Determination of equilibrium constants: Difference between revisions
No edit summary |
No edit summary |
||
Line 81: | Line 81: | ||
where p<sub>k</sub> is the kth parameter of the refinement. |
where p<sub>k</sub> is the kth parameter of the refinement. |
||
One or more equilibrium constants may be parameters of the refinement. However, examination of the expressions for the measured quantities (above) reveals that the equilibrium constants do not appear in the expressions, but that the concentrations, ''c'', are implicit functions of these parameters. Therefore the Jacobian must be obtained using [[implicit differentiation]]. |
One or more equilibrium constants may be parameters of the refinement. However, examination of the expressions for the measured quantities (above) reveals that the equilibrium constants do not appear in the expressions, but that the concentrations, ''c'', are implicit functions of these parameters. Therefore the Jacobian must be obtained using [[implicit differentiation]]. Parameter increments are calculated by solving the [[normal equations]] |
||
:<math> \mathbf{\delta p= \left(J^TJ +\lambda I\right)^{-1}J^T \left(y^{obs}-y^{calc} \right) }</math> |
|||
where <math>\lambda</math> is the Marquardt parameter. The increments are added to the current parameter estimates. |
|||
A particular issue arises with NMR and spectrophotometric data. For the latter, the Beer-Lambert law can be written as |
A particular issue arises with NMR and spectrophotometric data. For the latter the observed quantity is absorbance, A, and the Beer-Lambert law can be written as |
||
:<math> |
:<math>A_i=\sum\epsilon_{pq...} c_{pq...,i}</math> |
||
It can be seen that absorbance, A, is a linear function of the molar absorbptivities, <math>\epsilon</math>. |
It can be seen that absorbance, A, is a linear function of the molar absorbptivities, <math>\epsilon</math>, at the path length used. In matrix notation |
||
:<math>\mathbf{A=\Epsilon C}</math> |
|||
There are two approaches to the calculation of the unknown molar absorptivities |
There are two approaches to the calculation of the unknown molar absorptivities |
||
:1) The <math>\epsilon</math> values are considered to be parameters of the minimization and the Jacobian is constructed on that basis. However, the <math>\epsilon</math> values themselves are calculated at each step of the refinement by linear least-squares: |
:1) The <math>\epsilon</math> values are considered to be parameters of the minimization and the Jacobian is constructed on that basis. However, the <math>\epsilon</math> values themselves are calculated at each step of the refinement by linear least-squares: |
||
::<math>\mathbf{\ |
::<math>\mathbf{\Epsilon = \left(C^TC\right)^{-1}C^TA }</math> |
||
⚫ | |||
:where <math>J'_{jk}=\frac{\partial A_j^{calc}}{\partial c_k}</math> |
|||
⚫ | |||
:2) The Beer-Lambert law is written as |
:2) The Beer-Lambert law is written as |
||
::<math>A= |
::<math>\mathbf{A= \left( \left( C^TC \right)^{-1}C^TA \right) C}</math> |
||
⚫ | :Golub and Pereyra<ref name="Golub">G.H. Golub and V. Pereyra, ''SIAM J. Numer. Anal''., '''2''', 413-432 (1973) </ref> showed how the pseudo-inverse can be differentiated so that parameter increments for both molar absorptivities and equilibrium constants can be calculated by solving the normal equations. |
||
:where <math>J'_{jk}=\frac{\partial A_j^{calc}}{\partial c_k}</math> |
|||
⚫ | |||
== Implementations == |
== Implementations == |
||
Some simple systems are amenable to spreadsheet calculations.<ref>E.J. Billo, EXCEL for Chemists, Wiley-VCH, 2nd. edition 2001</ref> These calculations do not follow the general procedures outlined here and use SOLVER to perform the least-squares minimization. |
Some simple systems are amenable to spreadsheet calculations.<ref>E.J. Billo, EXCEL for Chemists, Wiley-VCH, 2nd. edition 2001</ref> These calculations do not follow the general procedures outlined here and use SOLVER to perform the least-squares minimization. |
||
A large number of computer programs for equilibrium constant calculation have been published. See <ref name="HQ"> P. Gans, A. Sabatini and A. Vacca, ''Talanta'', '''43''',1739-1753 (1996)</ref> for bibliography. The most |
A large number of computer programs for equilibrium constant calculation have been published. See <ref name="HQ"> P. Gans, A. Sabatini and A. Vacca, ''Talanta'', '''43''',1739-1753 (1996)</ref> for bibliography. The most frequently used programs are: |
||
* Potentiometric data: [http://www.hyperquad.co.uk/hq2000.htm Hyperquad], BEST<ref>A.E. Martell and R.J. Motekaitis, The determination and use of stability constants, Wiley-VCH, 1992.</ref> PSEQUAD<ref name=Leggett>D. J. Leggett (editor), Computational methods for the determination of formation constants, Plenum Press, 1985.</ref> |
* Potentiometric data: [http://www.hyperquad.co.uk/hq2000.htm Hyperquad], BEST<ref>A.E. Martell and R.J. Motekaitis, The determination and use of stability constants, Wiley-VCH, 1992.</ref> PSEQUAD<ref name=Leggett>D. J. Leggett (editor), Computational methods for the determination of formation constants, Plenum Press, 1985.</ref> |
||
* Spectrophotometric data:[http://www.hyperquad.co.uk/hq2000.htm Hyperquad], SQUAD<ref name="Leggett"/>,Specfit |
* Spectrophotometric data:[http://www.hyperquad.co.uk/hq2000.htm Hyperquad], SQUAD<ref name="Leggett"/>,Specfit,<ref>H. Gampp, M. Maeder, C.J.Mayer and a. Zuberbühler, ''Talanta'', '''32''', 95, 257 (1985)</ref> Specfit/32 is now available commercially |
||
* NMR data [http://www.hyperquad.co.uk/hypnmr.htm HypNMR], [http://www.nuigalway.ie/chem/Mike/eqnmr.htm EQNMR] |
* NMR data [http://www.hyperquad.co.uk/hypnmr.htm HypNMR], [http://www.nuigalway.ie/chem/Mike/eqnmr.htm EQNMR] |
||
Revision as of 08:07, 5 May 2007
This article is actively undergoing a major edit for a little while. To help avoid edit conflicts, please do not edit this page while this message is displayed. This page was last edited at 08:07, 5 May 2007 (UTC) (17 years ago) – this estimate is cached, . Please remove this template if this page hasn't been edited for a significant time. If you are the editor who added this template, please be sure to remove it or replace it with {{Under construction}} between editing sessions. |
- See Chemical equilibrium and stability constant for definitions and derivations
Experimental methods
A general expression for an equilibrium constant, expressed as a concentration quotient,
shows that it is a function of the concentrations [A], [B] etc. of the chemical species in equilibrium. The equilibrium constant value can be determined if any one of these concentrations can be measured. The general procedure is that the concentration in question is measured for a series of solutions with known analytical concentrations of the reactants. Typically, a titration is performed with one or more reactants in the titration vessel and one or more reactants in the burette. Knowing the analytical ooncentrations of reactants initially in the reaction vessel and in the burette, all analytical concentrations can be derived as a function of the volume (or mass) of titrant added.
There are four main experimental methods. For less commonly used methods see Rossotti and Rossotti[1]
Potentiometric measurements
A free concentration [A] or activity {A} is measured by means of an ion selective electrode such as the glass electrode. If the electrode is calibrated using activity standards it is assumed that the Nernst equation applies in the form
- E=E0+RT/nF ln{A}
where E0 is the standard electrode potential. When buffer solutions of known pH are used for calibration the meter reading will be pH. See IUPAC Compendium of Analytical Nomenclature, Table 8.4.2for primary pH standards.
When the electrode is calibrated with solutions of known concentration, by means of a strong acid/strong base titration, for example, a modified Nernst equation is assumed.
- E=E0+s lg[A]
s an empirical slope factor
Spectrophotometric measurements
It is assumed that the Beer-Lambert law applies.
where is the optical path length, is a molar absorbance at unit path length and c is a concentration. More than one of the species may contribute to the absorbance. In principle absorbance may be measured at one wavelength only, but in present-day practice it is common to record complete spectra.
Fluorescence (luminescence) intensity
It is assumed that the scattered light intensity is a linear function of species’ concentrations.
where is a proportionality constant.
NMR chemical shift measurements
Chemical exchange is assumed to be rapid on the NMR time-scale. An individual chemical shift is the mol-fraction weighted average of the shift of contributing species.
Calorimetric measurements
Simultaneous measurement of K and H for 1:1 adducts is routinely carried out using Isothermal Titration Calorimetry. Extension to more complex systems is limited by the availability of suitable software.
Range and limitations
- Potentiometry. The most widely used electrode is the glass electrode which is selective for the hydrogen ion. This is suitable for all acid-base equilibria. Lg values between about 2 and 11 can be measured directly by potentiometric titration using a glass electrode. This enormous range is possible because of the logarithmic response of the electrode. The limitations arise because the Nernst equation breaks down at very low or very high pH. The range can be extended by using the competition method.
- Absorbance and Luminescence. An upper limit on lg of 4 is usually quoted, corresponding to the precision of the measurements, but it also depends on how intense the effect is. Spectra of contributing species should be clearly distinct from each other
- NMR. Limited precision of chemical shift measurements also puts an upper limit of about 4 on lg . Limited to diamagnetic systems.
- Calorimetry. Insufficient evidence is currently available.
Computational methods
It is assumed that the experimental data which have been collected comprise a set of data points. At each data point, i, the analytical concentrations of the reactants, TA(i), TB(i) etc. are known along with a measured quantity, yi. A general computational procedure has three main components.
- Definion of a chemical model of the equilibria
- Calculation of the concentrations of all the chemical species in each solution
- Refinement of the equilibrium constants
The chemical model
The chemical model consists of a set of reactants and the complexes formed from them. Denoting the reactants by A, B ..., each complex is specified by its stoichiometric coefficients.
Ionic charges are omitted from these expression. For consistency all the equilibrium constants should be association constants. When using general-purpose computer programs it is usual to use cumulative constants. With aqueous solutions the constant for the self-disscotiation of water should be included.
It is quite usual to omit from the model species whose concentrations are considered to be negligible. For example it is usually assumed then there is no interaction between the reactants and/or complexes and the electrolyte used to maintain constant ionic strength or the buffer used to maintain constant pH. These assumptions may or may not be justified. Also, it is implicitly assumed that there are no other complex species present. When complexes are wrongly ignored a systematic error is introduced into the calculations.
Equilibrium constant values are usually estimated initially by reference to data sources.
Speciation calculations
The speciation calculations consist of solving the equations of mass-balance, at each data point, for the concentrations of free reactant, [A], [B] ...
In general solving these non-linear equations presents a formidable challenge because of the huge range over which the free concentrations may vary. Once the free reactant concentrations have been calculated the concentrations of the complexes are derived from them and the equilibrium constants. Note that the free reactant concentrations can be regarded as implicit parameters in the equilibrium constant refinement process.
The speciation calculations are repeated at each refinement step.
Equilibrium constant refinement
The objective of the refinement process it to find equilibrium constant values that give the best fit to the experimental data. This is usually acheived by minimising an objective function, U, by the method of non-linear least-squares.
The matrix of weights, W, should be, ideally, the inverse of the variance-covarance matrix of the observations. It is rare for this to be known. However, when it is the expectation value of U is one, which means that the data are fitted within experimental error. Most often only the diagonal elements are known, in which case the objective function simplifies to
Unit weights, Wi=1, are ofter used, but in that case the expectation value of U is the root mean square of the experimental errors. The minimization may be performed using the Gauss-Newton method with a Levenberg-Marquardt parameter. This method requires the Jacobian, J, to be calculated.
where pk is the kth parameter of the refinement.
One or more equilibrium constants may be parameters of the refinement. However, examination of the expressions for the measured quantities (above) reveals that the equilibrium constants do not appear in the expressions, but that the concentrations, c, are implicit functions of these parameters. Therefore the Jacobian must be obtained using implicit differentiation. Parameter increments are calculated by solving the normal equations
where is the Marquardt parameter. The increments are added to the current parameter estimates.
A particular issue arises with NMR and spectrophotometric data. For the latter the observed quantity is absorbance, A, and the Beer-Lambert law can be written as
It can be seen that absorbance, A, is a linear function of the molar absorbptivities, , at the path length used. In matrix notation
There are two approaches to the calculation of the unknown molar absorptivities
- 1) The values are considered to be parameters of the minimization and the Jacobian is constructed on that basis. However, the values themselves are calculated at each step of the refinement by linear least-squares:
- using the refined values of the equilibrium constants to obtain the speciation. The matrix is an example of a pseudo-inverse.
- 2) The Beer-Lambert law is written as
- Golub and Pereyra[2] showed how the pseudo-inverse can be differentiated so that parameter increments for both molar absorptivities and equilibrium constants can be calculated by solving the normal equations.
Implementations
Some simple systems are amenable to spreadsheet calculations.[3] These calculations do not follow the general procedures outlined here and use SOLVER to perform the least-squares minimization.
A large number of computer programs for equilibrium constant calculation have been published. See [4] for bibliography. The most frequently used programs are:
- Potentiometric data: Hyperquad, BEST[5] PSEQUAD[6]
- Spectrophotometric data:Hyperquad, SQUAD[6],Specfit,[7] Specfit/32 is now available commercially
- NMR data HypNMR, EQNMR
References
- ^ F.J,C. Rossotti and H. Rossotti, The determination of stability constants, McGraw-Hill, 1961.
- ^ G.H. Golub and V. Pereyra, SIAM J. Numer. Anal., 2, 413-432 (1973)
- ^ E.J. Billo, EXCEL for Chemists, Wiley-VCH, 2nd. edition 2001
- ^ P. Gans, A. Sabatini and A. Vacca, Talanta, 43,1739-1753 (1996)
- ^ A.E. Martell and R.J. Motekaitis, The determination and use of stability constants, Wiley-VCH, 1992.
- ^ a b D. J. Leggett (editor), Computational methods for the determination of formation constants, Plenum Press, 1985.
- ^ H. Gampp, M. Maeder, C.J.Mayer and a. Zuberbühler, Talanta, 32, 95, 257 (1985)