figure 4-35. If a generally applicable routine is envisaged, it is not possible to test the absolute difference between old and new square sum. Depending on the data, ssq can be very small or very large. Therefore, a convergence criterion analysing the relative change in ssq has to be applied. The iterations are stopped once the absolute change is less than a preset value μ, typically μ=10-4. abs(

ssqold − ssq )≤ μ ssqold

(4.70)

guess parameters, p=pstart

calculate residuals, r(p) and the sum of squares, ssq

ssq const ?

yes

end; display results

no calculate Jacobian J calculate shift vector δp, and p = p + δp

Figure 4-36. Improved Newton-Gauss algorithm, including a termination criterion.

154

Chapter 4

And now we introduce numerical derivatives. In the example above, we used explicit formulas for the derivatives of the residuals with respect to the parameters. Often it is not easy, or even impossible, to work out the correct equations. Numerical computation of the derivatives is always possible. Usually it is slower and also numerically less accurate. The general formula is: r(p + Δpi ) − r(p) ∂r ≅ Δpi ∂pi

(4.71)

This is a rather casual notation and we need to clarify what is meant. p+Δpi is a new parameter vector with only the i-th parameter pi shifted by the small amount Δpi. In Main_NG2.m, Δpi is calculated as 1×10-4 pi. The factor 1×10-4 is somewhat arbitrary and experimentation is usually the best way of ∂y ∂r ∂(y − y calc ) determining the optimal value. Note that = = − calc since ∂p ∂p ∂p ∂y = 0 due to the invariance of the measured data y; thus we only need to ∂p compute ycalc not r, which is slightly quicker. In equation (4.72) this is worked out for the example. j:,1 = − j:,2 = − j:,3 = −

( p + p2 e −1.0001 p1 t ) − ( p3 + p2 e − p1 t ) ∂y calc =− 3 ∂p1 10−4 p1

( p + 1.0001 p2 e − p1 t ) − ( p3 + p2 e − p1 t ) ∂y calc =− 3 ∂p2 10−4 p2

(4.72)

(1.0001 p3 + p2 e − p1 t ) − ( p3 + p2 e − p1 t ) ∂y calc =− ∂p3 10−4 p3

The Matlab program Main_NG2.m has implemented the additions for a termination criterion and numerical derivatives. Refer to the Matlab Help Desk for information on the while end loop and also the break command. MatlabFile 4-25. Main_NG2.m % Main_NG2  [t,y]=Data_exp;  p=[.01; ­5 ; 15 ];          % initial guesses [rate const, inf, amp] ssq_old=1e50;  while 1  y_calc=p(3)+p(2)*exp(­p(1)*t); r=y­y_calc; ssq=sum(r.*r)  if  (ssq_old­ssq)/ssq_old mp×5

mp/3

mp=0

calculate Jacobian J calculate shift vector δp, and p = p + δp

Figure 4-38. The Newton-Gauss algorithm after implementation of the Marquardt strategy

In most instances the algorithm converges straightaway. In order to test the Marquardt extension, we need more difficult data to analyse. The function Data_chrom.m generates an overlapping set of two Gaussian peaks. Each

158

Chapter 4

peak is characterised by peak height, peak position and peak width. Altogether there are six parameters to be fitted. Data_Chrom.m generates the data, they are represented in Figure 4-39. Recall the function gauss.m that generates Gaussians, as introduced in Chapter 3.2, Chromatography / Gaussian Curves. MatlabFile 4-26. Data_Chrom.m function [t,y]=Data_chrom  t=(1:2:100)'; y=2*gauss(t,40,25)+3*gauss(t,70,30); randn('seed',0); y=y+5e­2*randn(size(y));       % noise of standard deviation 5e­2 

The function nglm.m accepts the parameters p, the independent variables x, the measurements y and additionally the string fname, which contains the name of the function file that determines the residuals based on the model. If a different model is to be fitted, all that needs to be done is to replace fname in the calling routine, and, of course, supplying the function with that new name. Importantly, nglm.m is not affected at all. nglm.m returns the fitted parameters p and the minimal sum of squares. The function is a Matlab translation of the flow chart in Figure 4-38. The Jacobian is computed numerically, which makes the routine very generally applicable. The few parameters defined at the very beginning are chosen somewhat arbitrarily. Depending on specific problems, different values might have to be used. This is particularly the case for the Marquardt parameter. In stubborn cases, the user should try a different initial value.

The main program Main_chrom.m loads the measurement, generates initial guesses for the parameters, calls nglm.m and reports the fitted parameters. fprintf is a C based Matlab command that allows compact output. The reader is invited to read up on this command in the Matlab manuals. MatlabFile 4-27. Main_Chrom.m % Main_chrom  [t,y]=Data_chrom; p0=[2;40;20;3;50;10];   % guesses [height,center,width;...] [p,ssq]=nglm('Rcalc_chrom',p0,t,y);  % call ngl/m fprintf(1,'\n'); for i=1:length(p) fprintf(1,'p(%i) = %g\n',i,p(i)); end  it=0, ssq=107.204, mp=0, conv_crit=1 it=1, ssq=601.46, mp=0, conv_crit=­4.61043 it=2, ssq=96.9914, mp=1, conv_crit=0.0952631 it=3, ssq=85.0839, mp=0.333333, conv_crit=0.122769 it=4, ssq=47.7829, mp=0.111111, conv_crit=0.438403 it=5, ssq=23.7545, mp=0.037037, conv_crit=0.502867 it=6, ssq=4.22909, mp=0.0123457, conv_crit=0.821967 it=7, ssq=10.7692, mp=0.00411523, conv_crit=­1.54646 it=8, ssq=5.00616, mp=0.0205761, conv_crit=­0.183744 

Model-Based Analyses it=9, ssq=0.79927, mp=0.102881, conv_crit=0.811006 it=10, ssq=0.174235, mp=0.0342936, conv_crit=0.782008 it=11, ssq=0.104042, mp=0.0114312, conv_crit=0.402863 it=12, ssq=0.103976, mp=0.00381039, conv_crit=0.00063486 it=13, ssq=0.103976, mp=0.00127013, conv_crit=3.87491e­007 it=14, ssq=0.103976, mp=0, conv_crit=3.87501e­007 p(1) = 2.00292 p(2) = 40.4661 p(3) = 25.9485 p(4) = 2.97704 p(5) = 70.3259 p(6) = 29.57  MatlabFile 4-28. nglm.m function [p,ssq,Curv]=nglm(fname,p,t,y)  ssq_old=1e50; mp=0;  mu=1e­4;  delta=1e­6;  it=0; 

          % Marquardt parameter            % convergence limit            % step size for numerical diff 

while it0)                   % check if all absorbances>0  us_extrap=[US_bar(m,:)+inc*dir]     % move  plot(us_extrap(1),us_extrap(2),'k+'); y_extrap=us_extrap*V_bar;           % spect. in real absorbances inc=inc+.05; end  y_last=y_extrap;                        % first impossible spectrum  inc=0;                                  % init. step size for move y_extrap=Y(1,:); while all(y_extrap>0)                   % check if all absorbances>0  us_extrap=[US_bar(1,:)­inc*dir];    % move  plot(us_extrap(1),us_extrap(2),'k+'); y_extrap=us_extrap*V_bar;           % spect. in real absorbances inc=inc+.05; end  y_first=y_extrap;                       % first impossible spectrum subplot(1,2,2) plot(lam,Y,'­',lam,y_first,'­­',lam,y_last,'­­'); xlabel('wavelength');ylabel('absorbance'); 

Figure 5-16 displays the results of the analysis. On the left, the • markers ¯ , the + markers represent the measured spectra in the eigenvector space V the extrapolated values. The full lines in the right panel are the series of spectra that were used for the analysis and the dashed lines are the extrapolated boundary spectra. The spectrum of B is fairly well defined while the spectrum of A is not, there is a long extrapolation required until the its spectrum turns negative at 400nm. This exemplifies the limitation of model-free methods: they rely on very simple constraints but in certain cases the range of feasible answers can be very wide, sometimes too wide to be useful. This will be discussed later in Chapter 5.4.3, Rotational Ambiguity.

235

Model-Free Analyses

While the Lawton-Sylvestre method is very elegant and simple, it is virtually impossible to extend the principle to 3 and more components. 1

1. 1.8 8

0. 0.5 5

1. 1.6 6

B

1. 1.4 4

0

absorbance absor bance

(us (us ):,2

-1 -1. -1.5 -2

A

1 0. 0.8 8 0. 0.6 6 0. 0.4 4

-2. -2.5

0. 0.2 2

-3

0

-3. -3.5 -4

A

1. 1.2 2

-0. -0.5

-2 (us (us):,1

0

-0.2 -0.2 400

B

50 500 0 wav wavelength

600

Figure 5-16. The Lawton-Sylvestre analysis in action. The double arrows cover the feasible regions of positive absorbances.

Three and More Components So far we restricted our deliberations to 2-component systems. It is possible to increase this number to 3 and still comprehend the action in a 3­ dimensional space. We can even project the 3-dimensional space onto the plane of the paper or computer screen and 'see' what is going on. As usual, we demonstrate the procedures based on a chemical process. Instead of another kinetics example, we use a spectrophotometric titration. The experiment follows the deprotonation of a two-protic acid by measuring the absorption spectra of the solution as a function of pH. The equilibria are quantitatively described by equation (5.21). 1 ⎯⎯⎯⎯ → AH A + H ←⎯⎯⎯⎯

log K

2 ⎯⎯⎯⎯→ AH + H ←⎯⎯⎯⎯ AH 2

log K

(5.21)

236

Chapter 5

The concentrations of the differently protonated species, as a function of pH, are calculated with the explicit function we developed in Special Case: Explicit Calculation for Polyprotic Acids, (p.64). A data matrix Y is constructed as before. Data_eqAH2a.m generates the data, it is called by Main_eqAH2a.m. MatlabFile 5-11. Data_eqAH2a.m function [pH,lam,Y,C,A]=Data_eqAH2a  pH=[2:.1:12]';                   % pH range H=10.^(­pH); logK=[8 6];                      % protonation constants K=10.^logK; C_tot=1e­3;                      % [AH2]+[AH]+[A] n=length(logK);                  % number of protons  denom=zeros(size(H)); for i=0:n  num(:,i+1)=H.^i*prod(K(1:i)); % numerator denom=denom+num(:,i+1);       % denominator  end  alpha=diag(1./denom)*num;        % degree of dissociation C=C_tot*alpha;                   % concentration profiles  lam=400:10:600;                  % wavelength range A(1,:)=1000*gauss(lam,450,120);  % component spectra A(2,:)=2000*gauss(lam,350,120); A(3,:)=1000*gauss(lam,500,50); Y=C*A;                           % absorbance data  randn('seed',0); Y=Y+1e­3*randn(size(Y));         % noise level 0.001 

MatlabFile 5-12. Main_eqAH2a.m % Main_eqAH2a  [pH,lam,Y,C,A]=Data_eqAH2a;  subplot(2,1,1); plot(lam,A); xlabel('wavelength');ylabel('absorptivity'); subplot(2,1,2); plot(pH,C); xlabel('pH');ylabel('concentration'); 

Each spectrum, measured during the titration, forms a row of the data matrix Y, and is a vector in an nl-dimensional space (in the example nl=21). As this is a three component system, all the vectors lie in a 3-dimensional sub-space. Each measured spectrum is a linear combination of the 3 component spectra shown in Figure 5-17.

237

Model-Free Analyses

absorptivity

1500 1000 500 0 400

concentration

1

x 10

450

500 wavelength

550

600

-3

0.5

0 2

4

6

8

10

12

pH

Figure 5-17. Molar absorption spectra and concentration profiles for the titration of a diprotic acid with logK values of 8 and 6. ¯ and U ¯. Thus, the rank of Y is 3 − there are 3 significant eigenvectors in V ¯ form a set of three basis vectors in the spectral space. The row vectors of V The coordinates of each vector yi,:, in this new system of axes, are given by ¯¯ , see equations (5.17) and (5.18) or in a different the i-th row of the matrix US notation: YV t = USVV t = US

(5.22)

MatlabFile 5-13. Main_EV_space.m % Main_EV_space  [pH,lam,Y,C,A]=Data_eqAH2a; [U,S,Vt]=svd(Y,0); US=U*S;  plot3(US(:,1),US(:,2),US(:,3),'.',US(:,1),US(:,2),0*US(:,3)­.5) grid on xlabel('us_{:,1}');ylabel('us_{:,2}');zlabel('us_{:,3}'); 

To support comprehension of the 3-dimensional character of Figure 5-18, we added the projection of the curve onto the bottom of the plot at z=-0.5. The titration starts at the right hand end of the trace. As can be seen from Figure 5-17, the first spectrum at pH 2 is essentially pure AH2. There is not much change in the concentrations up to pH 4 and the spectrum vectors are very similar. Then, as the pH approaches the first logK-value, the measured spectrum starts to move towards the spectrum of the intermediate AH.

238

Chapter 5

1

us:,3

0.5

0

-0.5 1 -1

0 -2

-1 us

:,2

-2

-3

us

:,1

¯­ Figure 5-18. Representation of the measured spectra y:,i in V space for an AH2 titration. At pH 7 there is a maximum in the concentration of AH and the measured spectrum is close to that of pure AH. With further increase in pH the spectrum veers towards the spectrum of fully deprotonated A which is represented at the end of the series. A small side issue deserves mentioning: as discussed in connection with Figure 5-13 and equation (5.22), the system of axes is formed by the ¯ , the coordinates of the spectra in Y are the rows of the matrix eigenvectors V ¯¯ . It is not automatically clear how the axes in a plot like Figure 5-18 US ¯ or US ¯¯ ? The equivalent question can be should be labelled; should it be V posed for Figure 5-17; should the abscissa in the top panel be labelled 'wavelength' or 'nm'? Both are correct. The matrix Y can be regarded as a row or as a column matrix and consequently we can also concentrate on the columns of Y rather than the rows. The columns are linear combinations of the concentration profiles of the species and they all lie in a 3-dimensional space as well. The columns of ¯ form a basis in this space. And the coordinates of each column the matrix U ¯¯ . vector of Y are contained in the columns of the matrix SV MatlabFile 5-14. Main_EV_space.m …continued % Main_EV_space, ...continued  SV=S*Vt'; plot3(SV(1,:),SV(2,:),SV(3,:),'.',SV(1,:),SV(2,:),0*SV(3,:)­2) grid on xlabel('sv_{1,:}');ylabel('sv_{2,:}');zlabel('sv_{3,:}'); 

239

Model-Free Analyses

3 2 sv3,:

1 0 -1 -2 4 2

0 0

-5

-2 sv2,:

-4

-10

sv1,:

Figure 5-19. Representation of the measured absorption profiles in ¯ space. the U The lowest wavelength at 400nm is represented in Figure 5-19 at the end of the trace on the top left. For increasing wavelengths, the profiles move to the right. As expected, the trace in Figure 5-19 is less ordered than the equivalent in Figure 5-18. Concentration profiles are governed by the law of mass action ¯¯ , is structured and closure and thus the trace, following the rows of US accordingly. No such law governs the relative shape of the absorption ¯¯ . spectra and the trace following the columns of SV Mean Centring, Closure In Figure 5-10, we have seen that the law of conservation of mass dictates that in the 2-component case all measured spectra lie on a straight line. In the present context this property is called closure. In general terms it means that the sum of all species concentrations is constant during an experiment. In the 2-component case the spectral action occurs in a 2-dimensional subspace. If the system is closed, the action is concentrated in a 1­ dimensional space. Similarly in a closed 3-component case, the action is concentrated in a 2-dimensional subspace. Back to the data set for the titration of the 2-protic acid, Data_eqAH2a.m. Due to closure of the chemical system, the sum of all concentrations [A]+[AH]+[AH2] is constant, and as a result, the curve in Figure 5-18 lies in a plane.

240

Chapter 5

The fact that the spectral vectors in a closed system lie in an further reduced sub-space, in a 2-component system they lie on a straight line, in a 3­ component system, in a plane etc., suggests that we could move the origin of the system of axes into that sub-space and in this way the number of relevant dimensions is reduced by one. We subtract the mean spectrum from each measured spectrum yi,: and as a result, the origin of the system of axes is moved into the mean. In the above example, it is into the plane of all spectral vectors. This is called meancentring. Mean-centring is numerically superior to subtraction of one particular spectrum, e.g. the first one. The Matlab program, Main_MeanCenter.m, performs mean-centring on the titration data and displays the resulting curve in such a way that we see the zero us:,3­ component, i.e. the fact that the origin (+) lies in the (us:,1,us:,2)-plane. MatlabFile 5-15. Main_MeanCenter.m % Main_MeanCenter 

[pH,lam,Y,C,A]=Data_eqAH2a;  Y_mc=Y­repmat(mean(Y,1),length(pH),1);  [U,S,Vt]=svd(Y_mc,0); US=U*S; 

% subtract mean spectrum

plot3(US(:,1),US(:,2),US(:,3),'.',0,0,0,'+') grid on axis([­1.5 1.5 ­1.5 1.5 ­1.5 1.5]); view(­70, 2); xlabel('us_{:,1}');ylabel('us_{:,2}');zlabel('us_{:,3}'); 

1.5 1

us:,3

0.5

0 -0.5 -1 -1.5 1

0

-1 us:,2

-1

0

1 us:,1

Figure 5-20. Mean-centring moved the origin of the system of axes into the centre of the action. This reduces the dimension of the subspace by one.

Model-Free Analyses

241

The argument can be turned around. If mean-centring reduces the rank of the matrix by one, the data set is closed. We have to be careful. The symmetry between columns and rows of the matrix Y is not complete. Closure is a property of the concentration profiles only and thus applies only in one dimension. The command mean(Y,1) computes the mean of each column of Y and the resulting mean spectrum is subtracted from each individual spectrum. There is no equivalent in the other direction. Subtracting the mean column from the columns does not reduce the rank, however, it moves the origin to the centre of the action. While not reducing the rank, it does reduce the absolute values of the numbers and improves the numerical accuracy of the computations. The improvement is often insignificant and usually only marginal. We generally refrained from performing mean-centring, with the exception of PCR and PLS in Chapter 5.6. HELP Plots Plots of the kind represented in Figure 5-18 and in Figure 5-19 are more than just graphically appealing. A considerable amount of useful information can be extracted from these plots of the spectra or concentration profiles in their respective eigenvector spaces. Consider the multivariate chromatographic data, Data_Chrom2.m (p.219) of a 3-component system as shown in Figure 5-3. MatlabFile 5-16. Main_HELPP.m % Main_HELPP  [t,lam,Y,C,A]=Data_Chrom2; [U,S,Vt]=svd(Y,0); US=U*S;  plot3(US(:,1),US(:,2),US(:,3),'.',US(:,1),US(:,2),0*US(:,3)­1); hold on;plot3(0,0,0,'o','MarkerSize',10);hold off; grid on xlabel('us_{:,1}');ylabel('us_{:,2}');zlabel('us_{:,3}'); 

From Figure 5-21, there are a few observations we can make: This is a three component system, but as it is not closed, the action occupies all three dimensions. It does not occur in a plane.

The path starts and ends at the origin (marked by サ). Figure 5-22 reveals that there are no components eluting at the beginning and end of the chromatogram, and therefore the respective spectral vectors contain just noise.

242

Chapter 5

1

us:,3

0.5 0 -0.5 -1 2 1

2 0

0 -1

us:,2

-2 -2 -4

us

:,1

¯­ Figure 5-21. The spectra yi,: of a 3-component chromatogram in V space. The path takes off from the origin in an almost straight line and returns to the origin in an almost straight line. This is exploited in HELP plots (Heuristic Evolving Latent Projections). If a section of the path is on a straight line and its extension goes through the origin, this is an indication that there exists only one component in that section of the measurement. In the example, this is the case at the beginning and end of the overlapped concentration profiles. Figure 5-22 reveals that during times 15-25 only the first component is present and during times 70-95 only the third. MatlabFile 5-17. Main_HELPP.m …continued % Main_HELPP, ...continued  plot(t,C) xlabel('time');ylabel('concentration'); 

The useful aspect of this follows: we can determine the regions in the series of spectra in which there is only one component. The spectral vectors are all parallel and the average over all spectra in the region is a good estimate for the pure component spectrum. The main difficulty with this approach is to decide when exactly the deviation from a straight line starts and thus, which selection of spectra we need to average.

243

Model-Free Analyses

2

x 10

-3

1.8 1.6

concentration

1.4 1.2 1 0.8 0.6 0.4 0.2 0 0

20

40

60

80

100

time

Figure 5-22. The concentration profiles from Data_Chrom2.m.

Noise Reduction ¯ and their respective Retaining only the significant singular values S ¯ and V ¯ , as indicated in Figure 5-2 and equation (5.10), results eigenvectors U in a substantial reduction of the size of the matrices needed to represent the ¯ . There is an additional valuable benefit: Y ¯ not only original matrix Y as Y ¯ is represents all relevant information contained in the original Y, Y somewhat better as it contains much less noise than Y. This is demonstrated in Main_NoiseRed1.m, using the kinetic data set Data_AB.m (p.224). MatlabFile 5-18. Main_NoiseRed1.m % Main_NoiseRed1  [t,lam,Y0,C,A]=Data_AB; Y=Y0+.05*randn(size(Y0)); ne=2;  [U,S,Vt]=svd(Y,0); U_bar=U(:,1:ne);Vt_bar=Vt(:,1:ne);S_bar=S(1:ne,1:ne); Y_bar=U_bar*S_bar*Vt_bar';  subplot(3,1,1); plot(lam,Y0,'­');axis tight;ylabel('Y0'); subplot(3,1,2); plot(lam,Y,'­');axis tight;ylabel('Y'); subplot(3,1,3); plot(lam,Y_bar,'­');axis tight; xlabel('wavelength');ylabel('Y_{bar}'); 

244

Y0

Chapter 5

1 0.8 0.6 0.4 0.2 400

450

500

550

600

450

500

550

600

450

500 wavelength

550

600

Y

1 0.5 0 400

Ybar

1 0.5 0 400

Figure 5-23. Absorbance spectra with noise level 10-3 (top panel) and an increased noise level of 5×10-2 (second panel). The third ¯ retaining 2 significant factors. panel represents Y The graphs in Figure 5-23 are convincing. The top panel displays the original data for a simple first order reaction A→B. The next panel shows the same data after the addition of a substantial amount of noise. The third panel ¯ =U ¯S ¯V ¯ with 2 eigenvectors. Clearly a features the reconstructed matrix Y substantial amount, but not all, of the noise, was removed. It is worthwhile investigating this particular aspect of Factor Analysis more deeply. Data_AB2.m generates data for a first order reaction where only the first component A absorbs. The rank of Y is then only one. MatlabFile 5-19. Data_AB2.m function [t,lam,Y,C,A]=Data_AB2  A_0=1e­3;       % initial concentration  k=2e­2;         % rate constant  lam=400:10:600; A(1,:)=1000*gauss(lam,450,120);     % spectrum of component A only t=(1:2:100)'; 

245

Model-Free Analyses

C(:,1)=A_0*exp(­k*t);  Y=C*A; randn('seed',0); Y=Y+5e­2*randn(size(Y)); 

¯ against The program Main_NoiseRed2.m plots three columns of Y and Y each other.

MatlabFile 5-20. Main_NoiseRed2.m % Main_NoiseRed2  [t,lam,Y,C,A]=Data_AB2; ne=1;  [U,S,Vt]=svd(Y,0); U_bar=U(:,1:ne);Vt_bar=Vt(:,1:ne);S_bar=S(1:ne,1:ne); Y_bar=U_bar*S_bar*Vt_bar';  plot3(Y(:,5),Y(:,10),Y(:,15),'+',Y_bar(:,5),Y_bar(:,10),Y_bar(:,15),'.') xlabel('Y_{:,5}');ylabel('Y_{:,10}');zlabel('Y_{:,15}');grid on 

In Figure 5-24, the +'s are the noisy original (yi,5,yi,10,yi,15)-data points. The •'s represent the corresponding factor analytically reproduced data points. The noise reduction is obvious, however, note that the distribution of the •'s along the line is still noisy. This manifests in the irregular distribution of the markers.

0.3

Y:,15

0.2 0.1 0 -0.1 0.8 0.6

1 0.4

0.5

0.2 Y:,10

0

0

Y:,5

Figure 5-24. 3-dimensional plot of the 5th column of Y (+) or Y (•) versus the 10th versus the 15th.

246

Chapter 5

The Figure 5-25 attempts to provide a geometrical representation of the situation.

yii,:,:

v2,: ytruei,:

yi ,:

v1,:

Figure 5-25. The relationship between the original measurement vector yi,:, its projection into the eigenvector space and the 'true' vector ytruei,:. The eigenvectors v1,: and v2,: span the grey plane; yi,: is the i-th spectrum, its ¯ -plane is y ¯ . The coordinates of spectrum y projection onto the V ¯ i,:=(us ¯¯ )i,:V ¯ i,: in ¯ -space are (us ) V . The hypothetical true, noise-free spectrum ytruei,: (which is ¯¯ i,: not known) usually lies close to, but generally not exactly on, the plane. Figure 5-26 concentrates on the triangle defined by the tips of the vectors yi,:, y ¯ i,: and ytruei,:. These are represented as small circles on that figure. The ¯. difference vector y ¯ i,:-yi,: is orthogonal to the plane spanned by V yi,: i,: noise removed

real noise

ytrue

i,:

noise left

yi ,:

plane v1,:, v2,:

Figure 5-26. Detailed view of Figure 5-25. ¯ is y The projection of yi,: into V ¯ i,:, which is usually is much closer to the true spectrum ytruei,: than yi,: itself. A substantial amount of the noise is removed in the projection but not all.

5.2 Target Factor Analyses, TFA We continue considering multivariate data sets, e.g. a series of spectra measured as a function of time, reagent addition etc. In short, a matrix of

247

Model-Free Analyses

data that can be decomposed in the usual way: Y=CA. The spectra are measured at nl wavelengths and thus they are nl-dimensional vectors. The whole series of spectra follow a particular path in an nl-dimensional space. We have recognised in the preceding Chapter 5.1 Factor Analysis, that this path is concentrated in a much lower dimensional sub-space. Usually, for an nc component system, the sub-space has nc dimensions; e.g. for a two component system, all spectra lie in a plane. Recall that, if the system is closed, the dimension of the sub-space can be further reduced by meancentring. To start with, we do not know the spectra A of the components in the system under investigation. Factor Analysis delivers an orthonormal system of axes ¯ that defines the sub-space of Y and A in an optimal way. Importantly, this V is done automatically, and there is no input from the chemist regarding the components in the system or their spectra. The basic idea of Target Factor Analysis is very simple. In order to test whether a certain compound is taking part in the process, whether its spectrum exists in the measurement, we test whether that spectrum lies in ¯ . If such a test spectrum is outside V ¯ , there is no doubt that the component V does not take part in the process under investigation. If it is in the sub­ space, we cannot positively conclude that the species is there; the test spectrum could be a linear combination of the existing spectra. A typical application can be found in chromatography. A group of components elute in a strongly overlapping peak cluster. We suspect that a particular chemical, for which we know the spectrum, might be in the unknown mixture, but due to overlap, its spectrum does not appear pure in the matrix Y. Due to inevitable experimental noise, the test spectrum vector will never be ¯ and consequently the question is whether the test exactly in the subspace V ¯. vector is close to V The initial idea might be to compute the distance r of the test row vector t ¯ . As indicated in Figure 5-27, r is the difference between t and its from V ¯. projection tproj into V

v2,: t1

t2 r2

r1

v1,:

tproj,2

¯ . While r1 is Figure 5-27. The distance of two test vectors to V shorter t2 is a better test vector.

248

Chapter 5

Figure 5-27 shows the principle for two test vectors t1 and t2. The fact that ¯ is shorter than the distance t2 does the distance r1 to the sub-space V not mean that t1 is a better candidate. The test vectors need to be normalised in order to be able to compare these distances. One could also use the angle between a test vector t and its projection tproj as a measure. t v2,:

r

tn rn

α

tn,proj tproj proj v1,:

Figure 5-28. The angle α is a good measure for closeness of the ¯. test vector t to the space V The angle α is defined by the following equations: sin α =

r t

=

rn tn

= rn

(5.23)

The projection tn,proj is computed as given in equation (5.26). The Matlab file Main_TFA.m generates a three-component overlapping chromatogram, generated by Data_Chrom2.m (p.219). Two test spectra t1 and t2 are generated, t1 is the original spectrum of one of the components, t2 is slightly shifted, see Figure 5-29. Both are normalised. The output includes the length of the residuals and the angles between the test spectra and the ¯. plane V MatlabFile 5-21. Main_TFA.m % Main_TFA  [t,lam,Y,C,A]=Data_Chrom2; ne=size(C,2); [U,S,Vt]=svd(Y,0);V_bar=Vt(:,1:ne)'; t1=1000*gauss(lam,450,120);            % component spectrum, max at 450nm t2=1000*gauss(lam,460,120);            % slightly shifted spectrum plot(lam,t1,lam,t2); xlabel('wavelength');ylabel('absorptivity');  t1n=t1/norm(t1);                       % normalisation of t1 and t2  t2n=t2/norm(t2); t1n_proj=t1n*V_bar'*V_bar;             % projections t2n_proj=t2n*V_bar'*V_bar; r1n=t1n­t1n_proj;                      % residuals 

249

Model-Free Analyses

r2n=t2n­t2n_proj;  distance(1)=norm(r1n); distance(2)=norm(r2n) angles=asin(distance)/pi*180           % angles in degrees 

distance =  0.0003    0.0424  angles = 0.0160    2.4287 

While both angles (given in degrees) and distances are small, the ones for the correct spectrum are significantly smaller. The principle of Target Factor Analysis is not restricted to the testing of spectra or, more generally, to row vectors. Exactly the same principles apply, of course, to column vectors or concentration profiles. In mathematical terms, there is a complete symmetry between the two. However, in chemical terms the two dimensions are different. Along the concentration profiles, we usually have a function that quantitatively describes the action while there is nothing of that kind along the spectral dimension. In Chapter 5.2.3, Target Transform Search/Fit, we take advantage of the functional definition that is available in the column space.

1000 900 800

absorptivity

700 600 500 400 300 200 100 0 400

450

500 wavelength

550

600

Figure 5-29. Correct (—) and slightly shifted (...) species spectrum, both used as target spectra.

250

Chapter 5

5.2.1 Projection Matrices The projection of a vector into the subspace defined by eigenvectors, and the subsequent calculation of the residual vector between the original and its projection, is a very common task. Refer back to equations (5.15) and (5.16). It is worthwhile investigating the computations in some detail. The determination of the projections can be regarded as a linear least¯ =, as in Figure squares fit; only now we have an orthogonal set of vectors V 5-28, rather than a general set of non-orthogonal vectors in F in the equivalent Figure 4-12. The projected test vector tproj is a linear combination ¯. of the vectors V tproj = b V

(5.24)

The computation of the linear parameters b is easy, as the pseudo-inverse of an orthonormal matrix is equal to its transposed

b = t V+ = t V t

and thus the projected vector

and the residuals

tproj = t V t V

(5.25)

(5.26)

r = t − tproj

= t − t Vt V

= t (I − V V )

(5.27)

t

The equivalent operations are valid for columns:

tproj = U b

= UU t t

with

(5.28)

r = t − tproj

= t − UU t t

= (I − UU ) t

(5.29)

t

The above equations are valid for orthonormal sets of basis vectors. They can be written in very similar ways for general non-orthogonal bases (e.g. F in Figure 4-12). The only difference is the computation of the pseudo-inverse, which can be numerically demanding, but is trivial for orthonormal bases.

Model-Free Analyses

251

tproj = b F #

r = t (I − F F )

(5.30)

+

Similarly, for a column vector tproj, we can write in accordance to Figure 4-11

tproj = F b #

r = t (I − F F )

(5.31)

+

¯ tV ¯ ), r=t(I-F+F), r=(I- U ¯U ¯ t)t and r=(I-FF+)t are While the notations r=t(I- V elegant, they are inefficient ways of performing the calculations. The ¯ tV ¯ and U ¯U ¯ t are often very large square matrices which take time to matrices V compute, store and also to multiply with the vectors t. It is faster to ¯ (U ¯ tt) rather than r=(I-U ¯U ¯ t)t. The same, of course, is valid for calculate r=t–U the equations (5.29)-(5.31).

5.2.2 Iterative Target Transform Factor Analysis, ITTFA As the name Iterative Target Transform Factor Analysis indicates, this is an iterative extension of Target Factor Analysis. This time, we apply Target Factor Analysis to column vectors or concentration profiles. The basic idea is straightforward. First, we somehow guess a concentration profile, preferably close to a true one. Call it ctest. In the 3-component chromatographic example Data_Chrom2a.m, we use a delta function (often called a needle in the context of ITTFA) with a maximum (52) close to the true maximum of the second concentration profile (50). ¯ . We iteratively Such a test vector ctest normally is not in the sub-space U ¯ , applying equation (5.28). improve it in the following way: project ctest into U ¯ , is not correct, e.g. it contains negative This projected vector, while lying in U elements. A correction is applied that makes the profile physically possible ¯ . Nevertheless, this new vector is a better estimate but it removes it from U than the original one. As the projection invariably results in a shortening of the vector ctest, we re-normalise it to a maximum of one in each iteration. Ideally, the iterations are continued until things are perfect. Unfortunately this is easier said than done, as convergence is notoriously slow. This is illustrated in Figure 5-30, the correct profile will essentially 'never' be reached. MatlabFile 5-22. Data_Chrom2a.m function [t,lam,Y,C,A]=Data_Chrom2a  lam=400:10:600; A(1,:)=1000*gauss(lam,450,120);  A(2,:)=2000*gauss(lam,350,120); A(3,:)=1000*gauss(lam,500,50); 

  % component spectra 

252

Chapter 5

t=(1:1:100)'; C(:,1)=1e­3*gauss(t,35,30);  C(:,2)=9e­4*gauss(t,50,31); C(:,3)=2e­3*gauss(t,70,32); 

  % elution profiles 

Y=C*A; randn('seed',0); Y=Y+1e­3*randn(size(Y));  MatlabFile 5-23. Main_ITTFA.m % Main_ITTFA  [t,lam,Y,C_sim,A_sim]=Data_Chrom2a; ne=size(C_sim,2); [U,S,Vt]=svd(Y,0); U_bar=U(:,1:ne);  c_sim_n=C_sim(:,2)/max(C_sim(:,2)); % true conc profile, normalised c_test=zeros(size(t)); c_test(52)=1;                       % init guess,delta function at t=52  for i=1:20; C_test(:,i)=c_test;              % improved test vectors in C_test c_new=U_bar*(U_bar'*c_test);     % projection into U c_test=c_new.*(c_new>=0);  % negative values=0 c_test=c_test/max(c_test);  % normalisation to max=1  end  plot(t,C_test,'­',t,c_sim_n,':'); xlabel('time');ylabel('norm. conc.'); 

1 0.9 0.8

norm. conc.

0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0

20

40

60

80

100

time

Figure 5-30. Progress of the ITTFA algorithm for one particular concentration profile. The dashed line represents the correct concentration profile.

Model-Free Analyses

253

We are using the function Data_Chrom2a for data generation. It produces slightly wider concentration profiles than Data_Chrom2 (p.219). Starting from the initial needle, we observe a very quick improvement towards the correct profile; however, the iterative process slows down very quickly and essentially never reaches the correct profile. This is a typical result as convergence in algorithms of this kind tends to be fast in the beginning and subsequently slows down dramatically. Defining a reliable termination criterion for the iterative process that copes with such behaviour is very difficult. This is the exact opposite of what we have experienced in the Newton-Gauss type algorithms of Chapter 4.3.1, where convergence accelerates towards the minimum.

5.2.3 Target Transform Search/Fit Traditional Target Factor Analysis just determines whether a particular test vector is close to the sub-space spanned by the significant eigenvectors, ¯ or U ¯ . As an introduction to the method of Target Transform either V Searching, consider the example discussed in Figure 5-29. One of the ¯ , the other does not. The almost obvious idea is to move the spectra lies in V ¯ spectrum along the wavelength axis, to continuously check the distance to V and determine the minimum. Moving spectra around, in this or similar ways, does not have many applications for the chemist. The reason is that there is no functional relationship that usefully defines absorption spectra. The general idea of target testing has more potential in the column direction. There, we deal with concentration profiles that often are defined by mathematical functions, based on chemical/physical laws. In the following, we develop the principle vaguely defined above and see how concentration profiles and their parameters can be determined by analysing and minimising distances. We develop the idea using a kinetic example. Any reaction scheme that consists exclusively of first order reactions, results in concentration profiles that are linear combinations of exponentials. There is no limit to the number of reacting components nc. The set of differential equations describing such a scheme of exclusively first order reactions can always be written in the following way. ⎤ ⎡ [c1 ] ⎤ ⎡ [c1 ] ⎤ ⎡k1,1 k1,2 " ⎥ ⎢ [c ] ⎥ ⎢k ⎥ ⎢ % ⎥ ⎢ [c 2 ] ⎥ ⎢ 2 ⎥ = ⎢ 2,1 ⎥⎢ # ⎥ ⎢ # ⎥ ⎢ # ⎥⎢ ⎥ ⎢ ⎥ ⎢ knc ,nc ⎥⎦ ⎣[cnc ]⎦ ⎣[cnc ]⎦ ⎢⎣

or

c = K c

(5.32)

(5.33)

254

Chapter 5

d[ X ] , the derivative of the concentration of X with Recall the notation[ X ] = dt respect to time. The vector c contains all the derivatives, the vector c all concentrations and the matrix K is formed by the rate constants describing the reaction mechanism. Usually, most entries in K are zero. It is best to use an example: k2 k1 → B ⎯⎯⎯ →C , equation (5.32) reduces to For the reaction A ⎯⎯⎯

⎡[ A ]⎤ ⎡ −k1 0 0 ⎤ ⎡[ A ]⎤ ⎢  ⎥ ⎢ ⎥⎢ ⎥ ⎢[B ]⎥ = ⎢ k1 −k2 0 ⎥ ⎢[B ]⎥ ⎢[C ]⎥ ⎢⎣ 0 k2 0 ⎥⎦ ⎢⎣[C ]⎦⎥ ⎣ ⎦

(5.34)

Equation (5.34) is the equivalent to equation (3.75)(d) in matrix notation. For the same set of components but including reversible reactions k2 k1 ⎯⎯⎯ → B ←⎯⎯ ⎯⎯⎯ → C , equation (5.32) becomes: A ←⎯⎯ ⎯ ⎯ k3

k4

⎡[ A ]⎤ ⎡ −k1 k3 ⎢  ⎥ ⎢ ⎢[B ]⎥ = ⎢ k1 −k2 − k3 ⎢[C ]⎥ ⎢⎣ 0 k2 ⎣ ⎦

0 ⎤ ⎡[ A ]⎤ k 4 ⎥⎥ ⎢⎢[B ]⎥⎥ −k 4 ⎥⎦ ⎢⎣[C ]⎥⎦

(5.35)

Such systems of differential equations are called homogeneous. They have as solutions, linear combinations of exponential functions, where the eigenvalues, λi, of the matrix K are the exponentials. In the first, irreversible example, equation (5.34), the eigenvalues of K are λ1=-k1, λ2=-k2 and λ3=0. Thus, the concentration profiles are linear combinations of the vectors e-λit, where t is the vector of times. In matrix notation we can write

C = E TE

(5.36)

where C contains the concentration profiles in the usual way, E contains the column vectors e:,i=e-λit and TE is a transformation matrix that establishes the relationship between C and E. The elements of TE are defined by the reaction scheme and the initial concentrations of the reacting species. In the above example, equation (5.34), it is relatively straightforward to determine the eigenvalues of K. In the example (5.35) it is much more difficult to develop the equations. The Symbolic Toolbox of Matlab can be employed for the task. MatlabFile 5-24. Main_Sym_ABC.m % Main_Sym_ABC  syms k1 k2 K 

% symbolic variables 

K=[­k1   0    0;      % A­>B­>C  k1  ­k2   0; 0    k2   0]; 

Model-Free Analyses

lambdas=eig(K) 

255

% eigenvalues of K 

lambdas =  ­k1  ­k2     0   

and for the more interesting case of example (5.35) MatlabFile 5-25. Main_Sym_ABC_rev.m % Main_Sym_ABC_rev  syms k1 k2 k3 k4 K 

% symbolic variables 

K=[­k1   k3      0;     % ABC  k1  ­k2­k3   k4; 0    k2     ­k4];  lambdas=eig(K) 

      % eigenvalues of K 

lambdas =  0  ­1/2*k1­1/2*k2­1/2*k3­1/2*k4+1/2*(k1^2­2*k1*k2+2*k1*k32*k1*k4+k2^2+2*k2*k3+2*k4*k2+k3^2­2*k3*k4+k4^2)^(1/2) ­1/2*k1­1/2*k2­1/2*k3­1/2*k4­1/2*(k1^2­2*k1*k2+2*k1*k32*k1*k4+k2^2+2*k2*k3+2*k4*k2+k3^2­2*k3*k4+k4^2)^(1/2) 

or in a more civilised form λ1 = 0

(5.37) 1 2 1 λ2 = − (k1 + k2 + k3 + k4 ) + k1 + k22 + k32 + k42 − 2k1k2 + 2k1k3 − 2k1k4 + 2k2k3 + 2k2k4 − 2k3k4 2 2 1 1 2 λ3 = − (k1 + k2 + k3 + k4 ) − k1 + k22 + k32 + k42 − 2k1k2 + 2k1k3 − 2k1k4 + 2k2k3 + 2k2k4 − 2k3k4 2 2 Interestingly, analysis of measured data only delivers the 2 (or 3 if zero is included) λ-values. There is not enough information to resolve two equations into 4 rate constants. Or, in chemical terms, without independent additional information, it is impossible to determine all 4 rate constants. The Symbolic Toolbox can even cope with initial concentrations and thus delivers the equations for the concentration profiles. MatlabFile 5-26. Main_Sym_ABC_rev.m …continued % Main_Sym_ABC_rev, ...continued  C=dsolve('Da=­k1*a+k3*b','Db=k1*a­(k2+k3)*b+k4*c', ... 'Dc=k2*b­k4*c','a(0)=A0','b(0)=0','c(0)=0');  C.a  C.b  C.c 

The output of this short program is 9500 characters, too much to be included here. We leave it to the readers to perform the task on their computer.

256

Chapter 5

Back to Target Factor Analysis. C is both a linear combination of E (equation ¯ (see later in equation (5.49)). Combining the two (5.36)) and also of U equations

C = E TE

E=

C TE-1

and C = U TU

= U TU TE-1 = U T

(5.38)

¯ . Thus, demonstrates that the columns e:,i of E are linear combinations of U ¯ if the correct eigenvalue λi is used; otherwise it is at a e:,i lies in or close to U distance. This is nothing but target testing a test vector etest=e-λtestt. The application Main_TTF.m uses data generated in Data_ABC2.m for a consecutive reaction A→B→C. MatlabFile 5-27. Data_ABC2.m function [t,lam,Y,C,A]=Data_ABC2  % A ­> B ­> C  t   = [0:100]';     % reaction times  lam = 400:5:600;    % wavelengths k   = [0.1 0.03];   % rate constants  A_0 = 1e­3;         % initial concentration of A  C(:,1)=A_0*exp(­k(1)*t);       % concentrations of species A C(:,2)=A_0*(k(1)/(k(2)­k(1))*(exp(­k(1)*t)­exp(­k(2)*t))); % conc. of B C(:,3)=A_0­C(:,1)­C(:,2);      % concentrations of C  A(1,:)=1e3*gauss(lam,450,120); % molar spectrum of species A A(2,:)=2e3*gauss(lam,350,120)+1e3*gauss(lam,500,50); % mol. spect. of B A(3,:)=1e3*gauss(lam,500,50);  % molar spectrum of C  Y=C*A;                         % Beer's law  randn('seed',0);               % fixed start for random number generator Y=Y+0.001*randn(size(Y));      % standard deviation 0.001  MatlabFile 5-28. Main_TTF.m % Main_TTF  [t,lam,Y,C,A]=Data_ABC2; ne=size(C,2);  [U,S,Vt]=svd(Y,0); U_bar=U(:,1:ne);  lambda_test=­.05:.001:.2; for i=1:length(lambda_test) e_test=exp(­lambda_test(i)*t);   % test exponential vector e_test=e_test/norm(e_test);      % normalise  r=e_test­U_bar*(U_bar'*e_test);  % residual vector  distance(i)=norm(r);             % distance  end  plot(lambda_test,log10(distance)); xlabel('lambda_{test}');ylabel('log(\midr\mid)'); 

257

Model-Free Analyses

0 -0.5

log(⏐r⏐)

-1 -1.5 -2 -2.5 -3 -3.5 -0.05

0

0.05 0.1 lambda

0.15

0.2

test

Figure 5-31. Distance of normalised exponential functions, exp(­ ¯. λtest*t) to the subspace U Figure 5-31 clearly features three minima at the correct positions, λtest=0, and the two rate constants used to generate the data: λtest=0.03 and λtest=0.1. A very interesting feature of the whole method is that the rate constants are completely independent. Each minimum, or rate constant, is defined on its own, completely independent of all the others. This is in clear contrast to normal, hard-modelling data fitting where the residuals are a function of all parameters together. There are several extensions and comments worth making. 1. It is possible to use an iterative algorithm to determine the exact positions of the minima. Again, in such a program the rate constants can be fitted individually, irrespective of the others. 2. Under certain circumstances it is possible to represent concentration profiles encountered in titrations as linear combinations of the typical S-shaped profiles known from equilibrium studies. 3. It is feasible to target test not only column vectors of C or E but also the complete matrices C or E. Parameter Fitting via Target Testing

It is worthwhile examining point 3 above in some additional detail. Equation (5.38) C = U TU is, of course, not completely correct. It is only an approximation; it should be written as

258

Chapter 5

C = U TU + R U

(5.39)

The matrix C is defined by the non-linear parameters (rate constants). It is possible to minimise RU, i.e. the corresponding ssq, as a function of these parameters in a 'normal' Newton-Gauss algorithm. The chain of equations goes as follows TU = U t C

R U = C − U TU = C − UU t C = (I − UU ) C

(5.40)

t

The advantage is that there is no pseudo-inverse to be calculated in this way. The computation of TU, which comprises linear parameters is easier ¯ is an orthonormal matrix, U ¯ +=U ¯ t. As mentioned before, in than 'usual' as U equation (5.29), it is advantageous to compute the residuals as R U = C − U(U t C) ; it is considerably faster. The 'standard' chain of equations using Beer-Lambert's law is:

Y = CA +R

A = C+ Y

R = Y − CC+ Y

(5.41)

= (I − CC + ) Y

The main difference to equations (5.40) is the computation of the pseudoinverse C+. For the sake of completeness, we also include the relevant equations if data reduction, according to Reduced Eigenvector Space (p.180), is applied: Yred = C A red + R red = US

A red = C + US

R red = US − CC+ US

(5.42)

= (I − CC + ) US

The dimensions of the corresponding matrices in equation (5.42) are all the same as in equation (5.40), the main difference is still the computation of C+ ¯ t. instead of U The two approaches, equation (5.40) or (5.41)/(5.42), are not equivalent. Figure 5-32 attempts to represent the situation graphically. Due to the limitation of our mind to 3 dimensions, this endeavour is not easy, or rather ¯¯ or Y it is impossible as we are running out of dimensions. The 'vectors' US ¯¯ or the complete space Y; the curved line C=f(k) represent the subspace US represents the space defined by the whole matrix C; the 'vector' C(kC) represents the minimum for equation (5.41) or (5.42) and C(kU) the minimum

259

Model-Free Analyses

for equation (5.40); the 'vectors' R (Rred) and RU are the corresponding residual matrices.

R or Rred

C(kc) C(ku)

Ru

US or Y

C=f(k) Figure 5-32. Graphical representation of equations (5.40) to (5.42). The 'normal' residuals R are orthogonal to the space C, defined by the ¯¯ or Y into C. This is a straightforward projection of the column vectors in US linear least-squares calculation, equivalent to Figure 4-10. C(kC) is the ¯¯ or Y. closest the space C gets to the 'vectors' US The residuals Ru are defined by the projections of the 'vectors' C into the ¯ ; they are orthogonal to U ¯ . This projection is simpler due to the space U ¯. orthogonal base vectors. RU is the closest the 'vectors' C get to U The figure, however incomplete, demonstrates that the two minima are not the same. Often they are very similar. Probably a more important difference ¯¯ have a length defined by the measurement, thus there is that the vectors US is a weight given to the vectors which is relevant. Such information is completely lost in Target Transform Fitting, employing equation (5.40); Ru is ¯¯ . In fact, the shorter C, minimised without any reference to the length of US the shorter is Ru, thus some normalising of C, as a function of the parameters k, is required. Refer also to Figure 5-27 for a similar situation.

5.3 Evolving Factor Analyses, EFA ¯¯¯ is full The Singular Value Decomposition of a matrix Y into the product USV of rich and powerful information. The model-free analyses we discussed so ¯ and V ¯. far are based on the examination of the matrices of eigenvectors U Evolving Factor Analysis, EFA, is primarily based on the analysis of the ¯ of singular values. matrix S Previously, we have seen in Magnitude of the Singular Values (p.219) that the number of significant singular values in S equals the number of linearly

260

Chapter 5

independent rows or columns of the matrix Y of measurement, which ideally equals the number of changing chemical components in the process investigated. So far, the complete data matrix Y has been analysed and thus the result reflects the total measurement, the number of components existing anywhere in the measurement. Evolving Factor Analyses investigate the evolving character of the singular values − how they change as a function of the progress of the measurement. Information about the evolution of the rank and thus the appearance of new components is revealed. Naturally, this only makes sense if there is an inherent order in the data, usually an order in the acquisition of the spectra that make up the matrix Y. Factor analytical methods have been developed by social scientists; their samples are individuals for whom they have a 'spectrum' of properties. In this collection of samples there is no inherent order and thus, methods that rely on an inherent order of the samples, such as EFA, are of no use. As a typical example of ordered data we will investigate chromatography, where spectra are measured as a function of elution time.

5.3.1 Evolving Factor Analysis, Classical EFA The basic principle of EFA is very simple. Instead of subjecting the complete matrix Y to the Singular Value Decomposition, specific sub-matrices of Y are analysed. In the original EFA, these sub-matrices are formed by the first i spectra of Y where i increases from 1 to the total number of spectra, ns. The appearance of a new compound during the acquisition of the data is indicated by the emergence of a new significant singular value. The procedure is best explained graphically in Figure 5-33. The sub-matrix, indicated in grey, is subject to the SVD and the resulting ne significant singular values are stored as a row vector in a matrix EFA of the same number of rows as Y.

ne

1 SVD i

ns Figure 5-33. Schematic of Forward EFA

si ,:

Model-Free Analyses

261

The example used for the introduction of EFA is based on the threecomponent chromatogram, Data_Chrom2.m (p.219) we have used several times earlier. While most of the Matlab listing in Main_EFA1.m is close to self explanatory, a few statements might need clarification. The singular values are stored in the matrix EFA_f which has ns rows and ne columns. It is advantageous to plot the logarithms of the singular values; their values span several orders of magnitude and cannot be represented in a normal plot. The rank is the number of significant singular values. The significance level can be estimated as the first non-significant singular value of the total matrix Y. The three sub-plots in Figure 5-34 clearly indicate the relationship between the concentration profiles, the evolving singular values and the evolving rank.

MatlabFile 5-29. Main_EFA1.m % Main_EFA1  [t,lam,Y,C,A]=Data_Chrom2; [ns,nc]=size(C); ne=nc+1;                        % one extra singular value  EFA_f=NaN(ns,ne);               % NaNs to prevent log(0) for i=1:ns  s_f=svd(Y(1:i,:));           % svd of the first i rows of Y  if isig_level)'); % number of significant SV subplot(3,1,1); plot(t,C); ylabel('conc.'); subplot(3,1,2); plot(t,log10(EFA_f)); ylabel('log(s_f)'); subplot(3,1,3); plot(t,Rank_f,'x'); axis([0 t(ns) 0 ne]); xlabel('time');ylabel('rank'); 

262

Chapter 5

conc.

2

x 10

-3

1

0 0

20

40

60

80

100

20

40

60

80

100

20

40

60

80

100

log(s f)

2 0 -2 -4 0

rank

4

2

0 0

time

Figure 5-34. Forward EFA. Top panel: concentration profiles; second panel: evolving singular values; third panel: evolving rank. Evolving factor analysis can, and should be performed in both forward and backward directions. The forward plot, calculated above and shown in Figure 5-34, indicates the appearance of new components. The backward plots of Figure 5-36 are calculated similarly by determination of the singular values of the set of the last 1, 2, 3, ... spectra in Y, as seen in the schematic of Figure 5-35. These plots indicate the disappearance of the components.

ne 1 ns-i+1 SVD ns Figure 5-35. Schematic of Backward EFA

sns −i +1,: +1,:

263

Model-Free Analyses MatlabFile 5-30. Main_EFA1.m …continued % Main_EFA1 ...continued  EFA_b=NaN(ns,ne);  for i=1:ns  s_b=svd(Y(ns­i+1:ns,:));  if isig_level)');  subplot(3,1,1); plot(t,C); ylabel('conc.'); subplot(3,1,2); plot(t,log10(EFA_b)); ylabel('log(s_b)'); subplot(3,1,3); plot(t,Rank_b,'x'); axis([0 t(ns) 0 ne]); xlabel('time');ylabel('rank'); 

% number of species

conc.

2

x 10

% svd of the last i rows of Y  % relevant SV are stored 

-3

1

0 0

20

40

60

80

100

20

40

60

80

100

20

40

60

80

100

log(sb)

2 0 -2 -4 0

rank

4

2

0 0

time

Figure 5-36. Backward EFA. Top panel: concentration profiles; second panel: evolving singular values; bottom panel: evolving rank.

264

Chapter 5

The combined interpretation of the plots of the singular values in forward and backward direction is fairly straightforward. The increase of the rank by one, indicates the appearance of a species during the process monitored. In well behaved chromatograms, i.e. non-overloaded columns, the width of the elution profiles increases continuously with increasing elution time. In such instances, it is possible to connect the appearance and the disappearance of the individual components: the first compound to appear is also the first to disappear, etc. Concentration windows can be established for all compounds. These are regions along the time axis during which a component exists. Outside these windows the concentration is known to be zero. The connection between the forward and backward singular values can be made in a one-line Matlab command MatlabFile 5-31. Main_EFA1.m …continued % Main_EFA1 ...continued  for i=1:3                   % windows of existence for the components C_window(:,i)=EFA_f(:,i)>sig_level & EFA_b(:,ne­i)>sig_level; end  subplot(4,1,1) plot(t,C); ylabel('conc.'); subplot(4,1,2); plot(t,log10(EFA_f)); ylabel('log(s_f)'); subplot(4,1,3); plot(t,log10(EFA_b)); ylabel('log(s_b)'); subplot(4,1,4); plot(t,C_window(:,1),t,C_window(:,2)+0.3,t,C_window(:,3)+0.6); xlabel('time');ylabel('conc. window'); 

EFA plots can be used to estimate the rank of a matrix. EFA plots have similarities with the singular value plots shown in Figure 5-4 but they clearly contain more information and thus are more instructive. In order to demonstrate this enhanced capability of EFA plots for the determination of the number of components, we generate a series of spectrophotometric titrations of a diprotic acid with different noise levels, employing Data_eqAH2a.m (p.236) and analyse with Main_EFA2.m. Forward EFA plots, next to the original data, are presented in Figure 5-38. A few observations can be made: the significant singular values are not much affected by the noise level, only the non-significant ones move up continuously with increasing noise. This behaviour is similar to the one observed in Figure 5-4. The lowest panels of Figure 5-38 demonstrate that even at very high noise levels, EFA facilitates the determination of the correct number of components.

265

Model-Free Analyses

conc.

2

x 10

1

log(sf)

0 0 5

log(sb)

20

40

60

80

100

20

40

60

80

100

20

40

60

80

100

20

40

60

80

100

0 -5 0 5

conc. window

-3

0 -5 0 2 1 0 0

time

Figure 5-37. Complete EFA. Concentration profiles; forward and backward evolving singular values; bottom panel: concentration windows. The human eye is very good at detecting patterns − in this case the appearance of a new significant singular value. The appearance of a new component, as indicated by the point where a new significant singular value rises above the noise level, is delayed by increasing noise. MatlabFile 5-32. Main_EFA2.m % Main_EFA2  [pH,lam,Y,C,A]=Data_eqAH2a; [ns,nc]=size(C); ne=nc+1;                            % one extra singular value noise=[.05 .1 .2];  for j=1:3; Yn=Y+noise(j)*randn(size(Y));    % add different noise levels  EFA_f=EFA(Yn,ne); subplot(3,2,2*j­1); plot(pH,Yn,'­'); axis([2 12 ­1 2]); if j==3,xlabel('pH');end;ylabel('abs.'); subplot(3,2,2*j); 

266

Chapter 5

plot(pH,log10(EFA_f)); axis([2 12 ­.5 1.5]); if j==3,xlabel('pH');end;ylabel('log(s_f)'); end 

1.5 1

1

log(sf)

abs.

2

0 -1 2

0 4

6

-0.5 2

8 10 12

0

8 10 12

4

6

8 10 12

4

6 8 10 12 pH

0.5 0

4

6

-0.5 2

8 10 12

2

1.5 1

1

log(s f)

abs.

6

1

1

0 -1 2

4

1.5 log(sf)

abs.

2

-1 2

0.5

0.5 0

4

6 8 10 12 pH

-0.5 2

Figure 5-38. EFA forward plots for a data set with increasing noise levels. EFA.m is a short Matlab function that computes forward and backward EFA matrices for a given number, ne, of singular values. Its structure is essentially identical to the one discussed for Main_EFA2.m. MatlabFile 5-33. EFA.m function [EFA_f,EFA_b]=EFA(Y,ne)  [ns,nl]=size(Y); EFA_f=NaN(ns,ne); EFA_b=NaN(ns,ne);  for i=1:ns  s_f=svd(Y(1:i,:));                % forward SV  s_b=svd(Y(ns­i+1:ns,:));          % backward SV  EFA_f(i,1:min(i,ne))=s_f(1:min(i,ne))'; EFA_b(ns­i+1,1:min(i,ne))=s_b(1:min(i,ne))'; end 

267

Model-Free Analyses

Interestingly, EFA was originally developed for the analysis of spectrophotometric titration data. Concentration profiles in chromatography and equilibrium studies can be surprisingly similar. The main difference is that in chromatography, the data set generally starts and ends without any component present (Figure 5-37), while in titrations, there is usually one particular species at the beginning and another one at the end (Figure 5-39). While the algorithm is not affected, the concentration windows are different. MatlabFile 5-34. Main_EFA3.m % Main_EFA3  [pH,lam,Y,C,A]=Data_eqAH2a; [ns,nc]=size(C); ne=nc+1; [EFA_f,EFA_b]=EFA(Y,ne);  sig_level=EFA_f(ns,ne); for i=1:nc                    % windows of existence for the components C_window(:,i)=EFA_f(:,i)>sig_level & EFA_b(:,ne­i)>sig_level; end  subplot(3,1,1); plot(pH,C); ylabel('conc.'); subplot(3,1,2); plot(pH,log10(EFA_f),'­',pH,log10(EFA_b),':'); ylabel('log(s_f),log(s_b)'); subplot(3,1,3); plot(pH,C_window(:,1),pH,C_window(:,2)+0.3,pH,C_window(:,3)+0.6); xlabel('pH');ylabel('conc. window'); 

conc.

1

x 10

-3

0.5

0 2

4

6

8

10

12

4

6

8

10

12

4

6

8

10

12

b

log(s ),log(s )

2

f

0 -2 -4 2 conc. window

2

1

0 2

pH

Figure 5-39. Concentration profiles for the titration of a di-protic acid; EFA plots and concentration windows.

268

Chapter 5

5.3.2 Fixed-Size Window EFA, FSW-EFA There are different strategies for the selection of sub-matrices for evolving type factor analyses. The classical, original mode has been presented so far. The most important alternative procedure is based on a moving window of fixed size. In other words, a window of a pre-defined number of consecutive spectra is moved along the columns of the matrix Y. Each window is subjected to SVD, the singular values are stored and their logarithms are plotted.

SVD

si ,:

Figure 5-40. Schematic of FSW-EFA Fixed-size-window-EFA plots reveal the number of different species that co­ exist in the particular window. More precisely, it is the number of species with linearly independent concentration profiles. Here is the appropriate Matlab program Main_FSW_EFA.m. The data, generated by Data_eqAH4a.m, are mimicking a spectrophotometric titration of a tetra-protic acid AH4. with log(K) values of 8, 7, 6 and 2. The equilibria are quantitatively described by equation (5.43). 1 ⎯⎯⎯⎯ → AH A + H ←⎯⎯⎯⎯

log K

2 ⎯⎯⎯⎯→ AH + H ←⎯⎯⎯⎯ AH 2

log K

3 ⎯⎯⎯⎯→ AH 3 AH 2 + H ←⎯⎯⎯⎯

log K

(5.43)

4 ⎯⎯⎯⎯→ AH 3 + H ←⎯⎯⎯⎯ AH 4

log K

The concentrations of the differently protonated species as a function of pH are calculated with the explicit function we developed in Special Case: Explicit Calculation for Polyprotic Acids, p.64. MatlabFile 5-35. Data_eqAH4a.m function [pH,lam,Y,C,A]=Data_eqAH4a  pH=[0:.1:12]';                   % pH range H=10.^(­pH); 

Model-Free Analyses

logK=[8 7 6 2];                  % protonation constants K=10.^logK; n=length(logK);                  % number of protons  denom=zeros(size(H)); for i=0:n  num(:,i+1)=H.^i*prod(K(1:i)); % numerator denom=denom+num(:,i+1);       % denominator  end  alpha=diag(1./denom)*num;        % degree of dissociation C=1e­3*alpha;                    % concentration profiles  lam=400:10:600;                  % wavelength range A(1,:)=1000*gauss(lam,450,120);  % component spectra A(2,:)=2000*gauss(lam,350,120); A(3,:)=1000*gauss(lam,500,50); A(4,:)=1000*gauss(lam,550,50); A(5,:)=1000*gauss(lam,580,50); Y=C*A;                           % absorbance data  randn('seed',0); Y=Y+1e­3*randn(size(Y));         % noise level 0.001  MatlabFile 5-36. Main_FSW_EFA.m % Main_FSW_EFA  [pH,lam,Y,C,A]=Data_eqAH4a;  [ns,nc]=size(C); 

% eq. data 4­protic acid

size_w=3;    % small windows  EFA_w=zeros(ns­size_w+1,size_w); for i=1:ns­size_w+1  s_w=svd(Y(i:i+size_w­1,:)); EFA_w(i,:)=s_w'; end  subplot(3,1,1); plot(pH,C,'k'); ylabel('conc'); subplot(3,1,2); plot(pH(0.5*(size_w+1):ns­0.5*(size_w­1)),log10(EFA_w),'k'); ylabel('log(s_w)');  size_w=9;    % large windows  EFA_w=zeros(ns­size_w+1,size_w); for i=1:ns­size_w+1  s_w=svd(Y(i:i+size_w­1,:)); EFA_w(i,:)=s_w'; end  subplot(3,1,3); plot(pH(0.5*(size_w+1):ns­0.5*(size_w­1)),log10(EFA_w),'­k'); xlabel('pH'); ylabel('log(s_w)'); 

269

270

Chapter 5

conc

1

x 10

-3

0.5

0 0

2

4

6

8

10

12

2

4

6

8

10

12

2

4

6 pH

8

10

12

lo g (s w )

2 0 -2 -4 0

lo g (s w )

2 0 -2 -4 0

Figure 5-41. Concentration profiles for a titration of a 4-protic acid. Second panel: FSW-EFA plot for a window size of 3; and third panel for a window size of 9. Figure 5-41 displays the results of two FSW-EFA analyses with different window sizes, 3 and 9, the second and third panels. The small window of size 3 naturally cannot detect more than 3 different components within the window. The relatively high noise level using a window of only 3 spectra, and the high overlap of the concentration profiles, makes the third singular value in the middle plot, expected between pH 6-8, hardly discernable. With windows of size 9, up to 4 singular values are clearly identifiable. The price to pay for large window sizes is the spreading out of the information. Around pH 4, there are only 2 components co-existing but within the large window there is a total of 3 components. Due to that broadening effect, the

Model-Free Analyses

271

beginnings of concentration profiles are not easily detected. However, the 4 components coexisting at pH 7 are well distinguished in the FSW-EFA plot. There are advantages and disadvantages in this approach compared to classical EFA. a) In big systems with many species and spectra, the detection of new species deteriorates with increasing window size. This effect is clearly noise dependent and can qualitatively be observed in Figure 5-38. It is the effect of a continuous increase of the noise singular values. In FSW-EFA, the fixed size window maintains the magnitude of the singular values related to noise. b) the classical EFA plots are easier to interpret − compare Figure 5-37 with Figure 5-41. c) Consider the following, rather unlikely example: component 1 and 5 in a co-eluting peak system in chromatography have the same spectrum. Under such circumstances the detection of component 5 in the cluster is not detected at all in classical EFA. Window EFA does not suffer from these shortcomings as long as there is no overlap between the 2 components with identical spectra. d) a decision has to be made for the size of the window in FSW-EFA. As outlined above, this decision is important.

5.3.3 Secondary Analyses Based on Window Information The location of the concentration windows is the distinctive result of the classical EFA plots, as in Figure 5-37. Apart from information on peak purity in chromatography, there is not much that is directly useful in the information about concentration windows. In this section, we develop methods that, based on these concentration windows, result in complete concentration profiles C and subsequently the corresponding species spectra A. Iterative Refinement of the Concentration Profiles

This algorithm has many aspects similar to Iterative Target Transform Factor Analysis, ITTFA, as discussed in Chapter 5.2.2, and Alternating LeastSquares, ALS as introduced later in Chapter 5.4. The main difference is the inclusion of the window information as provided by the EFA plots. A brief description of the algorithm (as usual, everything is based on Y=CA): (a) Initial guess for the matrix C of concentrations; often this is not crucial; possible choices are combined EFA plots (see Figure 5-44), the window matrix of 0's and 1's is adequate as well (Figure 5-37). (b) Calculate A as A=C\Y (c) Corrections on A, e.g. negative values=0 (d) Calculate C as C=Y/A (e) Corrections on C, negative values and values outside the concentration windows =0. (This is the main difference to ITTFA).

272

Chapter 5

(f) If fit not 'perfect' return to (b). Instead of a proper termination criterion, which is difficult to develop, we just iterate 100 times. We employ the same chromatographic data data_chrom2a.m (p.251) as in ITTFA. Figure 5-42 displays the results.

MatlabFile 5-37. Main_It_EFA.m % Main_It_EFA  [t,lam,Y,C_sim,A_sim]=Data_Chrom2a; [ns,nc]=size(C_sim); ne=nc+1;                    % one extra singular value  [EFA_f,EFA_b]=EFA(Y,ne);    % perform EFA EFA_f(isnan(EFA_f)==1)=0;   % replace NaN's by zeros EFA_b(isnan(EFA_b)==1)=0;   % replace NaN's by zeros C=min(EFA_f(:,1:nc),fliplr(EFA_b(:,1:nc))); % combined SV curves sig_level=0.01;             % define cut off level  C_window=C>sig_level;       % build window matrix of 0's and 1's  for it=1:100  C=C/diag(max(C));        % normalization to max of 1  A=C\Y;                   % spectra A=A.*(A>0);              % positive  C=Y/A;                   % conc profiles C=C.*(C>0);              % positive C=C.*C_window;           % apply windows  R=Y­C*A;                 % residuals  ssq(it)=sum(sum(R.*R));  end  [C_n,A_n]=norm_max(C,A);    % norm. C and C_sim to unit height [C_sim_n,A_sim_n]=norm_max(C_sim,A_sim); % and recalc. A, A_sim subplot(3,1,1); plot(lam,A_n,'­',lam,A_sim_n,'.'); xlabel('wavelength');ylabel('absorptivity'); subplot(3,1,2); plot(t,C_n,'­',t,C_sim_n,'.'); xlabel('time');ylabel('concentration'); subplot(3,1,3); plot(log10(ssq)); xlabel('iteration');ylabel('log(ssq)'); 

273

Model-Free Analyses

absorptivity

2

1

0 400

450

500 wavelength

550

600

concentration

1

0.5

0 0

20

40

60

80

100

60

80

100

time

log(ssq)

2 0 -2 -4 0

20

40 iteration

Figure 5-42. Iterative EFA. Top panel: component spectra; middle panel: concentration profiles; bottom panel: progress of the quality of the fit. Markers (•) represent the true values and the lines, the iterative EFA estimates after 100 iterations. The width of the concentration windows is defined by the chosen level of significance of the singular values. Figure 5-43 represents the effect of choosing two different levels: the dotted horizontal line is defined by the first non-significant singular value (fine dotted line) of the complete matrix Y. The intersection of this horizontal line with the EFA trace of the second singular value is the beginning of the second concentration window. The lower significance level at 0.01, represented by the full line, results in an intersection at an earlier time and thus a wider concentration window. In all future iterative analyses of these chromatography data, data_chrom2a.m, the level of significance is set to 0.01.

274

Chapter 5

-1

log10(S)

-1.5

-2

-2.5 0

20

40

60

80

100

time

Figure 5-43. Different significance levels defining the concentration windows for Data_Chrom2a. The initial guess for the concentration profiles is computed as the combination of the forward and backward EFA graphs; the smaller of each forward/backward pair is used. We display them in Figure 5-44 as we use these initial guesses in most other upcoming model-free analyses. C=min(EFA_f(:,1:nc),fliplr(EFA_b(:,1:nc))); % combined SV curves  plot(t,C); xlabel('time');ylabel('concentration'); 

3.5 3

concentration

2.5 2 1.5 1 0.5 0 0

20

40

60

80

100

time

Figure 5-44. Initial guesses for the concentration profiles, computed as the combination of the singular value traces for forward and backward EFA.

Model-Free Analyses

275

Several additional comments are due. As observed in Chapter 5.2.2, Iterative Target Transform Factor Analysis, ITTFA, iterative progress is relatively fast at the beginning and slows down continuously with the number of iterations. The third panel of Figure 5-42 demonstrates that the minimum has not been reached at all after 100 iterations. While the concentration profiles are reasonably well reproduced, there are some problems with the absorption spectra; one spectrum has a substantial contribution from another. Nevertheless, considering the simplicity of the algorithm the results are astoundingly accurate. Model-free methods do neither supply absolute information about the concentrations or about the spectra. Essentially they only deliver the shapes for the profiles. In this and future examples, we normalise the concentration profiles in C to a maximum of 1 and adjust the species spectra of A in such a way that the product CA is correct. This is done in the function norm_max.m. MatlabFile 5-38. norm_max.m function [Cn,An]=norm_max(C,A)  coef=1./max(C);      % normalisation coefficients  Cn=C*diag(coef);      % apply to C  if nargin==2 An=diag(1./coef)*A;  % apply inverse to A end 

It is worthwhile to compare this iterative refinement of concentration profiles as given on p.271 with ITTFA, the other iterative process we introduced in Chapter 5.2.2. ¯S ¯V ¯ , see Instead of computing A=C+Y, as in (b) on p.271, we replace Y with U equation (5.10).

Y = CA = USV

(5.44)

The component spectra A can then be determined by the following ¯ tC)-1U ¯ t from the left results in rearrangements. Multiplication with (U (U t C )−1 U t C A = (U t C )−1 U t U S V or

A = (U C ) S V t

−1

(5.45)

The next step is C=YA+, as in (c) on p.271. Applying equation (5.45) to compute the pseudo-inverse of A

C = Y A + = U S V V t S-1 U t C = U U t C

(5.46)

which is exactly the formula used in ITTFA. If no changes are applied to C and A during both iterative processes, there is no difference between the two methods. The advantage in the refinement, as outlined in steps (a)-(f) on

276

Chapter 5

p.271, lies in the possibility of the incorporation of extra information on A, such as the non-negativity constraint. Explicit Computation of the Concentration Profiles

As we have seen with the previous iterative refinement and ITTFA, convergence generally is very sluggish. Even with moderately complex systems, it is often too slow to be useful. There are alternative, non-iterative methods that compare favourably with the above iterative algorithms. ¯¯¯ which We start the derivation with the standard equations Y=CA and Y=USV t −1 t we combine, see equation (5.44). Post-multiplication with V (A V ) results in

C A V t (A V t )−1 = C = U S (A V t )−1

(5.47)

To compute C, the only unknown is A. It is advantageous to regard the product S(A )−1 as the unknown. Dimensional analysis shows that it is a nc×nc square matrix (nc =number of components), we call it a transformation matrix T:

T = S (A V t )−1

(5.48)

C = UT

(5.49)

And now

This is a very interesting and useful equation and we will return to it several times in later parts of this chapter. Equation (5.49) relates the concentration ¯ of eigenvectors. It is worthwhile representing the matrix C to the matrix U equation graphically.

nc

nc

nc ×

ns

C

=

T

nc

U

Figure 5-45. Graphical representation of C = U T

Model-Free Analyses

277

For an nc=2 component system the square transformation matrix T has only 4 elements, for a 3 component system there are 9 elements, etc. This relatively small number of unknown elements can be calculated explicitly! The crucial information is contained in the concentration windows, as determined by EFA. We know the elements of the matrix C within the concentration windows are positive, while outside these windows, the ¯ . The known elements of C are zero. We also know the complete matrix U ¯ are represented as the shaded areas in Figure 5-46. The elements of C and U white parts of the matrices have to be calculated. ×

=

=

C

U

×

T

¯ are known. Figure 5-46. The shaded parts of C and U ¯T The idea is to compute the elements of T in such a way that the product U results in zeros for the shaded part of C. In order to achieve this, we can separate the columns of C and treat them individually. The i-th column c:,i of ¯ ×t:,i, where t:,i is the i-th column of T. C is the product of U c:,i = U × t:,i

(5.50)

Graphically: ×

=

c:,i

=

U

×

t:,i

¯ ×t:,i. Figure 5-47. The i-th column c:,i of C is the product U

278

Chapter 5

This equation can be split up again. We can remove the unknown, white part of c:,i, i.e. the window of existence of the i-th component, and also the ¯ (in white). What is left, in grey, is the product corresponding part (rows) in U

c:,0i = U0 × t:,i = 0

(5.51)

where the superscript '0' represents the zero parts of c:,i and corresponding ¯. parts of U × =

0

=

U0

× t:,i

Figure 5-48. The homogeneous system of equations U0 × t:,i = 0 This represents a homogeneous system of equations. There is, as always with homogeneous equations, a trivial solution: t:,i=0. There are, however, and fortunately, non-trivial solutions as well. This is due to the fact that U0 does not have full rank. We removed all the information on the i-th component and consequently the rank of U0 is one less than the ¯ . In such a system of equations, a solution t:,i is not rank of the complete U completely defined, as it can be multiplied by any factor (≠0). Thus, we can freely choose one element, e.g. the first element in t:,i as one, t1,i=1. Equation (5.51) can now be written as

0 = U0 × t:,i

0 0 ×1 + U:,2: = u:,1 nc × t2:nc ,i

(5.52)

× =

0

+

0 u:,1 :,1

t2:nc ,i

0 U:,2: nc

Figure 5-49. Graphical representation of equation (5.52).

Model-Free Analyses

279

This allows the computation of the other elements t2:nc,i by linear regression as + 0 0 t2:nc ,i = −(U:,2: nc ) u:,1

(5.53)

This process is repeated for all individual columns of C to result in the ¯ T is complete matrix T (containing 1's in the first row). Finally, the product U the complete matrix C. As always with model-free analyses, only the shapes of the concentration profiles are determined. They have to be normalised in some way. We have seen in The Structure of the Eigenvectors (p.221) that the sign of the eigenvectors is not defined. In the present application, a concentration profile can be positive or, if negative, needs to be multiplied by -1. The two lines neg=­min(C)>max(C);             % sign of conc profiles C=C*diag((­1).^neg);            % reverse if negative 

in Main_Non_It_EFA.m check for negative concentrations and if necessary correct them. The subsequent computation of the absorption spectra A from C and Y is a simple linear regression. This is followed by the normalisation of the concentration profiles to a maximum of one, as has been outlined already in the preceding chapter Iterative Refinement of the Concentration Profiles. The normalisation is done using the routine norm_max.m (p.275). MatlabFile 5-39. Main_Non_It_EFA.m % Main_Non_It_EFA  [t,lam,Y,C_sim,A_sim]=Data_Chrom2a; [ns,nc]=size(C_sim); ne=nc;  [EFA_f,EFA_b]=EFA(Y,ne+1);      % perform EFA sig_level=0.01;                 % define cut­off level  % build window matrix  C_window=EFA_f(:,1:nc)>sig_level & fliplr(EFA_b(:,1:nc))>sig_level; [U,S,Vt]=svd(Y,0); U_bar=U(:,1:ne);  T=ones(nc,nc); for i=1:nc  U_i_0=U_bar(~C_window(:,i),:); T(2:nc,i)=­U_i_0(:,2:nc)\U_i_0(:,1); end  C=U_bar*T;  neg=­min(C)>max(C);             % sign of conc profiles C=C*diag((­1).^neg);            % reverse if negative A=C\Y;  [C_n,A_n]=norm_max(C,A);        % normalisation of calc. C and A  [C_sim_n,A_sim_n]=norm_max(C_sim,A_sim); % norm. of true C and A  subplot(2,1,1) 

280

Chapter 5

plot(lam,A_n,'­',lam,A_sim_n,'.'); xlabel('wavelength');ylabel('absorptivity'); subplot(2,1,2); plot(t,C_n,'­',t,C_sim_n,'.'); xlabel('time');ylabel('concentration'); 

absorptivity

2 1 0 -1 400

450

500 wavelength

550

600

concentration

1 0.5 0 -0.5 0

20

40

60

80

100

time

Figure 5-50. Result of non-iterative EFA. It is instructive to compare Figure 5-50 with Figure 5-42. The explicit computation is not only much faster, it also produces better results. This clearly is the consequence of the poor convergence of the iterative version of EFA.

5.4 Alternating Least-Squares, ALS The method of Alternating Least-Squares, ALS, is very simple and exactly for that reason it can be very powerful. ALS has found widespread applications and it is an important method in the collection of model-free analyses. In contrast to most other model-free analyses, ALS is not based on Factor Analysis. ALS should more correctly be called Alternating Linear Least-Squares as every step in the iterative cycle is a linear least-squares calculation followed by some correction of the results. The main advantage and strength of ALS is the ease with which any conceivable constraint can be implemented; its main weakness is the inherent poor convergence. This is a property ALS shares with the very similar methods of Iterative Target Transform Factor Analysis, ITTFA and Iterative Refinement of the Concentration Profiles, discussed in Chapters 5.2.2 and 5.3.3.

281

Model-Free Analyses

We start with the flow diagram in Figure 5-51 demonstrating the basic ideas. Of course, the data matrix is still Y and the goal is to decompose it into the product of the concentration matrix C and matrix A of molar absorptivities according to Chapter 3.1, Beer-Lambert's Law.

Initial guess for C A = C + Y

Corrections to A → A C = Y Ã+

Corrections to C → C

A  R = Y-C


= < ssqold

>

do something

= end Figure 5-51. Flow diagram for the ALS algorithm. The diagram starts with initial guesses for the concentration profiles C. It is, of course, equally possible to start with initial guesses for the component spectra A and swapping the order of the linear regression/correction steps:  and then A  while the structure of the rest is the same. calculating first C

5.4.1 Initial Guesses for Concentrations or Spectra Experience shows that often the quality of the initial guesses made for the concentration matrix C (or the matrix A of component spectra) is not crucial. As demonstrated in Figure 5-42, progress in that kind of algorithm typically is fast initially and slows down dramatically towards the minimum. Nevertheless, it cannot harm to have good initial starting matrices. Commonly implemented options include: •



combined eigenvalue curves, such as in Figure 5-44 non-iterative EFA result, Figure 5-50

282



Chapter 5

concentration windows (matrices formed by 1's and 0's), such as the bottom panel of Figure 5-39

5.4.2 Alternating Least-Squares and Constraints By far the most important aspect of the ALS algorithm is the ease of implementing restrictions. In the following we demonstrate this using a number of examples. The program Main_ALS.m forms the backbone of the ALS algorithm. It reads in the data set Data_Chrom2a (p.251) which simulates an overlapping chromatogram of three components. It is the data set we used previously in Chapter 5.3.3 to demonstrate the concepts of iterative and explicit computation of the concentration profiles, based on the window information from EFA. In order to facilitate the comparison of the results and the progress of the iterative process, we start all iterative attempts with the same concentration profiles. They are the combined eigenvalue traces of EFA, as shown in Figure 5-44. The very basic ALS program does not include a termination criterion for the iterative cycle. Just 100 iterations are performed. As convergence invariably slows down towards the minimum, it is not trivial to introduce a generally reliable termination criterion. The algorithm also does not incorporate steps that are required if there is divergence in an iteration. This is indicated by 'do something' in Figure 5-51. Again, it is not easy to develop generally applicable measures that force the iterations towards a good direction. There is no equivalent to the Levenberg/Marquardt method that deals with divergence in the Newton-Gauss algorithm. MatlabFile 5-40. Main_ALS.m % Main_ALS  [t,lam,Y,C_sim,A_sim]=Data_Chrom2a; [ns,nc]=size(C_sim); nl=length(lam);  ne=nc+1;                             % one extra singular value [EFA_f,EFA_b]=EFA(Y,ne);             % perform EFA EFA_f(isnan(EFA_f)==1)=0;            % replace NaN's by zeros EFA_b(isnan(EFA_b)==1)=0;            % replace NaN's by zeros % combined singular value curves C=min(EFA_f(:,1:nc),fliplr(EFA_b(:,1:nc)));  for it=1:100  C=norm_max(C); 

% normalization 

[C,A]=constraints_positiveCA(Y,C); 

% constraints 

R=Y­C*A;  ssq(it)=sum(sum(R.*R)); 

% residuals 

283

Model-Free Analyses

end  [C_n,A_n]=norm_max(C,A);  % norm. C, C_sim to max. 1 [C_sim_n,A_sim_n]=norm_max(C_sim,A_sim);  % and recalc. A, A_sim  subplot(3,1,1); plot(lam,A_n,'­',lam,A_sim_n,'.'); xlabel('wavelength');ylabel('absorptivity'); subplot(3,1,2); plot(t,C_n,'­',t,C_sim_n,'.'); xlabel('time');ylabel('concentration'); subplot(3,1,3); plot(log10(ssq)); xlabel('iteration');ylabel('log(ssq)'); axis([0 100 ­3 0]); 

absorptivity

2

1

concentration

0 400

450

500 wavelength

550

600

1

0.5

0 0

20

40

60

80

100

60

80

100

time

log(ssq)

0 -1 -2 -3 0

20

40 iteration

Figure 5-52. 100 iterations of ALS using the simplest constraint of setting negative values of A and C to zero. The markers represent the true spectra and concentration profiles, the lines the ALS result. The bottom panel shows the progress of the sum of squares. There are several types of constraints that can be used and generally the more constraints applied, the better the convergence and the better defined the results.

284

Chapter 5

The most important and almost universally applicable constraint is the nonnegativity of all elements of C and A. Obviously, neither concentrations nor molar absorptivities can be negative. In many ALS algorithms, this constraint is enforced by simply setting all negative entries in C and A to zero: MatlabFile 5-41. constraints_positiveCA.m function  [C,A]=constraints_positiveCA(Y,C)  A=C\Y;         % spectra A=A.*(A>0);    % positive  C=Y/A;         % conc profiles C=C.*(C>0);    % positive 

There are exceptions to the universality of this non-negativity constraint: e.g. CD or ESR spectra can be negative. Apart from that, both spectroscopies produce a signal that is a linear function of concentration and thus the equivalent of Beer-Lambert's law holds. In other words, the equation Y=CA applies and thus also the ALS algorithm. The alternating computation of the matrices A and C in linear least-squares fits, each followed by setting negative values to zero, is simple but very crude. This fact is reflected in the slow process of the sum of squares minimisation. Matlab supplies the function lsqnonneg that performs a non-negative leastsquares fit of the kind y=Ca+r, where y and a are column vectors. The function computes the best vector a with only positive entries. This equation corresponds to data acquired at only one wavelength. In our application, the columns of A have to be computed individually in a loop over all wavelengths, in each instance using the appropriate column of Y. C is the complete matrix of concentrations. It is, of course, the same for all wavelengths. The following function constraints_lsqnonneg.m replaces the function constraints_positiveCA.m. (Naturally, the call in the main program, Main_ALS.m, needs to be adapted). All columns a:,j of A are computed sequentially in a loop. In the alternate computation of the best C from A, the same function can be used. It computes the rows of C in an analogue loop, using the appropriate row of Y and the complete matrix A. The computation of positive rows of C, using lsqnonneg requires the appropriate transpositions for the rows of C and Y and the matrix A. MatlabFile 5-42. constraints_lsqnonneg.m function  [C,A]=constraints_lsqnonneg(Y,C)  [ns,nl]=size(Y);  for j=1:nl      % pos spectra (MATLAB) A(:,j)=lsqnonneg(C,Y(:,j)); end 

285

Model-Free Analyses

for j=1:ns      % pos conc. (MATLAB) C(j,:)=lsqnonneg(A',Y(j,:)')'; end 

absorptivity

2

1

concentration

0 400

450

500 wavelength

550

600

1

0.5

0 0

20

40

60

80

100

60

80

100

time

log(ssq)

0 -1 -2 -3 0

20

40 iteration

Figure 5-53. ALS using the Matlab function lsqnonneg.m for non­ negative linear least-squares fitting. Compared to Figure 5-52, the resultant concentration profiles and spectra appear to be very similar. The plot of the development of the quality of the fit indicates that a smaller sum of squares is achieved in fewer iterations. However, the calculation is much slower and there is no obvious and significant benefit. In Constraint: Positive Component Spectra (p.168), we introduced an improved, much faster matrix based function nonneg.m (provided by C. Andersson) that is more efficient than the Matlab function lsqnonneg.m. The result of implementing the function constraints_nonneg.m is identical but achieved much faster. MatlabFile 5-43. constraints_nonneg.m function  [C,A]=constraints_nonneg(Y,C)  A=nonneg(Y',C')';   % pos spectra (Andersson) C=nonneg(Y,A);      % pos conc. (Andersson) 

The secondary hump in the third concentration profile in Figure 5-52 and Figure 5-53 obviously is not correct. In fact, often it is independently known

286

Chapter 5

that the concentration profiles are unimodal, i.e. they have only one maximum and continuously decrease on both sides of the maximum. This is certainly the case for chromatographic concentration profiles. The function constraints_nonneg_unimod.m implements this additional constraint by levelling off secondary maxima. It also uses nonneg.m for non-negative linear least-squares fits. The effect, as demonstrated in Figure 5-54, is clear; not only has the secondary maximum been suppressed, as a consequence, the absorption spectra also are closer to the true ones. MatlabFile 5-44. constraints_nonneg_unimod.m function  [C,A]=constraints_nonneg_unimod(Y,C) [ns,nc]=size(C);  A=nonneg(Y',C')';           % pos spectra (Andersson) C=nonneg(Y,A);              % pos conc. (Andersson) for j=1:nc                  % unimodal conc. profiles [m,p]=max(C(:,j)); for i=p:ns­1 if C(i+1,j)>C(i,j); C(i+1,j)=C(i,j);end end  for i=p:­1:2 if C(i­1,j)>C(i,j); C(i­1,j)=C(i,j);end end  end 

absorptivity

2

1

concentration

0 400

450

500 wavelength

550

600

1

0.5

0 0

20

40

60

80

100

60

80

100

time

log(ssq)

0 -1 -2 -3 0

20

40 iteration

Figure 5-54. ALS using the Matlab function constraints_ nonneg_unimod.m. performing positive linear least-squares and removing secondary maxima in the concentration profiles.

287

Model-Free Analyses

The unimodality constraint distorts the least-squares improvements and this is evident from the slower convergence. However, the constraint forces the concentration profiles to physically possible shapes. Another, most powerful constraint is based on concentration windows provided by EFA. MatlabFile 5-45. constraints_nonneg_window.m function  [C,A]=constraints_nonneg_window(Y,C,C_window)  [ns,nc]=size(C);  A=nonneg(Y',C')';    % pos spectra (Andersson) C=nonneg(Y,A);       % pos conc. (Andersson) C=C.*C_window;       % apply windows from EFA 

absorptivity

2

1

concentration

0 400

450

500 wavelength

550

600

1

0.5

0 0

20

40

60

80

100

60

80

100

time

log(ssq)

0 -1 -2 -3 0

20

40 iteration

Figure 5-55. ALS applying the Matlab function constraints_nonneg_window.m; using non-negative linear leastsquares and the window matrix applied to C. The matrix C_window contains the window information. It is composed of 1's and 0' indicating whether the particular value in the corresponding entry in

288

Chapter 5

the matrix C is known to be positive or zero, see Figure 5-37. This matrix C_window is computed directly before the ALS loop: sig_level=0.01;               % define cut off level  C_window=C>sig_level;         % build window matrix 

Of course the matrix C_window has to be passed as an argument into constraints_nonneg_window.m. Refer to Figure 5-43 for a discussion of the level of significance used above.

5.4.3 Rotational Ambiguity The one major problem with all model-free methods is the fact that often there is no unique solution for the task of decomposing the matrix Y of measurements into the product of two positive matrices C and A. In many instances, there is a whole range of possible solutions. Recall the original model-free method by Lawton-Sylvestre (p.231) that clearly results in bands of feasible solutions, see Figure 5-16. In the literature on model-free analyses, the expression 'rotational ambiguity' has been coined for such situations. In instances where there is rotational ambiguity, algorithms like ALS converge to one particular point within the range of possibilities. Importantly, the algorithm does not detect such situations and thus does not warn the user of the potential non-uniqueness of the result. It is difficult to generalise, but such ambiguous situations often occur when the concentration windows are overlapping in specific ways. Kinetic investigations are typical examples of rotational ambiguity as a result of very wide concentration windows. Using Data_ABC2.m (p.256), producing data mimicking a reaction A→B→C, instead of the chromatography data and applying constraints_nonneg.m results in Figure 5-56. The main program Main_ALS2.m is not listed here, it is virtually identical with Main_ALS.m. While the resulting concentration profiles, and in particular the computed spectra, seem to be reasonably close to the true ones, there are significant discrepancies, typical for model-free analyses. (a) The computed concentration profile for the intermediate component reaches zero at the end of the measurement. (b) The initial part of the concentration profile for the final product is wrong; it does not start with zero concentration. Both discrepancies are the result of rotational ambiguity. The minimal ssq, reached after relatively few iterations, reflects the noise of the data and not a misfit between CA and Y. ssq does not improve if the correct matrices C and A are used.

289

Model-Free Analyses

absorptivity

1

0.5

concentration

0 400

450

500 wavelength

550

600

1

0.5

0 0

20

40

60

80

100

60

80

100

time

log(ssq)

0 -1 -2 -3 0

20

40 iteration

Figure 5-56. ALS analysis of kinetic data Data_ABC2.m analysed using constraints_nonneg.m. Implementation of additional constraints can help remove, or at least reduce, rotational ambiguity. A possibility to narrow down the range is to utilize a known component spectrum. The simplest way of implementing known component spectra is to replace the appropriate spectrum within the iterations with the 'correct', known one. See constraints_nonneg_known_spec.m. We repeat, a powerful property of the ALS algorithm is the ease with which additional known information can be implemented. MatlabFile 5-46. constraints_nonneg_known_spec.m function  [C,A]=constraints_nonneg_known_spec(Y,C,A_sim)  A=nonneg(Y',C')';   % pos spectra (Andersson) A(2,:)=A_sim(2,:);  % known spectrum C=nonneg(Y,A);      % pos conc. (Andersson) 

290

Chapter 5

absorptivity

1.5 1 0.5

concentration

0 400

450

500 wavelength

550

600

1

0.5

0 0

20

40

60

80

100

60

80

100

time

log(ssq)

0 -1 -2 -3 0

20

40 iteration

Figure 5-57. ALS with known intermediate spectrum The improvement resulting from the incorporation of the correct spectrum for the intermediate is subtle but significant. The all resulting spectra are improved, with the intermediate spectrum, of course, correct. The new concentration profiles for the starting material A and product C are now correct, while the profile for the intermediate B is untouched. Compared to Figure 5-56, the minimal ssq after the incorporation of the correct spectrum, is not improved and this is a clear indication for rotational ambiguity. The incorrect concentration profile for the intermediate B indicates that there is still a reduced level of rotational ambiguity. In fact, the solution is still not unique; there is still a discrepancy in the concentration profile of the intermediate and small errors in the spectra. With the introduction of the one correct spectrum, the range of rotational ambiguity has been reduced but not totally removed.

5.5 Resolving Factor Analysis, RFA Resolving Factor Analysis, RFA, is an attempt to introduce the strengths of the Newton-Gauss algorithm into the model-free analysis methodology. As

Model-Free Analyses

we have seen in analyses can be accelerates as it methodologies is computations.

291

many instances, the iterative progress in model-free very slow. The Newton-Gauss algorithm, in contrast, converges towards the optimum. Combining the two a promising idea that should result in much faster

¯ T, is the core of RFA. See also its graphical Equation (5.49), C= U ¯ is known from the SVD of the measurement representation in Figure 5-45. U Y, C and also A, see equation(5.58), can be calculated as a function of a transformation matrix T. The residuals and the sum of squares are defined as R( T ) = Y − C(T ) × A(T )

ssq = ∑ ∑ Ri2, j

(5.54)

and are minimised by the Newton-Gauss method. For a three component system, the matrix T has nine elements and thus it appears that C and eventually the sum of squares are a function of nine parameters. As we will see in a moment there are actually fewer, only six, parameters to be fitted. The idea of RFA is to use the Newton-Gauss algorithm to fit this rather small number of parameters in T. To start the iterative cycle, we need a set of initial guesses, Tguess, for the parameters T. In order to compare the properties of RFA with the previous iterative methods, we use the same data set as generated by Data_chrom2a.m (p.251). Since the latest Newton-Gauss algorithm, nglm3.m (p.173), requires Matlab structures, the equivalent data are generated by the appropriate function Data_RFA_chrom2a.m. The Newton-Gauss algorithm requires initial estimates for the parameters in T. These can be computed from the same estimated concentration profiles Cguess as before (Figure 5-44). It is determined by

Tguess = U t C guess

(5.55)

An important issue needs to be discussed next. Multiplying a column of C with any number and its corresponding row of A with the inverse of that number, does not affect the product CA and thus this factor is not determined at all. It can be freely chosen. Due to this multiplicative ambiguity, only the shapes of the concentration profiles (and component spectra) can be determined by any model-free method and only additional quantitative information allows the absolute determination of C and A. Multiplying a concentration profile, or column of C, with a factor is equivalent to multiplication of the corresponding column of T with the same factor. Any one element of each column vector of T can be chosen freely while the other elements in that column define the shape of the concentration profile. In order to avoid numerical problems with very small or very large numbers in each column of T, we choose the largest absolute element of each column of the matrix of initial guesses Tguess and keep it

292

Chapter 5

fixed during the iterative refinement of the others. This reduces the number of parameters that need fitting to nc(nc-1) or in our example of a three component system from 9 to 6. The Newton-Gauss algorithm (nglm3.m), is called from Main_RFA.m, and requires a Matlab function that computes the residuals as a function of the parameters T, as defined in equation (5.54). This calculation is performed in the Matlab function Rcalc_RFA.m. ¯ T, see Figure 5-45. Elements of C that are First C is computed as C= U outside the concentration window and negative elements are set to zero. ¯¯ . This somewhat surprising equation can be derived A is computed as T-1SV in the following way

CA = USV

(5.56)

introducing the identity matrix TT-1 in the appropriate position: CA = UTT -1SV

(5.57)

A = T -1SV

(5.58)

¯ T, A must be: as C=U

Next, negative elements of A are set to zero and the residuals and the sum of squares are computed as indicated in Figure 5-51. The derivatives of the residuals with respect to the parameters are computed numerically by the Newton-Gauss algorithm. MatlabFile 5-47. Rcalc_RFA.m function [r,s]=Rcalc_RFA(s)  s.C=s.U_bar*s.T;                % calc. conc  s.C=s.C.*s.C_window;            % apply windows from EFA s.C(s.C 0  s.A=inv(s.T)*s.S_bar*s.V_bar;   % calc. A  s.A(s.A 0  R=s.Y­s.C*s.A;                  % residuals  r=R(:); s.ssq=sum(r.^2); s.ssq_all=[s.ssq_all;s.ssq]; 

There is one detail that necessitates a few additional comments: nglm3.m requires a vector (s.par_str) of strings that contains the names of all variables that are fitted. In the RFA application these are the nc(nc-1) elements of T, s.par_str=[s.T(1,1), s.T(1,2), …]. The function build_par_str.m does the job of finding the maximum element of each column of T and including the other elements into par_str. MatlabFile 5-48. build_par_str.m function par_str=build_par_str(T) [maxT,index]=max(abs(T)); k=0; for j=1:3 

Model-Free Analyses

for i=1:3  if i~=index(j) k=k+1; par_str{k}=['s.T(' int2str(i) ',' int2str(j) ')']; end  end  end  MatlabFile 5-49. Main_RFA.m % Main_RFA  s=Data_RFA_Chrom2a;  % get data into structure s s.fname='Rcalc_RFA';       % file to calc residuals  ne=s.nc; [EFA_f,EFA_b]=EFA(s.Y,ne); % perform EFA on ne sing. values EFA_f(isnan(EFA_f)==1)=0;  % replace NaN's by zeros EFA_b(isnan(EFA_b)==1)=0;  % replace NaN's by zeros % combined singular value curves C_guess=min(EFA_f(:,1:s.nc),fliplr(EFA_b(:,1:s.nc)));  sig_level=0.01;               % cut off level  s.C_window=C_guess>sig_level; % build window matrix C_guess=norm_max(C_guess);    % normalise  [U,S,Vt]=svd(s.Y,0); s.U_bar=U(:,1:ne); s.S_bar=S(1:ne,1:ne); s.V_bar=Vt(:,1:ne)';  s.T=s.U_bar'*C_guess;      % initial guesses for s.T from C_guess s.par_str=build_par_str(s.T);  % cell array of 's.T(i,j)' strings  % Newton­Gauss  s.ssq_all=[]; s.par=get_par(s);          % collects variable parameters into s.par s=nglm3(s);                % call ngl/m  fprintf(1,'s.T = \n');disp(s.T);fprintf(1,'\n');    % display T s.sig_r=sqrt(s.ssq/(s.ns*s.nl­length(s.par)));      % sigma_r s.sig_par=s.sig_r*sqrt(diag(inv(s.Curv)));          % sigma_par for i=1:length(s.par) fprintf(1,'%s: %g +­ %g\n',s.par_str{i}(3:end), ... s.par(i),s.sig_par(i)); end  fprintf(1,'sig_r: %g\n',s.sig_r);  [C_sim_n,A_sim_n]=norm_max(s.C_sim,s.A_sim); [C_n,A_n]=norm_max(s.C,s.A); subplot(3,1,1); plot(s.lam,A_n,'­',s.lam,A_sim_n,'.'); xlabel('wavelength');ylabel('absorptivity'); subplot(3,1,2); plot(s.t,C_n,'­',s.t,C_sim_n,'.'); xlabel('time');ylabel('concentration'); subplot(3,1,3); plot(log10(s.ssq_all),'.'); xlabel('iteration');ylabel('log(ssq)');  it=0, ssq=7.9338, mp=0, conv_crit=1 

293

294

Chapter 5

it=1, ssq=0.941161, mp=0, conv_crit=0.881373 it=2, ssq=0.176043, mp=0, conv_crit=0.812951 it=3, ssq=0.0444385, mp=0, conv_crit=0.74757 it=4, ssq=0.0145205, mp=0, conv_crit=0.673245 it=5, ssq=0.00511355, mp=0, conv_crit=0.647839 it=6, ssq=0.0029535, mp=0, conv_crit=0.422417 it=7, ssq=0.00241231, mp=0, conv_crit=0.183239 it=8, ssq=0.00237849, mp=0, conv_crit=0.0140165 it=9, ssq=0.00237848, mp=0, conv_crit=6.54114e­006 s.T =  ­5.7280   ­2.4717   ­3.9865  5.2605    0.7339   ­3.1774  2.4481   ­1.2217   ­0.3969  T(1,1): ­5.72798 +­ 0.0237521 T(2,1): 5.26049 +­ 0.0167212 T(2,2): 0.733879 +­ 0.0025855 T(3,2): ­1.2217 +­ 0.00220923 T(2,3): ­3.17735 +­ 0.0110746 T(3,3): ­0.396851 +­ 0.00511213 sig_r: 0.00106576 

absorptivity

2

1

concentration

0 400

450

500 wavelength

550

600

1

0.5

0 0

20

40

60

80

100

6

8

10

time

log(ssq)

2 0 -2 -4 0

2

4 iteration

Figure 5-58. RFA analysis of chromatography data.

Model-Free Analyses

295

It is illustrative to compare Figure 5-58 with the equivalent result of the ALS analysis in Figure 5-55. The most striking difference is the number of iterations required to reach the outcome. RFA arrives at the optimal resolution, within the constraints, in 10 iterations. ALS, using equivalent constraints results in acceptable matrices C and A but even after 100 iterations the optimum clearly has not been reached.

5.6 Principle Component Regression and Partial Least Squares, PCR and PLS Principal Component Regression, PCR, and Partial Least Squares, PLS, are the most widely known and applied chemometrics methods. This is particularly the case for PLS, for which there is a tremendous number of applications and a never-ending stream of proposed improvements. The details of these latest modifications are not within the scope of this book and we concentrate on the essential, classical aspects. One could argue whether PCR and PLS should be part of the chapter ModelBased Analyses or Model-Free Analyses. Both, PCR and PLS, are clearly not hard-model fitting methods in the way presented in Chapter 4, nor are they pure model-free analyses. They are somewhere in between, maybe closer to model-free analyses and that is the reason for discussing them here. PCR or PLS establish a mathematical relationship (calibration) between the matrix that is formed by the spectra taken of a collection of samples and the vector of properties or qualities for these same samples. Additionally, both methods allow the prediction of the quality for new samples, just based on their spectra. In contrast to most methods discussed so far, PCR and PLS do not require any order in the data set. In this chapter, we deviate from our well-established principle of generating 'measurements' and analysing them subsequently with the methods developed for the purpose. Such a procedure does not make much sense for PCR/PLS. At least it would be rather difficult to generate realistic data sets that are amenable to analysis by PCR or PLS. We decided to use a publicly available data set; the file corn.mat can be downloaded from http://software.eigenvector.com/Data/Corn/index.html. This data set contains near infrared (NIR) spectra of a collection of 80 corn samples measured on three different instruments, together with the qualities 'Moisture', 'Oil', 'Protein' and 'Starch' for each sample. We use the example of 'Protein' measured on instrument 'mp6' to demonstrate the principles of the PCR/PLS analyses. In order to chemically analyse a sample of corn for its protein content, a rather complex analytical procedure (e.g. Kjeldahl analysis) is required, a slow and expensive process. In our example, the PCR/PLS group of methods replaces this procedure with a much faster spectroscopic analysis. First, a mathematical relationship is established from a calibration set, comprising a matrix of NIR-spectra of the collection of samples and the vector of

296

Chapter 5

corresponding qualities. This calibration can subsequently be used to predict the particular quality for a new sample from its NIR-spectrum alone, thus avoiding an expensive experimental analysis. A more traditional spectroscopy based approach for corn analysis would be to investigate whether there is a peak in the NIR-spectrum that correlates well with the protein content, or whether there is a ratio of peaks that correlates well, or … whatever else the scientist can think of and is prepared to try. Evidently, there is a tremendous number of potential combinations and permutation one could try. PCR/PLS do this job in a much more elegant and efficient way.

5.6.1 Principal Component Regression, PCR In the example we deal with a collection of ns=80 corn samples for which we have the NIR spectra, measured at nl=700 wavelengths, and 80 corresponding qualities (protein contents). The Matlab script Main_PCR.m first reads in the complete corn data. Then it execute stepwise all the tasks that are described in the following. In order to test PCR and later PLS, we remove a random selection of 10 test samples from the total data set; the 10 test spectra are collected row-wise in the matrix Ys and the corresponding 'known' qualities in a column vector qs,known. The remaining spectra are organised in the same way in the matrix Y of dimensions 70×700. For each one of the samples we also know the protein content; we collect these qualities in the vector q with 70 entries. In the following, Y and q serve as the calibration set that is used later to predict the unknown qualities qs for the test set Ys. The predicted qs can then be compared with the 'known' qualities qs,known. MatlabFile 5-50. Main_PCR.m % Main_PCR  load corn.mat mp6spec propvals; % load corn data set Y_data=mp6spec.data;            % NIR spectra q_data=propvals.data(:,3);      % protein qualities  ns=length(q_data); lam=[1100:2:2498]; plot(lam,Y_data); xlabel('wavelength')  rand('seed',1);             % initialise random number generator s=ceil(rand(10,1)*ns);      % random selection of 10 samples Y=Y_data; Y(s,:)=[];        % calibration set excluding 10 samples q=q_data; q(s)=[]; Y_s=Y_data(s,:);            % 'unknown' test samples q_s_k=q_data(s);            % their 'known' qualities 

297

Model-Free Analyses

0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 -0.1 1000

1500

2000

2500

wavelength

Figure 5-59. Collection of NIR spectra of 80 corn samples. There is a fair amount of structure in the NIR spectra but the differences between the spectra are rather subtle. Obviously, the protein content cannot easily be read from any of the peaks. Mean-Centring, Normalisation

There are numerous publications proposing a glut of data treatment methods prior to PCR/PLS. Well established, tested and essentially universally applied are mean-centring and normalisation of the data. We have seen in Mean Centring, Closure (p.239) that mean centring reduces the dimensionality by one, which of course cannot harm. In PCR/PLS it is also common to normalise to the standard deviation of the signals. Both are implemented in Main_PCR.m. MatlabFile 5-51. Main_PCR.m … continued % Main_PCR ... continued (data pre­treatment)  meanY=mean(Y);              % mean centring Y and q meanq=mean(q); Y_mc=Y­repmat(meanY,size(Y,1),1); q_mc=q­meanq; norm_coef=1./std(Y_mc);     % normalisation of Y_mc  Y_mc_n=Y_mc*diag(norm_coef); 

Mean-centring and normalisation are optional. The PCR (and PLS) algorithm are essentially independent of the nature of pre-treatment of the data, only the centring has to be reversed in the prediction step. In the programs we

298

Chapter 5

indicate the levels of pre-treatment, i.e. Y→Ymc→Ymc,n and q→qmc, while in the equations in the text we do not. PCR Calibration

It is possible to develop the ideas behind PCR and to a lesser extent behind PLS, based on chemical ideas and intuition. Naturally, this is not the only way and both PCR and PLS methods have been developed on pathways that are theoretically oriented. While we do not know the components of the corn samples nor their component spectra, nor even how many components there are, we can still assume that Beer-Lambert's law holds and we can write Y=CA (see Chapter 3.1). There is nothing new here. For a chemist, it intuitively makes sense to assume that the quality 'protein content' is related to the concentrations of the components in the mixture. The simplest assumption is that the protein content is the weighted average of the concentrations of the relevant individual components. Some components have a high protein content others might even have a negative influence. This is best expressed in a matrix equation: nc ns

× nc

=

q

C

b’

+

ns

r

(5.59)

All we know at present is the vector of qualities q and that q might be approximated by the product Cb'. We do not have an idea about the number of components, nc, nor about C or b'. ¯ T, C is the product of U ¯ and a Now we remember equation (5.49) C= U transformation matrix T. Introduction of this equation into equation (5.59) results in

q = UTb' + r = Ub + r

(5.60)

where the product Tb' is replaced with the column vector b. The quality vector q is now approximated by a linear combination of the columns of the ¯ . This might be surprising but if (5.59) makes sense, eigenvector matrix U ¯ is known from the SVD of (5.60) does as well. The important aspect is that U Y. Note, we still do not know how many components there are, or how many ¯ . We come back to that question of the eigenvectors we should retain in U shortly. The computation of the best b, the one for which the residual vector r is ¯ minimal, is a linear least-squares calculation. Due to the orthonormality of U it is particularly easy:

299

Model-Free Analyses

b = Ut q

(5.61)

This allows us to compute the PCR approximation qPCR for the quality vector q: qPCR = Ub = UU t q

(5.62)

This should remind the reader of e.g. equation (5.28). The vector qPCR is ¯ . The PCR nothing but the projection of the quality vector q into the space U ¯ , it is bad otherwise. calibration is good if the vector q is close to the space U The Matlab function PCR_calibration.m performs the PCR calibration according to equation (5.62). Note that we use ne=12 eigenvectors in the above calculations. This is the optimal number for prediction, as we show in Cross Validation (p.303). The reader is invited to play with this number and observe the effect. The routine PCR_calibration.m also returns a 'prognostic vector' vprog. It is used for prediction and its function is explained in the next Chapter PCR Prediction. MatlabFile 5-52. Main_PCR.m … continued % Main_PCR ... continued  (calibration)  ne=12;                      % no of factors for calibration  [q_PCR,v_prog]=PCR_calibration(Y_mc_n,q_mc,ne); % calibration q_PCR=q_PCR+meanq;          % undo mean centering in q_PCR  plot(q,q_PCR,'.') xlabel('q');ylabel('q_P_C_R') 

10

9.5

q

PCR

9

8.5

8

7.5 7.5

8

8.5

9

9.5

10

q

Figure 5-60. PCR calibration of the corn data using 12 factors: qPCR versus the actual qualities q.

300

Chapter 5

MatlabFile 5-53. PCR_calibration.m function [q_PCR,v_prog]=PCR_calibration(Y_mc_n,q_mc,ne)  [U,S,Vt]=svd(Y_mc_n,0); U_bar=U(:,1:ne); S_bar=S(1:ne,1:ne); V_bar=Vt(:,1:ne)'; q_PCR=U_bar*U_bar'*q_mc;  v_prog=V_bar'/S_bar*U_bar'*q_mc;     % prognostic vector 

PCR Prediction

So far, the program Main_PCR.m covers the calibration part of PCR. As Figure 5-60 demonstrates, there is a very reasonable mathematical relationship between the quality 'protein' and the NIR-spectra of the collection of corn samples; the correlation between measured and PCRmodelled protein contents is convincing. How can we use these results to predict the quality qs of a new sample, just based on its NIR-spectrum? The ¯ b. Each individual property qi is approximated relevant equation is (5.60) q=U ¯ and b. The calculation of by the product of the corresponding row of u ¯ i,: of U the quality qs for a new sample is done in an analogous way: q s = us b

(5.63)

However, we need to determine the row vector u ¯ s corresponding to the new sample. Figure 5-61 attempts to represent the relationship between the spectrum of a new sample, ys, and the Singular Value Decomposition of Y itself.

Y

ys

= U

S

V

us

SV y s = u s SV

Figure 5-61. Relationship between a new sample spectrum ys and its representation u ¯ s in the eigenvector space. The new spectrum, ys, is shown as the grey row underneath Y and the ¯. corresponding u ¯ s as the grey row underneath U

y s = u s SV

(5.64)

¯¯ . u The spectrum of the new sample is the product of u ¯ s and the matrix SV ¯s contains the coordinates of spectrum ys in the eigenvector space spanned by ¯¯ . Rearranging equation (5.64), u SV ¯ s is computed as:

301

Model-Free Analyses

u s = y s V t S-1

(5.65)

Inserting this equation into equation (5.60) and substitution of b by equation (5.61) leads to: qs = u s b

= y s V t S-1U t q

(5.66)

¯, S ¯, V ¯ and the vector b are determined by the calibration set Y The matrices U ¯ tS ¯ -1 U ¯ tq is also completely determined by the and q; thus the product V calibration. It is a column vector of dimension nl×1 (nl is the number of wavelengths at which the spectra are taken), we call it the prognostic vector, vprog. The quality qs for any new sample can be predicted by the product of its spectrum ys and the prognostic vector vprog: qs = y s v prog

v prog = V t S-1U t q

(5.67)

This prognostic vector can reveal interesting insight into the relationship between the qualities q and the spectra of the calibration set Y. Note that the prognostic vector vprog has already been computed in the function PCR_calibration. MatlabFile 5-54. Main_PCR.m … continued % Main_PCR ... continued (calibration)  subplot(2,1,1);plot(lam,meanY);axis tight; subplot(2,1,2);plot(lam,v_prog);axis tight; xlabel('wavelength') 

0.6 0.4 0.2 0 1200

1400

1600

1800

2000

2200

2400

1200

1400

1600 1800 2000 wavelength

2200

2400

0.2 0.1 0 -0.1

Figure 5-62. The mean spectrum and the prognostic vector.

302

Chapter 5

Figure 5-62 displays the mean spectrum and the prognostic vector vprog. The vector product ys×vprog is the sum over all the products of pairs of elements in ys and vprog. Thus, if at a certain wavelength the prognostic vector vprog has a positive value, a high value in ys at this wavelength would add to the quality, a negative value in vprog would subtract. As an example, consider the wavelength 2100nm highlighted by the dotted line in Figure 5-62: the prognostic vector is negative, indicating that the peak in the spectrum at this wavelength is 'bad' for the protein content. Samples with shifted peaks towards longer wavelength would have an increased quality qs as the prognostic vector has a strong positive contribution at around 2200nm. For the sake of completeness, we introduce an alternative at this stage and present a more theoretical but quicker path for the development of the PCR calibration and prediction equations. The starting concept is to assume there is a linear relationship between q and the matrix Y:

q = Yb''

×

≈ q

(5.68)

Y

b”

Figure 5-63. Graphic representation of equation (5.68) The optimal vector b" cannot be computed as b"=Y+q since the pseudoinverse Y+ is not defined, its calculation would include the inversion of a ¯¯¯ , b" can then be rank deficient matrix. The way out is to replace Y with USV computed as

b'' = V t S −1U t q

(5.69)

and the prediction of qs of a new sample is simply

qs = y s b'' = y s V t S −1U t q = y s v prog

(5.70)

The equations, of course, are the same as developed in the derivations given previously. And it turns out that b'' is the prognostic vector b''=vprog. Now we use the information gathered so far for the prediction of the 10 test samples Ys removed from the complete data set at the very beginning. The function PCR_PLS_pred.m does the work according to equation (5.70). Importantly the mean-centring and normalisation have to be performed in exactly the same way as in the calibration. MatlabFile 5-55. PCR_PLS_pred.m function q_s_pred=PCR_PLS_pred(Y_s,meanY,meanq,norm_coef,v_prog) 

303

Model-Free Analyses

Y_s_mc   = Y_s­repmat(meanY,size(Y_s,1),1);  % mean centre  Y_s_mc_n = Y_s_mc*diag(norm_coef);           % normalise  q_s_mc   = Y_s_mc_n*v_prog;                  % predict q_s_pred = q_s_mc+meanq;                     % undo mean centring  MatlabFile 5-56. Main_PCR.m … continued % Main_PCR ... continued (prediction)  q_s_pred=PCR_PLS_pred(Y_s,meanY,meanq,norm_coef,v_prog); plot(q_s_k,q_s_pred,'.'); xlabel('q_{s,known}');ylabel('q_{s,pred}'); axis([7.5 10 7.5 10]) 

10

9.5

q s,pred

9

8.5

8

7.5 7.5

8

8.5 q

9

9.5

10

s,known

Figure 5-64. PCR prediction of 10 new corn samples. The prediction result, as shown in Figure 5-64, for the 10 samples that were removed from the total data set is convincing. ¯¯¯ . Remember, for the calculations so far we used ne=12 eigenvectors in USV Now, we need to return to the question on how this number is determined. The main goal of PCR/PLS is the prediction of the qualities of new samples based on prior calibration using a suitable known calibration set. The best number of eigenvectors is the number that results in the best prediction. It's as easy as that. Cross Validation

¯, S ¯ and V ¯ is determined prior to the The optimal number of eigenvectors in U predictions of new samples, it can be seen as part of the calibration. The

304

Chapter 5

process is only discussed here because it contains both calibration and prediction steps. The most common and intuitive method for the determination of this number of eigenvalues is called Cross Validation. The idea is to remove one (or several samples) from the calibration set, use what is left for the computation of a new calibration, and use it to predict the quality of the removed sample(s). Each prediction is compared with the actual quality that is known as the removed sample really is part of the total calibration set. In a loop all samples are removed either one by one or in groups and after recalibration with the reduced calibration set their qualities are predicted and compared with the true values. In order to determine the best number of eigenvectors this procedure is repeated in a big loop systematically trying all numbers of eigenvectors. This complete procedure is called Cross Validation. The continuation of Main_PCR.m calls the function PCR_cross.m. This function performs the systematic cross validation for up to nemax=40 eigenvectors. The computations result in a plot of the accuracy of the prediction as a function of the number of eigenvectors. Hopefully, the graph has a clear minimum! MatlabFile 5-57. PCR_cross.m function [q_s_cross,PRESS]=PCR_cross(Y,q,ne_max)  ns=length(q); for i=1:ns  i  Y_cal=Y(i~=1:ns,:);     % elim. i­th row of Y for new calib. set  q_cal=q(i~=1:ns,:);     % elim. i­th element of q, new qual. set y_s=Y(i,:);             % extract i­th row of Y as new sample  meanY_cal=mean(Y_cal);  % mean centring Y and q meanq_cal=mean(q_cal); Y_cal_mc=Y_cal­repmat(meanY_cal,ns­1,1); q_cal_mc=q_cal­meanq_cal; y_s_mc=y_s­meanY_cal;  norm_coef=1./std(Y_cal_mc);        % normalising Y Y_cal_mc_n=Y_cal_mc*diag(norm_coef); y_s_mc=y_s_mc.*norm_coef;  [U,S,Vt]=svd(Y_cal_mc_n,0);        % PCR calibration  for k=1:ne_max  U_bar=U(:,1:k); S_bar=S(1:k,1:k); V_bar=Vt(:,1:k)'; V_prog(:,k)=V_bar'/S_bar*U_bar'*q_cal_mc; end  q_s_cross(i,:)=y_s_mc*V_prog;             % prediction q_s_cross(i,:)=q_s_cross(i,:)+meanq_cal;  % undo mean centring end  PRESS=sum((q_s_cross­repmat(q,1,ne_max)).^2); 

305

Model-Free Analyses MatlabFile 5-58. Main_PCR.m … continued % Main_PCR ... continued (cross validation)  ne_max=40; [q_s_cross,PRESS]=PCR_cross(Y,q,ne_max); plot(1:ne_max,PRESS);ylabel('PRESS');xlabel('factors') 

16 14 12

PRESS

10 8 6 4 2 0 0

10

20 factors

30

40

Figure 5-65. PRESS-PCR for the corn data set. PRESS, the prediction sum of squares, is the measure for the accuracy of the prediction. It is the sum over all squared differences between crossvalidation predicted and true known qualities. PRESS = ∑ (q s ,crossi − qi )2 ns

i =1

(5.71)

¯, S ¯ In Main_PCR.m the PRESS values for all 1 to ne eigenvectors used in U ¯ to compute the predicted qualities qs,cross are stored in a vector PRESS and V that is displayed in Figure 5-65. The figure does not show a clear minimum. In Figure 5-66 we show the results of the cross-validation for ne=12; this number has already been used for the calibration in Figure 5-60. MatlabFile 5-59. Main_PCR.m … continued % Main_PCR ... continued (cross validation)  plot(q,q_s_cross(:,ne),'.'); xlabel('q');ylabel('q_{s,cross}'); axis([7.5 10 7.5 10]) 

306

Chapter 5

10

9.5

qs,cross

9

8.5

8

7.5 7.5

8

8.5

9

9.5

10

q

Figure 5-66. PCR cross-validation for the corn data set. Figure 5-66 shows the relationship between the true and cross-validation predicted qualities. Note the small but significant drop in the correlation compared to pure calibration, as shown in Figure 5-60. Calibration invariably produces a better correlation than prediction.

5.6.2 Partial Least Squares, PLS Partial Least Squares is the chemometrics method 'par excellence'. There is a tremendous number of published applications and also a large number of minor improvements to the original PLS algorithm. In order to understand the difference between the PCR and the PLS methods we first return to PCR. The two central equations of PCR are:

Y = USV and q = Ub

(5.72)

The first equation is the well-known Singular Value Decomposition. In the ¯ form the basis for the column vectors of Y. context of PCR the eigenvectors U The second equation in (5.72) attempts to also represent the column vector q ¯ . If both representations are good then PCR of qualities in the same space U works well, resulting in accurate predictions. A potential drawback of PCR is ¯ is defined solely by Y. Even if there is good reasoning for a the fact that U ¯ , as indicated in the derivation of equation relationship between q and U (5.60), it is somehow 'accidental'. The basic idea of PLS is to find a better set of basis vectors that represent adequately both Y and q. In the PLS literature, this basis T is often called

Model-Free Analyses

307

'scores'. Note, matrix T must not be confused with transformation matrices T we have used several times earlier in this chapter. As required for a decent set of basis vectors, the columns of T have to be orthogonal. The ideal basis for q would be q itself but it might not be a good basis for Y. Ideally, T is a compromise that serves as a good basis for q and Y. Obviously, there is not just one compromise and this to some extent explains the large number of modified PLS algorithms. Below, we give the complete program Main_PLS.m that runs the PLS computations. Its is identical to Main_PCR.m with the exception of the PLS functions PLS_calibration.m and PLS_cross.m that are called instead of the corresponding PCR routines; additional minor differences include the axis labelling etc. Mean-centring and normalisation are implemented as in PCR. Here, the complete listing is included while the equivalent Main_PCR.m has only been given in many little fragments. MatlabFile 5-60. Main_PLS.m % Main_PLS  load corn.mat mp6spec propvals; % load corn data set Y_data=mp6spec.data;            % NIR spectra q_data=propvals.data(:,3);      % protein qualities  ns=length(q_data); lam=[1100:2:2498]; plot(lam,Y_data); xlabel('wavelength')  rand('seed',1);             % initialise random number generator s=ceil(rand(10,1)*ns);      % random selection of 10 samples Y=Y_data; Y(s,:)=[];        % calibration set excluding 10 samples q=q_data; q(s)=[];          %  Y_s=Y_data(s,:);            % 'unknown' test samples q_s_k=q_data(s);            % their 'known' qualities  % Main_PLS ... continued (data pre­treatment)  meanY=mean(Y);              % mean centering Y and q meanq=mean(q); Y_mc=Y­repmat(meanY,size(Y,1),1); q_mc=q­meanq; norm_coef=1./std(Y_mc);     % normalisation of Y_mc  Y_mc_n=Y_mc*diag(norm_coef);  % Main_PLS ... continued (calibration)  ne=10;                      % no of factors for calibration  [q_PLS,v_prog]=PLS_calibration(Y_mc_n,q_mc,ne); % calibration q_PLS=q_PLS+meanq;          % undo mean centering in q_PLS  plot(q,q_PLS,'.') xlabel('q');ylabel('q_P_L_S')  subplot(2,1,1);plot(lam,meanY);axis tight; subplot(2,1,2);plot(lam,v_prog);axis tight; xlabel('wavelength') 

308

Chapter 5

% Main_PLS ... continued (prediction)  q_s_pred=PCR_PLS_pred(Y_s,meanY,meanq,norm_coef,v_prog); plot(q_s_k,q_s_pred,'.'); xlabel('q_{s,known}');ylabel('q_{s,pred}'); axis([7.5 10 7.5 10])  % Main_PLS ... continued (cross validation)  ne_max=40; [q_s_cross,PRESS]=PLS_cross(Y,q,ne_max); plot(1:ne_max,PRESS);xlabel('factors');ylabel('PRESS');  plot(q,q_s_cross(:,ne),'.'); xlabel('q');ylabel('q_{s,cross}'); axis([7.5 10 7.5 10]) 

PLS calibration

The PLS equations can be written in analogy to equation (5.72)

Y = TP and q = Tb

(5.73)

In the original PLS, the matrices T (scores), P (loadings) and the vector b are computed sequentially in the following way: (a) take q as a first estimate for the first basis vector t:,1. (b) q is assumed to be a basis for Y. Hence, we can approximate Y as Y=qw1,:, and calculate a best w1,: (loading weights) in a linear leastsquares fit

w1,: = q \ Y

(5.74)

In the standard PLS algorithm w1,: is normalised to unity length and subsequently, t:,1 is calculated as

t:,1 = Y / w1,:

(5.75)

This can be interpreted as one ALS iteration. If the iterations are continued, this process converges to the first eigenvector u:,1. The PLS compromise is to stop at one iteration. (c) this t:,1 is the first basis vector, it is the PLS analogue to the first eigenvector u:,1 in PCR. (d) Both Y and q are projected onto t:,1 and the residuals calculated. Importantly, the residuals are orthogonal to t:,1. rq = q − t:,1b1

Ry = Y − t:,1p1,:

where b1 and p1,: are computed in linear least-squares fits

(5.76)

Model-Free Analyses

b1 = t :,1\q

p1,: = t :,1\Y

309

(5.77)

(e) The remaining basis vectors t:,2:ne are computed in an adaptation of the NIPALS algorithm. q is replaced by the residual vector rq and is used as a new estimate for the next basis vector, t:,2; Ry replaces Y and the computation is continued at (a). The cycle (a)-(d) is repeated ne times. The optimal number ne is determined by cross-validation. The vectors t:,k, wk,:, pk,: and the scalars bk (k=1…ne) are collected in the matrices T, W, P and vector b. The function PLS_calibration.m is the PLS equivalent to PCR_calibration.m. The iterative loop implements equations (5.73)-(5.77). The prognostic vector vprog is introduced in the next section. MatlabFile 5-61. PLS_calibration.m function [q_PLS,v_prog]=PLS_calibration(Y_mc_n,q_mc,ne)  rq=q_mc; Ry=Y_mc_n; for k=1:ne  W(k,:)=rq\Y_mc_n; W(k,:)=W(k,:)/norm(W(k,:)); T(:,k)=Ry/W(k,:); P(k,:)=T(:,k)\Ry; b(k,1)=T(:,k)\rq; Ry=Ry­T(:,k)*P(k,:); rq=rq­T(:,k)*b(k,1); end  q_PLS=T*b;  v_prog=W'*((P*W')\b);             % prognostic vector 

PLS Prediction / Cross Validation

PLS prediction can be performed in analogy to PCR by PCR_PLS_pred.m (p.302) already introduced earlier. For this, we need to determine the prognostic vector, vprog. Using the results of the PLS calibration, it can be computed as:

v prog = W t (PW t )-1 b

(5.78)

During cross validation, it is most convenient to compute this prognostic vector as a function of the number of factors. This results in a collection of prognostic vectors, conveniently stored in a matrix V_prog. After determination of the optimal number of factors, the appropriate column can be selected as prognostic vector. The predicted quality qs for an unknown sample with the spectrum ys is then calculated as: qs = y s v prog

(5.79)

310

Chapter 5

For PCR and PLS, the principle of cross validation is identical. We refer to Cross Validation (p.303). The following function PLS_cross.m differs from PCR_cross only in the few lines performing the calibration and calculating the prognostic vector. MatlabFile 5-62. PLS_cross.m function [q_s_cross,PRESS]=PLS_cross(Y,q,ne_max)  ns=length(q); for i=1:ns  i  Y_cal=Y(i~=1:ns,:);     % elim. i­th row of Y for new calib. set  q_cal=q(i~=1:ns,:);     % elim. i­th elem. of q for new qual. set y_s=Y(i,:);             % extract i­th row of Y as new sample  meanY_cal=mean(Y_cal);  % mean centring Y and q meanq_cal=mean(q_cal); Y_cal_mc=Y_cal­repmat(meanY_cal,ns­1,1); q_cal_mc=q_cal­meanq_cal; y_s_mc=y_s­meanY_cal;  norm_coef=1./std(Y_cal_mc);               % normalising Y Y_cal_mc_n=Y_cal_mc*diag(norm_coef); y_s_mc=y_s_mc.*norm_coef;  rq=q_cal_mc;                              % PLS calibration  Ry=Y_cal_mc_n; for k=1:ne_max  W(k,:)=rq\Y_cal_mc_n; W(k,:)=W(k,:)/norm(W(k,:)); T(:,k)=Ry/W(k,:); P(k,:)=T(:,k)\Ry; b(k,1)=T(:,k)\rq; Ry=Ry­T(:,k)*P(k,:); rq=rq­T(:,k)*b(k,1); V_prog(:,k)=W(1:k,:)'*inv(P(1:k,:)*W(1:k,:)')*b(1:k,1); end  q_s_cross(i,:)=y_s_mc*V_prog;             % prediction q_s_cross(i,:)=q_s_cross(i,:)+meanq_cal;  % undo mean centring end  PRESS=sum((q_s_cross­repmat(q,1,ne_max)).^2); 

5.6.3 Comparing PCR and PLS Figure 5-67 displays the results of the cross validation computations of the corn data with PCR and PLS. The graph is fairly typical: PLS is consistently better at small numbers of factors and predictions are very similar at the optimal number of factors ne, which is 10 for PLS and 12 for PCR. Experience has shown that it is 'dangerous' to use an excessive number of factors (over-fitting) for thew prediction of new unknown samples. This is why we selected ne=12 rather than 23 for PCR.

311

Model-Free Analyses

PCR PLS

14

PRESS

12 10 8 6 4 2 5

10

15

20 25 factors

30

35

40

Figure 5-67. Comparison of the cross validation results for PCR and PLS. In view of the similarity of the PRESS results for PCR and PLS, it is not surprising that the predicted qualities are very similar for the two methods if the optimal numbers of factors is used. Figure 5-68 summarises the comparison.

qs,cross

10 9 PCR PLS

8 7.5

8

8.5

9

9.5

10

q PCR PLS

vprog

0.2 0 -0.2 1200

1400

1600 1800 2000 wavelength

2200

2400

qs,pred

10 9 PCR PLS

8 7.5

8

8.5 q

9

9.5

10

s,known

Figure 5-68. Comparison of predictions and prognostic vectors for PCR and PLS.

312

Chapter 5

The top panel compares the cross-validation predictions with the optimal number of factors of 12 for PCR and 10 for PLS. The middle panel shows the similarity of the two prognostic vectors. The bottom panel compares the prediction of the 10 test samples that were removed from the total data set prior to cross-validation. If a conclusion can be drawn, it is that PCR and PLS are virtually indistinguishable in their outcome. Yet, PLS appears to reach optimal prediction with fewer factors than PCR.

Further Reading Kinetics John Ross, Igor Schreiber, Marcel O. Vlad, and Adam Arkin. Determination of Complex Reaction Mechanisms: Analysis of Chemical, Biological, and Genetic Networks. Oxford University Press 2005 Robert W. Hay. Reaction Mechanisms of Metal Complexes. Albion/Harwood Pub 2000 James H Espenson. Chemical Kinetics and Reaction Mechanisms (2nd edition). McGraw-Hill 1995 S.K. Scott. Oscillations, Waves, and Chaos in Chemical Kinetics. Oxford. Oxford Chemistry Press 1994. Ralph G. Wilkins. Kinetics and Mechanisms of Reactions of Transition Metal Complexes. VCH 1991 N.M. Rodiguin, E.N. Rodiguina. Consecutive Chemical Reactions. Mathematical Analysis and Development. D. van Nostrand New York 1964 S.W. Benson. The Foundations of Chemical Kinetics. McGraw-Hill New York 1960

Equilibria Arthur Martell, Robert Hancock. Metal Complexes in Aqueous Solutions (Modern Inorganic Chemistry). Springer 1996 Arthur Martell, Ramunas J. Motekaitis. The Determination and Use of Stability Constants. (2nd Edition). Wiley 1992 Juergen Polster, Heinrich Lachmann. Spectrometric Titrations: Analysis of Chemical Equilibria. VCH 1989 Kenneth A. Connors. Binding Constants: The Measurement of Molecular Complex Stability. Wiley 1987 M.T. Beck, I. Nagypal. The Chemistry of Complex Equilibria. Van NostrandReinhold. London 1970

Chemometrics/Statistics/Data Fitting Richard G.Brereton. Applied Chemometrics for Scientists. Wiley 2007

314

Further Reading

Paul Gemperline (editor). Practical Guide to Chemometrics (2nd edition). CRC Press 2006 Bruns, Scarmino, de Barros Neto. Statistical Design - Chemometrics, Volume 25 (Data Handling in Science and Technology). Elsevier 2006 James Miller, Jane Miller. Statistics and Chemometrics for Analytical Chemistry (5th edition). Prentice Hall 2005 D. Brynn Hibbert, J. Justin Gooding. Data Analysis for Chemistry. Oxford University Press 2005 M.J. Adams. Chemometrics in Analytical Spectroscopy (2nd edition). The Royal Society of Chemistry 2004. Richard G. Brereton. Chemometrics: Data Analysis for the Laboratory and Chemical Plant. Wiley 2003 George A. F. Seber, C. J. Wild. Nonlinear Regression (Wiley Series in Probability and Statistics). Wiley 2003 Edmund R. Malinowski. Factor Analysis in Chemistry (3rd edition). Wiley 2002 Philip R. Bevington, D. Keith. Robinson. Data Reduction and Error Analysis (3rd edition). McGrawHill New York 2002 Peter C. Meier, Richard E. Zünd. Statistical Methods in Analytical Chemistry (2nd edition). Wiley 2000 Matthias Otto. Chemometrics: Statistics and Computer Application in Analytical Chemistry. Wiley 1999 Sigmund Brandt, Glen Gowan. Data Analysis: Statistical and Computational Methods for Scientists and Engineers (3rd edition). Springer 1998 Richard Kramer. Chemometrics Techniques for Quantitative Analysis. Marcel Dekker 1998 Kenneth R. Beebe, Randy J. Pell, Mary Beth Seasholtz. Chemometrics: A Practical Guide. Wiley 1998 E.J. Karjalainen, U.P. Karjalainen. Data Analysis for Hyphenated Techniques (Data Handling in Science and Technology, Vol. II). Elsevier 1996 Meloun Milan, Jiri Militky, and Michele Forina. Chemometrics for Analytical Chemistry (Vol I+II). Ellis Horwood 1994 Harald Martens, Tormod Naes. Multivariate Calibration. Wiley 1993 John Kalivas. Mathematical Analysis of Spectral Orthogonality. Marcel Dekker 1993 Richard C. Graham. Data Analysis for the Chemical Sciences: A Guide to Statistical Techniques. VCH 1993 Stephen Haswell (editor). Practical Guide to Chemometrics (1st edition). Marcel Dekker 1992

315

Further Reading

Peter Gans. Data Fitting in the Chemical Sciences by the Method of Least Squares. Wiley 1992 Ed Morgan. Chemometrics: Experimental Design. Wiley 1991 D.L. Massart, B.G.M. Vandeginste, S.N. Chemometrics: A Textbook. Elsevier 1988

Deming,

and

Y.

Michotte.

B. G. M. Vandeginste, L. M. C. Buydens, S. De Jong, and P. J. Lewi. Handbook of Chemometrics and Qualimetrics A/B. Elsevier 1988

Numerical Methods Michael B. Cutlip, Mordechai Shacham. Problem Solving in Chemical Engineering with Numerical Methods. Prentice Hall 2007 Bruce A. Finlayson. Introduction to Chemical Engineering Computing. Wiley 2006 Daniel Dubin. Numerical and Analytical Methods for Scientists and Engineers using Mathematica. Wiley 2003 James B. Riggs. Introduction to Numerical Methods for Chemical Engineers (2nd edition). Texas Tech University Press 1999. Alejandro Garcia. Numerical Methods for Physics (2nd edition). Prentice Hall 1999 William H. Press, Brian P. Flannery, Saul A. Teukolsky, and William T. Vetterling. Numerical Recipes in C: The Art of Scientific Computing (2nd edition). Cambridge 1996 Peter Pelikan, Michal Ceppan, and Marek Liska. Applications of Numerical Methods in Molecular Spectroscopy. CRC Press 1994 R. Bulirsch, J. Stoer. Introduction to Numerical Analysis. Springer New York 1993 Louis Lyons. A Practical Guide to Data Analysis for Physical Science Students. Cambridge University Press 1991

Matlab and Excel Kenneth Beers. Numerical Methods for Chemical Engineering: Applications in MATLAB. Cambridge University Press 2006 Rudra Pratap. Getting Started with MATLAB 7: A Quick Introduction for Scientists and Engineers (The Oxford Series in Electrical and Computer Engineering). Oxford University Press 2005 Gerard Verschuuren. Excel for Scientists Professionals series). Holy Macro! Books 2005

and

Engineers

(Excel

for

316

Further Reading

Steve Chapra. Applied Numerical Methods with MATLAB for Engineers and Scientists. McGraw-Hill 2004 Robert de Levie. Advanced Excel for Scientific Data Analysis. Oxford University Press 2004 S.C. Bloch. Excel for Engineers and Scientists (2nd edition). Wiley 2003 Bernard Liengme. A Guide to MS Excel 2002 for Scientists and Engineers. Butterworth, Heinemann 2002 E. Joseph Billo. Excel for Chemists: A Comprehensive Guide (2nd edition). Wiley-VCH 2001 Robert de Levie. How to Use Excel in Analytical Chemistry and in General Scientific Data Analysis. Cambridge University Press 2001 Alkis Constantinides, Navid Mostoufi. Numerical Methods for Chemical Engineers with Matlab Applications. Prentice Hall 1999 Michael B. Cutlip, Mordechai Shacham. Problem Solving in Chemical and Biochemical Engineering with POLYMATH, Excel, and MATLAB (2nd Edition). Prentice Hall 1998 Dermot Diamond, Venita C. A. Hanratty. Spreadsheet Applications in Chemistry using MS Excel. Wiley 1997 William J. Orvis. Excel for Scientist and Engineers (2nd edition). SYBEX 1995

List of Matlab Files MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE

3-1. GAS_LAWS.M 3-2. GAS_LAWS.M …CONTINUED 3-3. BEER_LAMBERT.M 3-4. GAUSS.M 3-5. GAUSS_CURVE.M 3-6. GAUSS_CURVE2.M 3-7. GAUSS_SK.M 3-8. GAUSS_SKEWED.M 3-9. NEWTONRAPHSON.M 3-10. EQ1.M 3-11. EQ2.M 3-12. EDTA.M 3-13. EGG_CARTON.M 3-14. EGG_CARTON.M …CONTINUED 3-15. EGG_CARTON.M …CONTINUED 3-16. TWO_EQUATIONS.M 3-17. NONLINEQ.M 3-18. MAIN_NONLINEQ.M 3-19. ATOB.M 3-20. ODE_AUTOCAT.M 3-21. AUTOCAT.M 3-22. ODE_ZERO_ORDER.M 3-23. ZERO_ORDER.M 3-24. ZERO_ORDER.M …CONTINUED 3-25. ODE_LOTKA_VOLTERRA.M 3-26. LOTKA_VOLTERRA.M 3-27. LOTKA_VOLTERRA.M …CONTINUED 3-28. ODE_BZ.M 3-29. BZ.M 3-30. ODE_LORENZ.M 3-31. LORENZ.M 3-32. LORENZ.M …CONTINUED 4-1. DATA_MXB.M 4-2. MAIN_MXB.M 4-3. MAIN_MXB.M …CONTINUED 4-4. DATA_DECAY.M 4-5. MAIN_DECAY_2D.M 4-6. MAIN_DECAY_SSQ.M 4-7. TAN_POLY.M 4-8. MAIN_DECAY.M 4-9. DATA_DECAY_OFFSET.M 4-10. MAIN_DECAY_OFFSET.M 4-11. MAIN_SAVGOL.M 4-12. SAVGOL_BAD.M

29 30 34 37 37 38 39 39 53 57 59 66 72 72 73 74 75 75 79 87 88 89 90 91 93 93 94 96 96 98 98 99 104 104 104 106 106 107 124 127 129 129 132 133

318 MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE

List of Matlab Files 4-13. SAVGOL.M 4-14. MAIN_SAVGOL_DERIV.M 4-15. SAVGOL_DERIV.M 4-16. MAIN_LOLIPOP.M 4-17. LOLIPOP.M 4-18. DATA_ABC.M 4-19. MAIN_ABC_3D 4-20. MAIN_ABC_LIN1.M 4-21. MAIN_ABC_LIN2.M 4-22. DATA_EXP.M 4-23. MAIN_EXP_2D.M 4-24. MAIN_NG1.M 4-25. MAIN_NG2.M 4-26. DATA_CHROM.M 4-27. MAIN_CHROM.M 4-28. NGLM.M 4-29. RCALC_CHROM.M 4-30. MAIN_CHROM2.M 4-31. MAIN_ABC.M 4-32. NGLM2.M 4-33. RCALC_ABC.M 4-34. RCALC_ABC2.M 4-35. DATA_EQAH2.M 4-36. MAIN_EQAH2.M 4-37. GET_PAR.M 4-38. PUT_PAR.M 4-39. NGLM3.M 4-40. RCALC_EQAH2.M 4-41. DATA_EQFIX.M 4-42. RCALC_EQFIX.M 4-43. MAIN_EQFIX.M 4-44. MAIN_ABC_RED.M 4-45. ODEAPB_C_REV.M 4-46. DATA_GLOB.M 4-47. RCALC_GLOB.M 4-48. MAIN_GLOB.M 4-49. DATA_EMISSION.M 4-50. MAIN_EMISSION_LIN.M 4-51. MAIN_EMISSION_LIN.M …CONTINUED 4-52. MAIN_EMISSION_WEIGHTED.M 4-53. RCALC_EMISSION_WEIGHTED.M 4-54. MAIN_DECAY_SIMPLEX.M 4-55. SSQCALC_DECAY.M 4-56. MAIN_ABC_SIMPLEX.M 4-57. SSQCALC_ABC.M 5-1. MAIN_SVD1.M 5-2. DATA_CHROM2.M 5-3. MAIN_SVD2.M

134 136 137 138 139 143 143 144 145 150 150 151 154 158 158 159 160 161 165 166 167 168 170 172 173 173 173 174 178 178 179 182 185 185 186 187 191 192 194 195 195 205 206 207 207 216 219 219

List of Matlab Files MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE

5-4. MAIN_SVD2.M …CONTINUED 5-5. MAIN_SVD2.M …CONTINUED 5-6. MAIN_SVD2.M …CONTINUED 5-7. MAIN_SVD2.M …CONTINUED 5-8. DATA_AB.M 5-9. MAIN_PLOT_AB.M 5-10. LAWTONSYLVESTRE.M 5-11. DATA_EQAH2A.M 5-12. MAIN_EQAH2A.M 5-13. MAIN_EV_SPACE.M 5-14. MAIN_EV_SPACE.M …CONTINUED 5-15. MAIN_MEANCENTER.M 5-16. MAIN_HELPP.M 5-17. MAIN_HELPP.M …CONTINUED 5-18. MAIN_NOISERED1.M 5-19. DATA_AB2.M 5-20. MAIN_NOISERED2.M 5-21. MAIN_TFA.M 5-22. DATA_CHROM2A.M 5-23. MAIN_ITTFA.M 5-24. MAIN_SYM_ABC.M 5-25. MAIN_SYM_ABC_REV.M 5-26. MAIN_SYM_ABC_REV.M …CONTINUED 5-27. DATA_ABC2.M 5-28. MAIN_TTF.M 5-29. MAIN_EFA1.M 5-30. MAIN_EFA1.M …CONTINUED 5-31. MAIN_EFA1.M …CONTINUED 5-32. MAIN_EFA2.M 5-33. EFA.M 5-34. MAIN_EFA3.M 5-35. DATA_EQAH4A.M 5-36. MAIN_FSW_EFA.M 5-37. MAIN_IT_EFA.M 5-38. NORM_MAX.M 5-39. MAIN_NON_IT_EFA.M 5-40. MAIN_ALS.M 5-41. CONSTRAINTS_POSITIVECA.M 5-42. CONSTRAINTS_LSQNONNEG.M 5-43. CONSTRAINTS_NONNEG.M 5-44. CONSTRAINTS_NONNEG_UNIMOD.M 5-45. CONSTRAINTS_NONNEG_WINDOW.M 5-46. CONSTRAINTS_NONNEG_KNOWN_SPEC.M 5-47. RCALC_RFA.M 5-48. BUILD_PAR_STR.M 5-49. MAIN_RFA.M 5-50. MAIN_PCR.M 5-51. MAIN_PCR.M … CONTINUED

319 220 221 222 224 224 225 234 236 236 237 238 240 241 242 243 244 245 248 251 252 254 255 255 256 256 261 263 264 265 266 267 268 269 272 275 279 282 284 284 285 286 287 289 292 292 293 296 297

320 MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE MATLABFILE

List of Matlab Files 5-52. 5-53. 5-54. 5-55. 5-56. 5-57. 5-58. 5-59. 5-60. 5-61. 5-62.

PCR_CALIBRATION.M MAIN_PCR.M … CONTINUED MAIN_PCR.M … CONTINUED PCR_PLS_PRED.M MAIN_PCR.M … CONTINUED PCR_CROSS.M MAIN_PCR.M … CONTINUED MAIN_PCR.M … CONTINUED MAIN_PLS.M PLS_CALIBRATION.M PLS_CROSS.M

300 299 301 302 303 304 305 305 307 309 310

List of Excel Sheets EXCELSHEET EXCELSHEET EXCELSHEET EXCELSHEET EXCELSHEET EXCELSHEET EXCELSHEET EXCELSHEET EXCELSHEET EXCELSHEET EXCELSHEET EXCELSHEET

3-1. 3-2. 3-3. 3-4. 3-5. 3-6. 4-1. 4-2. 4-3. 4-4. 4-5. 4-6.

CHAPTER2.XLS-FESCN CHAPTER2.XLS-EQML2 CHAPTER2.XLS-CASO4 CHAPTER2.XLS-H3PO4 CHAPTER2.XLS-EQSYS CHAPTER2.XLS-RUNGEKUTTA CHAPTER3.XLS-TRENDLINE CHAPTER3.XLS-LINEST CHAPTER3.XLS-PSEUDOINVERSE CHAPTER3.XLS-CHROM CHAPTER3.XLS-KINETICS CHAPTER3.XLS-EMISSION

42 61 63 67 75 83 111 126 147 208 210 212

Index χ2-fitting,

D

# data generation, 34

189, 211

linear, 190

chromatograms, 36

non-linear, 195

spectra, 36

standard deviation of the residuals, 194

degrees of freedom, 122, 161, 165, 180, 189, 194, 195

A

design matrix, 115

activity coefficients, 44, 62

E

Debye-Hückel, 63 Alternating Least-Squares (ALS), 280 constraints, 282

eigenvalues, 215 eigenvectors, 181, 215

concentration windows, 287

noise, 221

known component spectrum, 289

significant, 218

non-negativity, 284

structure, 221

unimodality, 286

equilibria, 32, 40

initial guesses, 281

beta-values, 43 complex, 48, 62

B Beer-Lambert's law, 33 absorption, 33 calculation of component spectra using known concentrations, 144 calculation of concentrations using known component spectra, 145 component spectrum, 34 concentration profiles, 34 molar absorptivity, 33 path length, 33

components, 40, 44, 49 concentration profiles, 42, 56 degree of dissociation, 65 deprotonation of coordianted water, 47 equilibrium concentrations, 47 equilibrium constant, 44 explicit equations, 41, 64 formation constant, 44 general case, 43 general solution, 48 hydroxide ion concentration, 47 ionic product of water, 58

C

model, 45 notation, 43

chemical equilibrium, 40 chemometrics, 231 chromatography, 36

numerical solution. See Newton-Raphson algorithm simple case Fe/SCN, 40

elution profiles, 36

species, 40, 44, 49

overlapping peaks, 36

stoichiometry, 47, 53

closure, 227, 239 constraints positive component spectra, 168 coordination chemistry, 45 curvature matrix, 122, 161, 202

total concentrations, 41, 43, 47 Evolving Factor Analysis (EFA), 259 classical, 260 analysis of titration data, 267 backward EFA, 262 concentration windows, 264, 271

323

Index evolving singular values, 260

spectrophotometric titration of a 2-protic acid, 236, 264

forward EFA, 261 rank analysis, 261

spectrophotometric titration of a 4-protic acid, 268

significance level, 261, 273 Fixed-Size Window EFA, 268

steady state, 91

secondary analyses, 271

time resolved single photon counting, 191, 211

explicit computation of concentration profiles, 276 initial guesses of concentration profiles, 274 iterative concentration refinement, 271 examples

titration curve for a 2-protic acid, 170 titration of acetic acid, 58

χ2-fitting, 211

Excel

$ operator, 14

0th order reaction, 89

element-wise operations, 15

autocatalysis, 87

equilibria, 60

Beer-Lambert's law and multivariate linear

introduction, 7

regression, 144, 145, 146

linear regression, 111, 125

Belousov-Zhabotinsky (BZ), 95

linest, 125

chaos, 97

matrix operations, 11

citric acid titration, 68

multivariate linear regression, 146

Cu/en/H+, 45

polynomial fitting, 125

distorted Gaussian, 38

pseudo-inverse, 146

EDTA species distribution, 66

solver, 61, 74, 207

exponential curve fitting, 150, 154

solver constraints, 61

exponential decay, 105, 205

straight line fit, 111

Fe/SCN titration, 40 gas law, 29

trendline, 111 explicit equations, 29

Gaussian curve, 37

F

general 3-component titration, 56 H3PO4 titration, 67 linearisation of exponential decay, 127 Lorenz attractor, 97 Lotka-Volterra, 92 metal/ligand titration, 60 multivariate chromatogram, 219, 241, 253, 261, 272, 282 overlapping Gaussians, 158, 161, 207 PCR and PLS of corn data, 295 polynomial fitting, 124 predator-prey, 92 reaction 2A→B, 79, 80 reaction A→B, 78, 224, 233, 244 reaction A→B→C, 143, 162, 165, 182, 206,

Factor Analysis, 213 geometrical interpretations, 224 closure, 239 HELP plots, 241 Lawton-Sylvestre, 231 mean centring, 239 noise reduction, 243 reduction in the number of dimensions, 228 three and more components, 235 two components, 224 number of significant factors, 224 feasible regions. See model-free analyses

209, 254, 255, 256, 288

G

reaction A+B↔C, 185 solubility product, 31, 62 spectrophotometric metal/ligand titration, 177

Gaussian curves, 36 distorted, 38 linear combination of, 36 global analysis, 183

324

Index

augmented data matrices, 184

multivariate, 139 numerical difficulties, 120

H hard-modelling. See model-based analyses Hessian matrix, 202 Heuristic Evolving Latent Projections (HELP), 241

polynomials, 114, 124 pseudo-inverse, 117, 118, 122, 140 standard deviation of the residuals, 122 straight line fit, 109 using Excel, 111 using Matlab, 121

I

linearisation of non-linear problems non-uniform error distribution, 130

ionic strength, 44

loadings, 215

Iterative Target Transform Factor Analysis target testing, 251

M K

Matlab / \ operators, 48, 109, 117, 118, 121, 156,

kinetics, 32, 76, Also see examples

165

boundary conditions, 77

cell arrays, 169

chemical model, 77

element-wise operations, 15, 19

complex mechanisms, 80

introduction, 7

concentration profiles, 78

linear regression, 117, 121

Euler method, 80

multivariate linear regression, 144, 145

explicit solutions, 77

optimisation toolbox, 203

initial concentrations, 77

polynomial fitting, 124

mechanism, 76

pseudo-inverse, 142

numerical integration. See numerical

Singular Value Decomposition, 215

integration

structures, 169

ordinary differential equations (ODEs), 77 oscillating reaction, 95

symbolic toolbox, 79 matrix

rate constants, 77

addition and subtraction, 12

rate law, 76

diagonal, 22

Runge-Kutta, 82

dimension, 8

L

identity, 23 inverse, 24

law of mass action, 31, 40, 44

multiplication, 16

Lawton-Sylvestre, 231

notation, 8

least-squares methods, 102

operator, 11

ligand, 45

orthogonal, 25

linear dependence, 119

orthonormal, 25

concentration profiles, 217

pseudo-inverse, 49

in concentration profiles, 175

rank, 119

linear least-squares. See linear regression

size, 8

linear regression, 109, 163

square, 21

applications, 127

submatrix, 9

design matrix, 115

symmetric, 22

errors in parameters, 121

transposition, 10

generalised matrix notation, 114

mean centring, 239

linearisation of non-linear problems, 127

metal ion, 45

matrix notation, 113

model-based analyses, 101

325

Index best model, 101, 197 model-free analyses, 213 feasible regions, 234, 288

multivariate, 162 Newton-Gauss algorithm. See NewtonGauss algorithm standard deviation of the residuals, 161,

limitations, 234 multiplicative ambiguity, 291 rotational ambiguity, 234, 288 multiplicative ambiguity. See model-free

165, 180, 189 non-white noise, 189 normal equations, 115 normalisation

analyses multivariate data, 34, 139, 162

unity hight of concentration profiles, 275 number of species, 217

N Newton-Gauss algorithm, 148, 290 calculation of the residuals, 163 convergence criterion, 153 curvature matrix, 161, 202 explicit derivatives, 151

numerical integration accuracy, 97 Euler, 80 Runge-Kutta, 82, 88, 93 step size, 86 stiff problems, 86, 90, 97

fitting in reduced eigenvector space, 180

O

fixing parameters, 169 flow diagram, 157 Hessian matrix, 202 initial guesses, 148

optimisation Newton-Gauss algorithm. See NewtonGauss algorithm

Jacobian, 149, 163

simplex, 204, See Simplex optimisation

known spectra, 175, 177

solver. See Excel

Levenberg/Marquardt extension, 155

P

Marquardt parameter, 156, 161 minimal algorithm, 149 numerical derivatives, 153, 173 separation of linear and non-linear parameters, 163 termination criterion, 153 uncolored species, 175, 177 Newton-Raphson algorithm, 48, 69 equilibrium model, 53 flow diagram, 52 initial guesses, 49, 70 input arguments, 53 Jacobian, 50, 74 numerical accuracy, 56 output arguments, 54 shift vector, 50 termination criterion, 56 noise reduction, 243 non-linear least-squares. See non-linear regression non-linear regression, 148 errors in parameters, 161 initial guesses, 148 Jacobian, 149

parameters linear, 105 Partial Least Squares (PLS), 295, 306 calibration, 295, 308 comparing with PCR, 310 mean-centring and normalisation, 307 prediction, 309 prognostic vector, 309 physical/chemical models, 29 polynomial fitting, 114, 124 Savitzky-Golay filter, 130 using Excel, 125 using Matlab, 124 polynomial interpolation, 138 lolipop, 138 Principal Component Regression (PCR), 295, 296 calibration, 295 cross validation, 303 mean-centring and normalisation, 297 prediction, 300, 302 PRESS, 305

326

Index

prognostic vector, 300

sum of squares, 103, 109, 140

principal components, 218, 221

systems of linear equations, 26

pseudo-inverse, 117, 140, 142

systems of non-linear equations, 69

using Excel, 146

T R

Target Factor Analysis, 246 projection matrices, 250

rank deficiency of concentration profiles, 175, 184 rank of a matrix, 119, 120, 217 residuals, 34, 103, 140

target testing, 247, 250, 251 parameter fitting, 257 Target Transform Search/Fit, 253

standard deviation, 223

parameter fitting via target testing, 257

structure, 222

system of homogeneous differential

Resolving Factor Analysis (RFA), 290 rotational ambiguity. See model-free analyses

equations, 254 transformation matrix, 254 Taylor series, 48, 80, 149, 199

S Savitzky-Golay filter, 130 derivative of a curve using, 135 smoothing using, 131, 132 scalar, 8 scores, 215 simplex optimisation, 204 Singular Value Decomposition (SVD), 181, 214, 260, 268 singular values, 181, 215

titration, 40, Also see examples acetic acid, 58 acid-base, 40 complexometric, 40 equilibrium constants, 40 Fe/SCN, 40 general 3-component titration, 56 metal/ligand titration, 60 pH, 40 polyprotic acids, 64

magnitude, 219

V

significant, 218 soft-modelling. See model-free analyses

van der Waals coefficients, 30

straight line fit

vector, 8

explicit equations, 111

dimension, 8

matrix notation, 113

scalar product, 17