Tutorial on the fitting of kinetics model to multivariate spectroscopic measurements with nonlinear...

download Tutorial on the fitting of kinetics model to multivariate spectroscopic measurements with nonlinear least-squares regression

of 7

Transcript of Tutorial on the fitting of kinetics model to multivariate spectroscopic measurements with nonlinear...

  • 8/9/2019 Tutorial on the fitting of kinetics model to multivariate spectroscopic measurements with nonlinear least-squares re…

    1/16

  • 8/9/2019 Tutorial on the fitting of kinetics model to multivariate spectroscopic measurements with nonlinear least-squares re…

    2/16

    reacting species. This enables structural interpretation even for 

    intermediates that only transiently exist during the reaction. (b)

    Multivariate data allow the application of a wide range of 

    model-free analyses, from simple factor analysis [19] to indicate

    the number of reaction species to sophisticated complete

    analyses based on evolving factor analysis   [20]   or the

    alternating least-squares algorithm   [21]. These model-freetechniques can also be combined with the hard-modelling

    techniques described in this tutorial to give the benefits of both

    approaches   [22,23]. (c) The need to determine a   “good”

    wavelength to follow the reaction is eliminated. (d) The analysis

    of multiwavelength data is often significantly more robust. The

    disadvantages of multiwavelength data include the large

    number of data that are acquired in a short time and the large

    number of parameters that need to be fitted. Readily available

     personal computers with large memory solve the first problem;

    appropriate algorithms solve the second.

    The most recent development, and not yet widely

    accepted, is the globalisation of the analysis of several setsof kinetic data. The investigation of reasonably complex

    mechanisms requires the measurement of data acquired under 

    different conditions, e.g., variation of initial concentrations.

    Global analysis of the complete collection of data invariably

    results in increased robustness, reduced model ambiguity and

    also significantly less operator input as only one analysis has

    to be performed.

    All of the algorithms and analyses described in this tutorial

    were completed using the commercially available software

     package Pro-KII  [24] or in-house Matlab® [25] software. Other 

    software packages offering some similar features have also been

    reported in the literature or are available commercially such as

    OPKINE  [26], KINAJDC (MW)  [27,28]  and SPECFIT/32™

    [29].

    2. Multivariate absorption data and Beer–Lambert's law

    Light absorption spectroscopy is a common technique used

    for chemical reaction monitoring, with the ultraviolet-visible(UV–Vis, practical range of 190–700 nm) and near-infrared

    (near-IR, range of 700–3000 nm) wavelength regions com-

    monly used. While measurements in the mid-IR (3000–12,500

    nm) are less commonly used for kinetic studies, the use of this

    region is increasing with the development of faster scanning

    instruments. Any of the above wavelength ranges can be used to

    follow the time-dependent spectral changes during chemical

    reactions and for the subsequent determination of reaction

    mechanisms, its kinetic parameters and the spectra of the

    absorbing reacting species. These spectra can be used evaluate

    the structure of intermediates and products. Instruments for 

    making absorbance measurements are widely available with bench-top spectrophotometers now being sold covering a large

     portion of the UV–Vis to near-IR wavelength range in a single

    device. Generally, experiments are made using stopped-flow

    techniques for rapid reaction kinetics (milliseconds–minutes),

    and manually mixed or batch reactor measurements for slower 

    kinetics (minutes–hours).

    To follow the evolution of a reaction, spectra are acquired as

    a function of the reaction time. This type of measurement results

    in data that can be arranged into a two-dimensional matrix

    which will be named  Y. If each measured spectrum is arranged

    as a row, then the resulting row dimension is the number of 

    measured spectra (nt   rows) and the column dimension is the

    Fig. 1. Structure of multivariate absorbance data arranged into a matrix  Y.

    150   G. Puxty et al. / Chemometrics and Intelligent Laboratory Systems 81 (2006) 149 – 164

  • 8/9/2019 Tutorial on the fitting of kinetics model to multivariate spectroscopic measurements with nonlinear least-squares re…

    3/16

    number of measured wavelengths (nλ  columns). This arrange-

    ment is shown in Fig. 1.

    According to Beer –Lambert's law the absorbance of a

    solution at one particular wavelength  λ  at time   t  is the sum of 

    the contributions from all absorbing species. Each element in  Y

    represents this sum for a specific time  t  and wavelength λ. The

    contribution from each species to the absorbance is linearly proportional to the concentration and the optical path length (the

    distance travelled through the solution by the incident light 

     beam). If the path length and proportionality constant are

    expressed by the single constant  εi(λ) then Beer –Lambert's law

    and each element in Y is represented by Eq. (1) (for  nc absorbing

    species).

     yðt ;kÞ ¼ c1ðt Þe1ðkÞ þ c2ðt Þe2ðkÞ þ : : : þ cncðt ÞencðkÞ¼

    Xnci¼1

    ciðt ÞeiðkÞ ð1Þ

    The structure of Eq. (1) is such that when spectra arearranged according to the matrix Y, Beer –Lambert's law can be

    elegantly expressed using matrix notation [4].

    Y ¼ CA þ R    ð2ÞThis equation is represented graphically in   Fig. 2. The

    columns of the matrix  C  contain the concentration profiles ci of 

    the nc absorbing species at the nt  measurement times. The rows

    of the matrix  A  contain the molar absorptivities  εi(λ) for each

    species at the nλ measured wavelengths. These will be referred

    to as pure component spectra. Because measurements are never 

     perfect, the matrix  Y  will contain some measurement noise and

    as such cannot be perfectly represented by the product of  C  andA. This difference is captured in the matrix of residuals   R ,

    which shares the same dimensions as   Y. The fact that each

    element in Y is an application of Beer –Lambert's law of Eq. (1)

    is illustrated by the shading of  Fig. 2. Each element in  Y  is the

    vector product of the corresponding row in  C  and column in  A

     plus the noise component in the matrix   R . Also shown are

    typical plots of the type of data contained in each matrix. They

    were produced by simulation of simple first-order kinetics.

    3. Calculation of kinetic concentration profiles

    The central step in the fitting of a kinetic model to

    multivariate absorbance data is being able to calculate the

    concentration profiles of all the species involved in a chemical

    reaction. The concentration profile of a species is its change in

    concentration with time. The concentration profiles of all species

    are arranged into column vectors and they correspond to the

    columns of the matrix  C  in  Fig. 2. According to kinetic theory,

    the concentration profiles of the species in a reaction mechanism

    are defined by a system of ordinary differential equations

    (ODEs) [30]. With knowledge of the initial conditions, ODEs

    can be integrated to any time point, allowing the calculation of the concentration profiles corresponding to data acquisition

    times.

    For a number of simple first- and second-order reactions, it is

     possible to explicitly integrate the resulting system of ODEs.

    For example, consider the reaction mechanism of Eq. (3).

    Shown is the system of ODEs and their integrated form for this

    second-order mechanism.

    2AYk 2A

    B

    d½Adt   ¼

    −2k 2A

    ½A

    2

    ;

    d½Bdt   ¼

     k 2A

    ½A

    2

    A½ ¼   ½A01 þ 2½A0k 2At 

    Fig. 2. The matrix  Y  of multivariate absorbance data expressed using Beer –Lambert's law. A plot of typical data is given below each matrix.

    151G. Puxty et al. / Chemometrics and Intelligent Laboratory Systems 81 (2006) 149 – 164

  • 8/9/2019 Tutorial on the fitting of kinetics model to multivariate spectroscopic measurements with nonlinear least-squares re…

    4/16

    B½ ¼ ½B0 þ ½A0−½A

    2  ð3Þ

    For the majority of multiple step reaction mechanisms, it is

    not possible to explicitly integrate the resulting systems of 

    ODEs. To overcome this, numerical integration is used.

     Numerical integration allows an approximation to the explicit 

    solution to be calculated for any system of ODEs, within limits

    of numerical accuracy and computation time  [31]. Numerical

    integration routines are, at least in basic principle, based on

    Euler's method [31]. Euler's method uses a form of the Taylor 

    series expansion truncated to the first derivative (see the

    Appendix for details of the Taylor series expansion).

    The principle of Euler's method is represented graphically in

    Fig. 3. Starting with the concentration at time   t , the

    concentration at  t +Δt  is estimated by moving along the tangent 

    (the derivative  dciðt Þ

    dt   ) at  t . It can be seen that the accuracy of the

    approximation is dependent on the magnitude of the increment 

    size Δt  and the shape of the function. The greater the curvatureand the steeper the tangent, the smaller the increment size that 

    must be used to give an accurate approximation.

    Eq. (4) formalises the graph and Eq. (5) shows the equation

    as would be applied to calculate the concentration profiles of 

    species A and B from Eq. (3).

    ciðt  þ Dt Þcciðt Þ þ dciðt Þdt 

      Dt    ð4Þ

    ½At þDt c½At −2k 2A½A2t Dt ; ½Bt þDt c½Bt  þ k 2A½A2t Dt    ð5Þ

    In practice, Euler's method as described here is not used in

    modern numerical integration routines due to its lack of accuracy. Some of the modern strategies include using higher-

    order terms from the Taylor series expansion or calculating the

    first derivative at multiple points within an increment. Adaptive

    increment size control is also used to achieve a specified level of 

    accuracy. The most robust of the modern numerical integration

    routines use the idea of extrapolating from a particular result to

    the value that would have been obtained if a much smaller,

    ideally infinitesimally small, increment size were used [31].

    A system of ODEs can be considered either non-stiff or stiff 

    depending upon the relative difference in the rate of change of 

    the concentration profiles. If the rate of change of each

    concentration profile is similar, the system of ODEs is said to

     be non-stiff. For non-stiff systems of ODEs, the fourth-order 

    Runge–Kutta method   [31]   is the   “workhorse”   of numerical

    integration and it was used in this tutorial. It is called fourth

    order because the first derivative is calculated at four points

    along the increment of integration, allowing much larger 

    increments and thus dramatically reduced computation timescompared to the simple Euler method. Alternatively, if one

    concentration profile changes rapidly while another does not,

    the system of ODEs is said to be stiff. Such a situation can occur 

    if either rate constants or concentrations vary by orders of 

    magnitude or the mechanism is a mixture of elementary steps

    with different reaction orders. In this case, if a traditional

    method is used the increment size must be reduced to such a

    small value to yield accurate integration of the rapidly changing

    concentration profiles, that the computation time becomes

    excessive for the more slowly changing concentration profiles.

    A modified approach to carrying out the integration is then

    required. For stiff systems the Bulirsch–Stoer method using

    semi-explicit extrapolation [31] was used.To carry out numerical integration, it is first necessary to

    construct the system of ODEs that represent a given reaction

    mechanism. This can be done manually, however such an

    approach is both error-prone and cumbersome for mechanisms

    of any complexity. Rather, an automated method commonly

    used to construct ODEs from chemical equations can be adopted

    [32]. Firstly, the number of species present in the reaction

    mechanism and the stoichiometry of each reactant and product is

    determined. Consider the reaction mechanism of Eq. (6).

    A

     þ B Y

    k AB

    C

    2CYk 2C

    D   ð6ÞTwo matrices are then constructed; one containing the

    reactant stoichiometries   Xr , and one containing the product 

    stoichiometries   X p. Each matrix has  ns   columns, representing

    all the reacting species in the mechanism (A, B, C and D as

    distinct from the number of absorbing species  nc≤ns), and  np

    rows, representing reactions or rate constants (k BA and k 2C).  Xr 

    contains the stoichiometry of each species as a reactant and  X pcontains the stoichiometry of each species as a product. If  Xr  is

    subtracted from   X p, the result is the matrix of stoichiometric

    coefficients  X  (see Fig. 4).Fig. 3. Application of Euler's method to the calculation of a concentration profile.

    Fig. 4. Reactant and product matrices for the reaction mechanism of Eq. (6).

    152   G. Puxty et al. / Chemometrics and Intelligent Laboratory Systems 81 (2006) 149 – 164

  • 8/9/2019 Tutorial on the fitting of kinetics model to multivariate spectroscopic measurements with nonlinear least-squares re…

    5/16

    Once   Xr    and   X   have been determined, the differential

    equations can be constructed in a completely generalised way,

    according to Eq. (7). The expression defining v   j  is the rate law of 

    the j -th elementary step in the mechanism [30].

    v  j  ¼ k  j Ynsi¼1

    c xrj ;ii   where   j  ¼ 1 to  np

    dci

    dt  ¼

    Xnp j ¼1

     x j ;iv  j    where   j  ¼ 1 to   ns   ð7Þ

    The ODEs that result from the application of Eq. (7) to thereaction mechanism of Eq. (6) are given by Eq. (8). The result of 

    integrating the ODEs in Eq. (8) with [A]0= 1, [B]0=0.8,  k BA= 2

    and k 2C= 1 from 0 to 10 time units in 100 increments is shown in

    Fig. 5.

    v 1 ¼ k AB ½A1½B1½C0½D0

    v 2 ¼ k 2C½A0½B0½C2½D0

    d½Adt 

      ¼ d½Bdt 

      ¼ 1v 1 þ 0v 2

    d½Cdt 

      ¼ 1v 1−2v 2

    d½Ddt 

      ¼ 0v 1 þ 1v 2   ð8Þ

    4. Linear and non-linear least-squares regression

    4.1. Linear parameters

    Once the concentration profiles have been calculated, the

     pure component spectra can be calculated in a single step. This

    is because they are linear parameters, and as such, there is no

    need to pass them through the non-linear optimisation routine[4,33].

    Fig. 5. Integrated concentration profiles for the mechanism of Eq. (6) using the

    ODEs of Eq. (8).

    Fig. 6. Flow diagram of the Newton–Gauss–Levenberg/Marquardt (NGL/M) method.

    153G. Puxty et al. / Chemometrics and Intelligent Laboratory Systems 81 (2006) 149 – 164

  • 8/9/2019 Tutorial on the fitting of kinetics model to multivariate spectroscopic measurements with nonlinear least-squares re…

    6/16

    Recall that a multivariate spectroscopic measurement   Y

    can be expressed as the product of  C  and   A  plus a matrix of 

    residuals (Eq. (2)). This is a system of linear equations and,

    for such systems, there is an explicit least-squares solution

    [34]. Given   Y   and   C, the best estimate for   A   can be

    calculated as shown in Eq. (9).   C+ is called the pseudo-

    inverse of  C. It is necessary to use the pseudo-inverse as  C   isnot a square matrix precluding direct calculation of its

    inverse.

    Cþ ¼ ðCTCÞ−1CT

    A ¼ CþY   ð9Þ

    C+ can be calculated as shown in Eq. (9); this, however, is

    not recommended as there are numerically superior methods

    [31].

    4.2. Non-linear parameters

    The non-linear parameters to be fitted are the rate

    constants of the reaction mechanism that define the matrix

    of concentration profiles   C. The rate constants are refined

    so as to minimize the sum of squares of the residuals

    matrix of Eq. (2) ( ssq, the sum over the all elements in

    R 2). The parameters are non-linear because the relationship

     between the parameters and the residuals is not linear. A

    vast range of non-linear regression algorithms exists. One of 

    the most commonly chosen methods of non-linear regres-

    sion is the Newton–Gauss–Levenberg/Marquardt (NGL/M)method   [4,35,36], and it will be described in detail here.

    This method is a gradient method, which means it relies on

    calculation of the derivative of the function being optimised

    (the residuals). Methods also exist that do not rely on

    calculation of the derivative, such as the Simplex method

    [37]   or genetic algorithms   [38]. In general, the convergence

    of gradient methods is superior to other methods if the

    initial parameter estimates lie in the region of the global

    optimum.   Fig. 6   is a flow diagram showing the NGL/M

    method. The NGL/M method is able to vary smoothly

     between an inverse Hessian method and a linear decent 

    method. It evolved in essentially two stages. The inverseHessian Newton–Gauss method was developed first; this

    will be the starting point for a description of the NGL/M

    method.

    4.2.1. Newton – Gauss method 

    The first step in any gradient method is to define initial

    estimates for the non-linear parameters, the rate constants, to

     be refined. The next step is to evaluate the target function to

     be minimised, the   ssq   over the residuals. To do this, the

    concentration profiles must be calculated according to

    Section 3 using the initial rate constants. The linear 

     parameters, the pure component spectra, can then be

    calculated by Eq. (9). This gives the matrices   C   and   A.

    The residuals matrix and its   ssq   can now be calculated by

    rearranging Eq. (2) to yield Eq. (10).

    R  ¼ Y−CA

     ssq ¼ X

    nt 

    i¼1X

    nk

     j ¼1r 2

    i;

     j    ð10

    ÞOnce the   ssq   has been determined, the next step is to

    calculate a shift in the non-linear parameters in such a way that 

    the ssq moves towards its minimum value. To do this, it needs to

     be emphasised that  R , and subsequently the ssq, are functions of 

    the non-linear parameters only. By substituting Eq. (9) into Eq.

    (10),  R  can be written as a function of  C  only, which is itself a

    function of the rate constants. If the non-linear parameters to be

    fitted,  p1  to  pnp, are arranged into a vector  p = ( p1,  p2,  …,  pnp),

    this relationship is given by Eq. (11).

    ðp

    Þ ¼ Y−CCþY

      ð11

    ÞIf the initial parameter estimates are given by the vector  p0,

    the Taylor series expansion can be used to estimate  R  following

    a small shift in the parameters Δ p = (Δ p1, Δ p2, …, Δ pnp) (see the

    Appendix for details of the Taylor series expansion). If only the

    first derivative of  R  is used a linear expression results and is

    given by Eq. (12).

    R ðp0 þ DpÞcR ðp0Þ þAR ðp0Þ

    A p1D p1 þAR ðp0Þ

    A p2D p2

    þ: : : þ AR ðp0ÞA pnp

    D pnp   ð12Þ

    While this is a crude approximation, the fact that it is a linear expression makes it easy to deal with. The goal is to determine

    the vector of parameter shifts that moves   R (p0+Δp) towards

    zero. So, if   R (p0+Δp) is replaced by zero and Eq. (12) is

    rearranged, Eq. (13) results.

    R ðp0Þc−AR ðp0Þ

    A p1D p1−

    AR ðp0ÞA p2

    D p2− : : : −AR ðp0ÞA pnp

    D pnp

    ð13ÞR (p0) is calculated as described and the partial derivative

    AR ðp0ÞA pi

    can be calculated by the method of finite differencing

    according to Eq. (14). To calculate the partial derivative for 

     parameter   pi,   pi   is shifted by a small amount   Δ pi   and the

    residuals are calculated to yield   R (p0+Δ pi).  AR ðp0Þ

    A piis then

    calculated by subtracting R (p0+Δ pi) from  R (p0) and dividing

    Δ pi  (equivalent to calculating the tangent in two dimensions).

    AR ðp0ÞA pi

    cR ðp0 þD piÞ−R ðp0Þ

    D pið14Þ

    In its current form, it is not clear how Δp can be calculated

    using Eq. (13) as it is not a single matrix–vector product. One

    solution to this problem is to vectorise (unfold into long column

    vectors) the residuals and partial derivative matrices. This

    expression can then be easily collapsed into a matrix–vector 

     product. This procedure is illustrated graphically in Fig. 7. The

    154   G. Puxty et al. / Chemometrics and Intelligent Laboratory Systems 81 (2006) 149 – 164

  • 8/9/2019 Tutorial on the fitting of kinetics model to multivariate spectroscopic measurements with nonlinear least-squares re…

    7/16

    matrix of vectorised AR ðp0Þ

    A pi

    matrices is called the Jacobian J  and

    results in Eq. (15).

    rðp0Þ ¼ −JDp   ð15ÞThis equation is now in a form that has a structure that can be

    solved by the linear least-squares method for Δp. This solution

    is given in Eq. (16).

    Jþ ¼ ðJTJÞ−1JT

    Dp ¼ −Jþrðp0Þ ð16Þ

    Since both the truncated Taylor series expansion and the

     partial derivatives used are only approximations, the calculated

     parameter shifts of Eq. (13) will not be perfect. Thus, aniterative procedure is adopted, where the approximations are

    successively improved. The calculated shifts are applied to the

     parameters (p0,  j +1= p0,  j +Δp  going from the   j -th to the   j +1-th

    iteration) and the process of calculating   C   through to  Δp   is

    repeated. This process is iterated until the relative change in the

     ssq   from one iteration to the next falls below some threshold

    value ( ssqold≈ ssq). A relative change of less then 10−4 is

    generally appropriate.

    4.4. The Marquardt modification

    Generally, the Newton–Gauss method, as described so far,

    converges rapidly, quadratically near the minimum. However, if 

    the initial estimates are poor, the functional approximation by

    the Taylor series expansion and the linearization of the problem

     becomes invalid. This can lead to divergence of the   ssq   and

    failure of the algorithm.

    The modification suggested by Marquardt  [36], based on the

    ideas of Levenberg   [35], was to add a certain number, the

    Marquardt parameter   mp, to the diagonal elements of theHessian matrix, H = JTJ, during the calculation of the parameter 

    shifts, as shown in Eq. (17).

    H ¼ JTJ

    Dp ¼ −ðH þ mp  IÞ−1JTrðp0Þ

    where   I ¼ the identity matrix   ð17ÞWhen the value of mp is significantly larger than the elements

    of  H, the expression  H + mp × I   becomes diagonally dominant.

    This means the inverse is effectively a diagonal matrix, with   1

    mpin the diagonal elements. This causes Eq. (17) to collapse to Eq.

    (18), which has the form of the linear descent method.

    Dp ¼   1mp

      I

    JTrðp0Þ ð18Þ

    When the value of  mp is small, Eq. (17) reverts to that of the

    inverse Hessian method. The Marquardt parameter is initially set 

    to zero. There are many strategies to manage the Marquardt 

     parameter, ours is the following. If divergence of the ssq occurs,

    then the Marquardt parameter is introduced (given a value of 1)

    and increased (multiplication by 10 per iteration) until the  ssq

     begins to converge. Once the ssq converges the magnitude of the

    Marquardt parameter is reduced (division by ffiffiffiffiffi

    10p    per iteration)and eventually set to zero when the break criterion is reached.

    4.5. Error estimates and correlation coefficients

    A spin-off from the NGL/M algorithm is that it allows

    direct estimation of the errors in the non-linear parameters.

    The inverted Hessian matrix   H−1, without the Marquardt 

     parameter added, is the variance–covariance matrix of the

     parameters. The diagonal elements contain information on the

     parameter variances and the off-diagonal elements the

    covariances. The formula for the standard error    σi   in

     parameter   pi   is given by Eq. (19).

    ri ¼  rY ffiffiffiffiffiffiffi

    h−1i; j 

    q   ð19Þ

    h− 1i,i   is the   i-th diagonal element of the inverted Hessian

    matrix   H− 1 and   σY   is the standard deviation of the residuals

    R   (see Eq. (20)).

    rY ¼ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

     ssq

    nt   nk−ðnk þ nc  nk Þr 

      ð20Þ

    The denominator of  σY is the number of degrees of freedom.

    This equals the number of experimental values, the number of 

    elements in  R  (nt × nλ), minus the number of fitted parameters

    Fig. 7. Vectorising and collapsing Eq. (13) into a matrix and vector product  [39].

    155G. Puxty et al. / Chemometrics and Intelligent Laboratory Systems 81 (2006) 149 – 164

  • 8/9/2019 Tutorial on the fitting of kinetics model to multivariate spectroscopic measurements with nonlinear least-squares re…

    8/16

    (np + nc × nλ, that is the number non-linear and linear para-meters). If  H  is normalised to one in the diagonal elements, this

    yields the correlation coefficients between parameters [40]. The

    element   hi,  j   of the normalised   H   is the correlation coefficient 

     between parameter i and j . This is the cosine of the angle between

    the columns in J for these parameters. The closer the value of the

    correlation coefficient to one (the cosine of 0°) the more

    correlated the parameters. High correlation between parameters

    means that the two parameters cannot be distinguished from one

    another and one needs to be removed from the fit.

    5. Second-order global analysis

    Up to this point, the analyses discussed have involved a

    single multivariate measurement. This approach has been

    dubbed first-order global analysis  [1,4]. A powerful extension

    of this approach is second-order global analysis   [2,41]. In

    second-order global analysis, the procedures already described

    are applied simultaneously to multiple measurements of the

    same system made under different conditions. Typically,

    measurements where the initial concentrations have been

    varied. Thus, a global model is simultaneously fitted to multiple

    measurements. This has a number of advantages, including

    more robust determination of the parameters, determination of 

     parameters that are not defined in any single measurement andthe breaking of linear dependencies  [2,42]. Also, by fitting a

    global model to multiple measurements made under different 

    conditions, more confidence can be had in the model.

    A convenient way of organising a second-order global

    analysis process is to organise the data sets in such a way that 

    existing matrix based software can be adapted. If   nm

    multivariate measurements have been made, the first step for 

    carrying out second-order global analysis is to concatenate

    (stack one atop the other) the data matrices to form one large

    matrix   Ytot . It is assumed that each measurement covers the

    same wavelength range, and thus, each data matrix has the same

    number of columns. The individual concentration profile

    matrices are then concatenated in a similar manner to the data

    matrices, to form one large matrix   Ctot . Although multiple

    multivariate measurements have been concatenated, because

    each measurement obeys the same reaction mechanism, Eq. (2)

    can still be applied using  Ytot  and  Ctot  (see Eq. (21)).

    Ytot  ¼ Ctot A þ R tot    ð21ÞThis equation is represented in Fig. 8. This means calculation of 

    the linear parameters, and fitting of the non-linear parameters,

    can be carried out in the same manner as previously described

     by replacing  Y  with  Ytot  and  C  with  Ctot .

    Second-order global analysis is at its most powerful when a

    single matrix of pure component spectra,  A, can be calculated

    for  Ytot  and Ctot , as shown in Fig. 8. This is known as the global

    spectra mode of analysis. It is not always possible however. If 

    there is baseline drift, or some other inconsistency between the

    measurements such as temperature-dependent spectra, a single

    A matrix cannot be used. This is because all measurements will

    no longer share the same pure component spectra. In this

    situation, a separate   A  matrix is calculated for each measure-ment. This arrangement is shown in Fig. 9, and is known as the

    local spectra mode of analysis.

    To illustrate the power of second-order global analysis,

    consider the second-order reaction Eq. (22) where the concen-

    tration profiles for a single measurement are linearly dependent.

    A þ B   Yk A

    B ¼1

    C   ð22ÞFor a single multivariate measurement of this mechanism,

    starting with the initial concentrations [A]0   and [B]0, the

    concentration profiles of the species A, B and C are linearly

    dependent. The smallest number of linearly independent termsthat can be used to represent the data is called the rank of the data.

    The rank of the data can be estimated by carrying out factor 

    analysis [19] and in this case the rank of C and subsequently Y is

    two and the calculation of C+ and thus A, Eq. (9), is not possible.

    However, if a second measurement of the same reaction, having

    different initial concentrations for A and B, was also made the

    linear dependence can be broken. By fitting the mechanism

    Fig. 8. Application of second order global analysis using a single matrix of pure

    component spectra  A  (global mode).

    Fig. 9. Application of second order global analysis where an individual  A i  iscalculated for each measurement (local mode).

    156   G. Puxty et al. / Chemometrics and Intelligent Laboratory Systems 81 (2006) 149 – 164

  • 8/9/2019 Tutorial on the fitting of kinetics model to multivariate spectroscopic measurements with nonlinear least-squares re…

    9/16

    using second-order global analysis to both measurements, using

    a single A matrix, the pure component spectrum of all the species

    can be resolved. The rank of the resulting combined data matrix

    Ytot  is three as it can no longer be represented by two linearly

    independent terms.

    This is illustrated in   Fig. 10, which shows simulated

    concentration profiles and data for the mechanism of Eq. (22),with one set of initial concentrations in part (a) and with two

    different sets and second-order global analysis in part (b). Both

    concentration profiles have [A]0=0.5 and for the first [B]0=0.2

    and the second [B]0=0.4. A single Gaussian peak was used for 

    the pure component spectrum of each species and the

    absorbance data was calculated as   Y = CA. If one single data

    file and matrix of concentration profiles are used to calculate the

     pure component spectra according to Eq. (9) ((a) of  Fig. 10), the

     pure component spectra cannot be resolved. However, if both

    data files and concentration profiles are used according to

    second-order global analysis ((b) of  Fig. 10), the spectra of all

    three species can be calculated and are well defined.

    6. Examples

    In this section a number of examples are given where the

    techniques described have been applied, first to a complex

    simulated example, and then to three real examples. Each

    example will be used to highlight particular benefits of the

    global analysis of multivariate data. This tutorial does not deal

    directly with how to discriminate between different models.

    However, in summary, the basic approach adopted is that the

    residuals are used as an indication of the validity of the

    Fig. 10. Calculation of pure component spectra using (a) one single measurement and (b) two measurements and second order global analysis using global spectra.

    157G. Puxty et al. / Chemometrics and Intelligent Laboratory Systems 81 (2006) 149 – 164

  • 8/9/2019 Tutorial on the fitting of kinetics model to multivariate spectroscopic measurements with nonlinear least-squares re…

    10/16

    Fig. 11. Reaction scheme used for the chlorination of benzene. The chlorination reagent is omitted.

  • 8/9/2019 Tutorial on the fitting of kinetics model to multivariate spectroscopic measurements with nonlinear least-squares re…

    11/16

    underlying model. In all cases, the simplest model (in terms of 

    number of parameters) that yields residuals with a noise level

    and error structure expected from the instrumentation is chosen.

    The identifiabilty of model parameters is estimated based on the

    error values and correlation coefficients calculated during the

    non-linear regression. If the estimated error in a parameter is

    large or two parameters are highly correlated then this is used toindicate that the model is incorrect or overly complex.

    Convergence of the model parameters to the same values

    from different starting guesses is also used as an indication of 

    model correctness and parameter identifiability. For a discus-

    sion of this topic, see Vajda and Rabitz [43].

    6.1. Simulated chlorination of benzene

    As a first example of the power of fitting kinetic models to

    multivariate data a simulation will be considered. The

    chlorination of benzene to hexachlorobenzene involves a

    complex series of reactions. The scheme is represented in Fig.11. To highlight the complexity of the mechanism, consider the

    third chlorination step. Three dichloro isomers react to produce

    three trichloro isomers. 1,2-dichlorobenzene reacts only to the

    1,2,3- and 1,2,4-trichlorobenzenes; 1,3-dichlorobenzene reacts

    to form all three isomers and 1,4-dichlorobenzene reacts only to

    form 1,2,4-trichlorobenzene. Data were generated for the

    mechanism by assigning hypothetical rate constants defined

     by the probability of the reaction given by the number of 

    substitution possibilities that leads to a particular product. For 

    example, for 1,3-dichlorobenzene to 1,2,4-trichlorobenzene

    there are two possibilities so  k 1,3–1,2,4  was assigned a value of 

    2 and for 1,3,5-trichlorobenzene to 1,2,3,5-tetrachlorobenzene

    there are three possibilities so   k 1,3,5–1,2,3,5=3. This modelignores electronic effects of the substituents which in reality

    slow down additional substitutions. This is not relevant for the

     present purpose. There are a total of 13 species and 20 rate

    constants in the mechanism.

    Data were generated by modelling the pure spectra of the

     benzenes as severely overlapping Gaussian peak s. The

    chlorination reagent was assumed to be non-absorbing.

    Absorbance data was generated by multiplying simulated

    concentration profiles by the simulated pure spectra. White

    noise with a standard deviation of 0.001 (a realistic value for 

    UV–Vis spectrophotometers) was added to the data. Anexample of the generated data is shown in  Fig. 12. Due to the

    complexity of this system it cannot be resolved from a single

    measurement, even with perfect noise-free data as serious rank 

    deficiencies result and simplified models can be fit. 10

    measurements were generated, each one starting with the

    unsubstituted benzene and with one of the isomers in the steps

    involving three isomers (di-, tri- and tetra-chloro isomers). This

    was necessary to break the rank deficiencies in the concentra-

    tion profiles. A concentration of 1 with a ten-fold excess of 

    chlorination reagent was used, with 100 spectra generated at 

    equal intervals between 0 and 0.3 time units.

    All 10 measurements were analysed globally using second-order global analysis. The determined rate constants were

    essentially correct (exactly correct to two significant figures).

    Such an analysis would be practically impossible without the

    use of multivariate data and second-order global analysis. Due

    to the severely overlapped spectra, there is no single wavelength

    that could be chosen to follow a single species. Furthermore,

    due to the complexity of the mechanism, there is no single

    measurement or subset of the measurements that could be made

    to elucidate all the rate constants and spectra due to the severe

    rank deficiencies of the concentration profiles.

    6.2. Complexation of Cu(II) by cyclam using stopped-flow and 

     standard spectrometry

    In this example the reaction between Cu(II) and the

    macrocyclic ligand cyclam (1,4,8,11-tetraazacyclotetradecane)

    in aqueous solution was investigated. The details of the

    Fig. 12. Generated data for the benzene example with a 1 : 10 ratio between benzene and the chlorination reagent: (a) calculated concentration profiles and pure spectra(inset) and (b) calculated absorbance data.

    159G. Puxty et al. / Chemometrics and Intelligent Laboratory Systems 81 (2006) 149 – 164

  • 8/9/2019 Tutorial on the fitting of kinetics model to multivariate spectroscopic measurements with nonlinear least-squares re…

    12/16

    experiments and results can be found elsewhere [13]. Cyclam is

    a tetradentate ligand and forms a 1:1 complex with Cu(II) in

    aqueous solution. The reaction rate depends upon the level of 

     protonation of cyclam, and as such is strongly pH-dependent.

    Multiwavelength kinetic data of the reaction was collected

    using both stopped-flow and manually mixed measurements

    with [Cu2+]0=3.9×10−3 M and [cyclam]0=4.3×10

    −3 M and

    different starting acid concentrations (1×10−7

    −0.057 M). Thekinetic model given in Eq. (23), coupled to the protonation

    equilibria given by Eq. (24) (L-cyclam), was fitted simulta-

    neously to all the measurements. A description of the method

    used to couple the kinetic model and protonation equilibria is

     beyond the scope of this tutorial. The reader is referred to

    Maeder et al.   [13]   for details. However, beyond some

    modification to the method used to calculate the concentration

     profiles, the fitting was done as described in this tutorial.

    Cu2þ þ LHþ Yk Cu

    LH

    CuL2þ þ Hþ

    Cu2þ þ LH2þ2   Yk Cu

    LH

    2 CuL2þ þ 2Hþ   ð23Þ

    L þ HþW K 1 LHþ

    LHþ þ HþW K 2 LH2þ2

    LH2þ2  þ

     HþW K 3

    LH3þ3

    LH3þ3   þ HþW K 4

    LH4þ4   ð24ÞThe calculated concentration profiles, calculated pH as a

    function of reaction time (calculated as − log10([H+])), fits of the

    absorbance data at two wavelengths and the calculated pure

    component spectra of the absorbing species are shown in  Fig.

    13. In this case, it was necessary to calculate the pure

    component spectra in the local mode due baseline shifts

     between measurements. However, because of the multiwave-

    length data and the fact that only Cu2+ and CuL2+ absorbed the

    calculated pure component spectra were well defined for all the

    measurements. This allowed measurement to measurement comparison as well as comparison with independently deter-

    mined spectra for verification purposes. However, the real

     benefit for the study of this system comes from the fitting of a

    global reaction model to a number of measurements simulta-

    neously. All the parameters associated with this mechanism

    cannot be defined by a single measurement due to the effect of 

     pH on the reaction velocity. Only through fitting multiple

    multivariate measurements simultaneously are all the para-

    meters defined.

    6.3. Acid-induced dissociation of tris(ethylenediamine)

    nickel(II)

    In this example, the acid-induced dissociation of tris

    (ethylenediamine) nickel(II) (Ni(en)32+) in aqueous solution is

    considered. This reaction has been studied in detail   [44–47],

    and in the presence of an excess of acid the Ni(en) 32+ complex

    undergoes irreversible dissociation in the three first-order steps

    given by Eq. (25).

     NiðenÞ2þ3 YHþ

    k  NiðenÞ3 NiðH2OÞ2ðenÞ2þ2   þ en

     Ni

    ðen

    Þ2þ2 Y

    k  Niðen

    Þ2 Ni

    ðH2O

    Þ4

    ðen

    Þ2þ

     þ en

     NiðenÞ2þYHþ

    k  NiðenÞ3 NiðH2OÞ2þ6   þ en   ð25Þ

    The reaction proceeds as the bidentate en ligand becomes

     protonated and dissociates from the metal centre.

    Multiwavelength kinetic data of this reaction was measured

    using a stopped-flow spectrophotometer. Initially a 0.040 M

    solution of Ni(en)32+ was prepared by dissolving Ni2+ in the

     presence of an excess of en and 1 M sodium perchlorate. This

    solution was then mixed with a 1 M solution of perchloric acid

    in the stopped-flow and the reaction was followed between 430

    and 640 nm for 10 s. Experimental details can be found

    elsewhere [48]. The reaction mechanism of Eq. (25) was fitted

    Fig. 13. Fit results for a manually mixed measurement with [Cu2+]0=3.9×

    10−3 M, [LH22+]0=4.3×10

    −3 M and [H+]0=1.0×10−7 M. (a) Calculated

    concentration profiles and pH and (b) measured (•••) and calculated (—)

    absorbance data and calculated pure component spectra (inset).

    160   G. Puxty et al. / Chemometrics and Intelligent Laboratory Systems 81 (2006) 149 – 164

  • 8/9/2019 Tutorial on the fitting of kinetics model to multivariate spectroscopic measurements with nonlinear least-squares re…

    13/16

    to the data and   Fig. 14   shows the resulting calculated

    concentration profiles and fits at selected wavelengths.

    For this example the benefits of using multivariate data are

    significant and three-fold. Firstly, without knowledge of the

    spectrum of the intermediate species selection of a single

    wavelength to follow the reaction is difficult. Choosing a single

    wavelength below 500 nm results in k  Ni(en) being poorly definedand choosing a wavelength above 640 nm means   k  Ni(en)3 is

     poorly defined. The second advantage is related to the fact that 

    the mechanism consists of three first-order consecutive

    reactions. The parameter values determined for the mechanism

    are k  Ni(en)3 =99.0±0.3 s−1, k  Ni(en)2 =4.10± 0.01 s

    −1 and k  Ni(en)=

    0.184± 0.001 s−1. Swapping of the values of  k  Ni(en)3 and k  Ni(en)2or  k  Ni(en)2 and k  Ni(en) results in fits of identical quality. This fast-

    slow ambiguity exists for all consecutive first-order reactions

    and it has been well documented in literature [49]. However, the

    correct ordering of the rate constants is immediately apparent 

    upon examination of the pure component spectra, as can be seen

    in Fig. 15. When the wrong ordering is used, severely distorted

    and often negative spectra will result. When using single

    wavelength data, such physically impossible spectra may not 

    occur or may not be detected. For example, using 640 nm and

    swapping the values of   k  Ni(en)3 and   k  Ni(en)2 still results in all

     positive molar absorptivities and reasonable values. The last 

    advantage is that the calculated pure component spectra allow

    structural determination of the intermediate species. In this case,

    the determined spectrum of Ni(en)22+ indicates that it is in the cisform [50].

    6.4. Epoxidation of 2,5-di-tert-butyl-1,4-benzoquinone

    As a final example the epoxidation of 2,5-di-tert -butyl-1,4-

     benzoquinone (TBB) is considered. The experimental details

    can be found elsewhere [17]. In summary, TBB and tert -butyl-

    hydroperoxide (TBH) were added to a small volume (b45 mL)

    stirred and thermostated reactor. The solvent used was a mixture

    of 1,4-dioxane, ethanol and water. The reaction was initiated by

    addition of a catalyst, Triton-B (benzyltrimethylammonium

    hydroxide), in methanol. The two-step epoxidation reaction wasthen followed by IR spectroscopy using an in situ ATR probe.

    Fig. 15. Calculated pure component spectra for all species with (a) the rate

    constants in the correct order and (b) with the values of   k  Ni(en)3 and   k  Ni(en)2swapped.

    Fig. 14. (a) Calculated concentration profiles and (b) measured (•••) and

    calculated (—) absorbance data for the dissociation of Ni(en)32+. The time axis

    has been plotted with a logarithmic scale so initial fast changes are visible.

    161G. Puxty et al. / Chemometrics and Intelligent Laboratory Systems 81 (2006) 149 – 164

  • 8/9/2019 Tutorial on the fitting of kinetics model to multivariate spectroscopic measurements with nonlinear least-squares re…

    14/16

    Calorimetry data were measured simultaneously but were not 

    considered here.

    To highlight the ability of second-order global analysis to

     break rank deficiencies two measurements made at 30 °C, with

    different initial concentrations of reagents have been chosen.

    The reaction mechanism used to fit the data is shown in  Fig. 16.

    The reaction was fitted as irreversible (TBH is in excess) andthird order in each step (with the Triton-B catalyst as a reactant 

    and product) to allow inclusion of the catalyst.

    If the reaction mechanism of   Fig. 16   is fitted to a single

    measurement with TBB, TBH, monoEp and diEp set as

    absorbing species ([TBB]0=0.22 M, [TBH]0=2.2 M and

    [Triton-B]0=0.064 M), the calculated pure component spectra

    of   Fig. 17(a) result. The poor outcome is due to linear 

    dependencies amongst the concentration profiles. The rank of 

    the concentration profiles matrix is three, but four species have

     been set as absorbing. To address this, one of the species could

     be set as non-absorbing. The resulting calculated pure

    component spectrum would then represent a mixture of multiplespecies. A better alternative is to include a second measurement 

    into the analysis, with different initial concentrations

    ([TBB]0=0.24 M, [TBH]0=1.5 M and [Triton-B]0= 0.072 M),

    and calculate a global pure component spectra matrix according

    to the method of second-order global analysis. The result of 

    doing this is shown in Fig. 17(b). The linear dependence of the

    concentration profiles is now broken and the pure component 

    spectrum of all the species can be resolved.

    7. Conclusion

    The methods required to carry out the fitting of a kinetic

    chemical model to measured multivariate spectroscopic datahave been outlined in detail. From the postulation of the model

    and the derivation of the differential equations, through the

    numerical integration of the model to yield the concentration

     profiles and finally the calculation of the pure component 

    spectra and fitting of the model's rate constants to measured

    data.

    The benefits of fitting kinetic models to multivariate data

    have been explained and demonstrated by simulated and real

    examples. These benefits include: more robust model and

     parameter determination; calculation of pure component 

    spectra; the breaking of linear dependencies (second-order 

    global analysis); and elimination of the need for single

    wavelength selection and a reduction in the number of 

    measurements required for analysis. Probably the most 

    important point of all is that, particularly with the

    Fig. 17. Calculated pure component spectra with TBB, TBH, monoEp and diEp

    as coloured for (a) calculation with a single measurement and (b) calculation

    using two measurements with different initial concentrations and global spectra

    (shaded area shown in inset). NB: The absorbance data has not been divided by

    the optical path length for visual clarity so the absorptivity is in M −1 rather than

    M−1 cm−1.

    Fig. 16. Reaction mechanism for the epoxidation of TBB by TBH with Triton-B as a catalyst.

    162   G. Puxty et al. / Chemometrics and Intelligent Laboratory Systems 81 (2006) 149 – 164

  • 8/9/2019 Tutorial on the fitting of kinetics model to multivariate spectroscopic measurements with nonlinear least-squares re…

    15/16

    availability of instrumentation capable of delivering multi-

    variate data, there is no reason not to take advantage of the

    techniques that exist for their treatment. Furthermore, a

    n umber o f simulated and real examp les h ave b een

    considered to illustrate the benefits.

    An additional point to note is that although the techniques

    described in this tutorial relate directly to absorbance data, theycan be applied to any multivariate data that shows a linear 

    response to the concentration of species in the reaction

    mechanism. For example the techniques can just as easily be

    applied to fluorescence or time resolved NMR data. It is also

    straightforward to extend the kinetic models used for non-

    isothermal conditions using either Arrhenius of Erying based

    rate constant temperature dependencies [3,15] (although careful

    attention must be paid to the temperature dependence of 

    spectra). It is also possible to mix measurements of different 

    types (for example, absorbance data covering different 

    wavelength ranges or absorbance and fluorescence data) or 

    use different types of modelling to calculate the concentration profiles (for example, equilibria models [42]).

    Appendix A. The Taylor series expansion

    The Taylor series expansion [51] is a mathematical means of 

    approximating the value of a function that cannot be explicitly

    calculated. It is used both for the numerical integration of a

    kinetic model and during the calculation of parameter shifts

    during non-linear regression. It relies on knowing the value

    taken by a function at some point   x. The derivative(s) of the

    function at  x  is (are) then used to extrapolate the value taken by

    the function at some other nearby point,   x +Δ x   (Δ x   is some

    small increment in x). As an example, consider some function f  ,for which the value of  f  ( x) is known. The full form of the Taylor 

    series expansion as would be used to calculate its value at   f  ( x

    +Δ x), and its truncated form as used in Euler's method and the

     Newton–Gauss–Levenberg/Marquardt method, are shown in

    Eqs. (26) and (27), respectively.

     f   ð x þ D xÞ ¼ f   ð xÞ þ  11!

    d f   ð xÞd x

      ðD xÞ þ   12!

    d2 f   ð xÞd2 x

    ðD xÞ2

    þ: : : þ   1n!

    dn f   ð xÞdn x

      ðD xÞn ð26Þ

     f   ð x þ D xÞ ¼ f   ð xÞ þ  11!

    d f   ð xÞd x

      ðD xÞ ð27Þ

    For practical reasons, generally only the first or second

    derivative of the function is used. However, the higher the order 

    of derivative that is used, the greater the accuracy with which

    the prediction of  f  ( x +Δ x) is made.

    References

    [1] J.R. Knutson, J.M. Beechem, L. Brand, Simultaneous analysis of multiple

    fluorescence decay curves: a global approach, Chem. Phys. Lett. 102

    (1983) 501–507.

    [2] P. Bugnon, J.C. Chottard, J.L. Jestin, B. Jung, G. Laurenczy, M. Maeder,

    A.E. Merbach, A.D. Zuberbühler, Second-order globalisation for the

    determination of activation parameters in kinetics, Anal. Chim. Acta 298

    (1994) 193–201.

    [3] M. Maeder, K.J. Molloy, M.M. Schumacher, Analysis of non-isothermal

    kinetic measurements, Anal. Chim. Acta 337 (1997) 73–81.

    [4] M. Maeder, A.D. Zuberbühler, Nonlinear least-squares fitting of 

    multivariate absorption data, Anal. Chem. 62 (1990) 2220–2224.

    [5] S. Bijlsma, H.F.M. Boelens, H.C.J. Hoefsloot, A.K. Smilde, Constrained

    least squares methods for estimating reaction rate constants fromspectroscopic data, J. Chemom. 16 (2002) 28–40.

    [6] S. Bijlsma, H.F.M. Boelens, A.K. Smilde, Determination of rate constants

    in second-order kinetics using UV–Visible spectroscopy, Appl. Spectrosc.

    55 (2001) 77–83.

    [7] S. Bijlsma, A.K. Smilde, Estimating reaction rate constants: comparison

     between traditional curve fitting and curve resolution, Anal. Chim. Acta

    419 (2000) 197–207.

    [8] E. Furusjö, L.-G. Danielsson, A method for the determination of reaction

    mechanisms and rate constants from two-way spectroscopic data, Anal.

    Chim. Acta 373 (1998) 83–94.

    [9] E. Furusjö, O. Svenssen, L.-G. Danielsson, Estimation of kinetic

     parameters from non-isothermal batch experiments monitored by in

    situ vibrational spectroscopy, Chemom. Intell. Lab. Syst. 66 (2003)

    1–14.

    [10] P.G. Lye, G.A. Lawrance, M. Maeder, Formation kinetics of pendant arm polyamine macrocycles with copper(II), J . Chem. Soc., Dalton Trans. 16

    (2001) 2376–2382.

    [11] D. Chatterjee, M.S.A. Hamza, M.M. Shoukry, A. Mitra, S. Deshmukh, R.v.

    Eldik, Interaction of [RuIII(edta)(H2O)]- with amino acids in aqueous

    solution. Equilibrium, kinetic and protease inhibition studies, J. Chem.

    Soc., Dalton Trans. (2003) 203–209.

    [12] S. Yamada, T. Kido, M. Tanaka, Enhanced reactivity of nickel(II)

    complexes involving multidentate ligands in their complexation, Inorg.

    Chem. 23 (1984) 2990–2993.

    [13] M. Maeder, Y.-M. Neuhold, G. Puxty, P. King, Analysis of reactions in

    aqueous solution at non-constant pH: no more buffers? Phys. Chem.

    Chem. Phys. 5 (2003) 2836–2841.

    [14] L. Balland, N. Mouhab, J.-M. Cosmao, L. Estel, Kinetic parameter 

    estimation of solvent-free reactions: application to esterification of acetic

    anhydride by methanol, Chem. Eng. Process. 41 (2002) 395–402.[15] P. Gemperline, G. Puxty, M. Maeder, D. Walker, F. Tarczynski, M.

    Bosserman, Calibration-free estimates of batch process yields and

    detection of process upsets using in situ spectroscopic measurements and

    nonisothermal kinetic models: 4-(dimethylamino)pyridine-catalyzed es-

    terification of butanol, Anal. Chem. 76 (2004) 2575–2582.

    [16] Y.-S. Duh, C.-C. Hsu, C.-S. Kao, S.W. Yu, Applications of reaction

    calorimetry in reaction kinetics and thermal hazard evaluation, Thermo-

    chim. Acta 285 (1996) 67–79.

    [17] A. Zogg, U. Fischer, K. Hungerbühler, Identification of kinetic and

    thermodynamic reaction parameters from online calorimetric and IR-ATR 

    data using a new combined evaluation algorithm, Chem. Eng. Sci. 59

    (2004) 5795–5806.

    [18] G. Donnati, R. Paludeto, Batch and semibatch catalytic reactors (from

    theory to practice), Catal. Today 52 (1999) 183–195.

    [19] E.R. Malinowski, Factor Analysis in Chemistry, John Wiley & Sons, Inc., New York, 1991.

    [20] M. Maeder, A.D. Zuberbühler, The Resolution of overlapping peaks by

    evolving factor analysis, Anal. Chim. Acta 181 (1986) 287–291.

    [21] R. Tauler, Multivariate curve resolution applied to second order data,

    Chemom. Intell. Lab. Syst. 30 (1995) 133–146.

    [22] A. de Juan, M. Maeder, M. Martinez, R. Tauler, Combining hard- and soft-

    modelling to solve kinetic problems, Chemom. Intell. Lab. Syst. 54 (2000)

    123–141.

    [23] A. de Juan, M. Maeder, M. Martinez, R. Tauler, Application of a novel

    resolution approach combining soft- and hard-modelling features to

    investigate temperature-dependent kinetic processes, Anal. Chim. Acta

    442 (2001) 337–350.

    [24] G. Puxty, M. Maeder, Y.-M. Neuhold, P. King, Pro-Kineticist II, Applied

    Photophysics Ltd, Leatherhead, Surrey, 2001 http://www.photophysics.

    com.

    163G. Puxty et al. / Chemometrics and Intelligent Laboratory Systems 81 (2006) 149 – 164

    http://www.photophysics.com/http://www.photophysics.com/http://www.photophysics.com/http://www.photophysics.com/

  • 8/9/2019 Tutorial on the fitting of kinetics model to multivariate spectroscopic measurements with nonlinear least-squares re…

    16/16

    [25] MATLAB 7.1. The Mathworks Inc., Natick, MA, 2005,  http://www.

    mathworks.com.

    [26] J.J.B. Baeza, G.R. Ramos, F.P. Pla, R.V. Molina, Multicomponent analysis

    using OPKINE, a program for the nonlinear treatment of kinetic problems,

    Analyst 115 (1990) 721–724.

    [27] M.M. Canedo, J.L. Gonzalez-Hernandez, A new computational applica-

    tion of the AGDC algorithm for kinetic resolution of multicomponent 

    mixtures (static and dynamic), Chemom. Intell. Lab. Syst. 66 (2003)63–78.

    [28] J.L. Gonzalez, M.M. Canedo, C. Grande, Optimization of kinetic

     parameters. Multipurpose KINAGDC(MW) non-linear regression pro-

    gram, Chemom. Intell. Lab. Syst. 39 (1997) 77–84.

    [29] SPECFIT/32. Spectrum Software Associates, Marlborough, MA, 2004,

    http://www.bio-logic.info/rapid-kinetics/specfit.html.

    [30] J.H. Espensen, Chemical Kinetics and Reaction Mechanisms, McGraw-

    Hill, New York, 1981.

    [31] W.H. Press, B.P. Flannery, S.A. Teukolsky, W.T. Vetterling, Numerical

    Recipes in C: The Art of Scientific Computing, Cambridge University

    Press, Cambridge, 1992.

    [32] R. Dyson, M. Maeder, G. Puxty, Y.-M. Neuhold, Simulation of complex

    kinetics, Inorg. React. Mech. 5 (2003) 39–46.

    [33] A.D. Zuberbühler, T.A. Kaden, Handling of electronic absorption spectra

    with a desk-top computer: II. Calculation of stability constants fromspectrophotometric titrations, Talanta 26 (1979) 1111–1118.

    [34] D.C. Lay, Linear Algebra and Its Applications, Addison-Wesley

    Publishing Company, New York, USA, 2000.

    [35] K.Q. Levenberg, A method for the solution of certain problems in least 

    squares, Appl. Math. 2 (1944) 164–168.

    [36] D.W. Marquardt, An algorithm for least-squares estimation of nonlinear 

     parameters, J. Soc. Ind. Appl. Math. 11 (1963) 431–441.

    [37] J.C. Lagarias, J.A. Reeds, M.H. Wright, P.E. Wright, Convergence

     properties of the Nelder –Mead simplex method in low dimensions, SIAM

    J. Optim. 9 (1998) 112–147.

    [38] R. Leardi, Genetic algorithms in chemometrics and chemistry: a review, J.

    Chemom. 15 (2001) 559–569.

    [39] G. Puxty, M. Maeder, P. Gemperline, R.R. Rhinehart, A. Samir, S. Moore,

    Modeling of batch reactions with in-situ spectroscopic measurements and

    calorimetry. J. Chemom. Accepted Copyright© 2005 John Wiley & SonsLtd.

    [40] J. Lumsden, Elementary Statistical Method, University of Western

    Australia Press, Nedlands, 1971.

    [41] R. Dyson, M. Maeder, Y.-M. Neuhold, G. Puxty, Analyses of three-way

    data from equilibrium and kinetic investigations, Anal. Chim. Acta 490

    (2003) 99–108.

    [42] R.M. Dyson, S. Kaderli, G.A. Lawrance, M. Maeder, A.D. Zuberbühler,

    Second order global analysis  –  the evaluation of series of spectrophoto-

    metric titrations for improved determination of equilibrium constants,Anal. Chim. Acta 353 (1997) 381–393.

    [43] S. Vajda, H. Rabitz, Identifiability and distinguishability of general

    reaction systems, J. Phys. Chem. 98 (1994) 5265–5271.

    [44] A.K.S. Ahmed, R.G. Wilkins, Factors influencing the rates of dissociation

    of metal complexes: Part I. The mechanism of dissociation of mono

    (ethylenediamine)–nickel(II) ion, J. Chem. Soc. (1959) 3700–3708.

    [45] A.K.S. Ahmed, R.G. Wilkins, Factors influencing the rates of dissociation

    of metal complexes: Part III. The dissociation of bis(ethylenediamine)–

    and tris(ethylenediamine)–nickel complexes and of the racemic 2,3-

    diaminobutane analogues, J. Chem. Soc. (1960) 2901–2906.

    [46] A.K.S. Ahmed, R.G. Wilkins, Factors influencing the rates of dissociation

    of metal complexes: Part II. The dissociation of complexes of nickel with

    various diamines, J. Chem. Soc. (1960) 2895–2900.

    [47] K.J. Wannowius, K. Krimm, H. Elias, Visible spectra of the complex

    species Ni(en)32+

    , Ni(en)2(H2O)22+

    and Ni(en)(H2O)42+

    from kinetic studies,Inorg. Chim. Acta 127 (1987) L43–L44.

    [48] G. Puxty, Novel Chemometric Methods for the Hard Modelling of 

    Multivariate Spectroscopic Data, The University of Newcastle, Newcastle,

    Australia, 2004.

    [49] W.G. Jackson, J.M. Harrowfield, P.D. Vowles, Consecutive, irreversible

    first-order reactions. Ambiguities and practical aspects of kinetic analyses,

    Int. J. Chem. Kinet. 9 (1977) 535–548.

    [50] T.J. Kemp, P. Moore, G.R. Quick, Kinetic and spectroscopic studies with a

    rapid-scanning Spectrometer: Part 1. Transient intermediates in the

    dissociation of nickel(II) polyamine complexes, J. Chem. Soc. 92 (1979)

    1377–1383.

    [51] C.H. Edwards, D.E. Penney, Calculus with Analytical Geometry, Prentice

    Hall International Inc., New Jersey, 1994.

    164   G. Puxty et al. / Chemometrics and Intelligent Laboratory Systems 81 (2006) 149 – 164

    http://www.mathworks.com/http://www.mathworks.com/http://www.bio-logic.info/rapid-kinetics/specfit.htmlhttp://www.bio-logic.info/rapid-kinetics/specfit.htmlhttp://www.mathworks.com/http://www.mathworks.com/