Big Chemical Encyclopedia

Chemical substances, components, reactions, process design ...

Articles Figures Tables About

Chemical component variable number

The application of principal components regression (PCR) to multivariate calibration introduces a new element, viz. data compression through the construction of a small set of new orthogonal components or factors. Henceforth, we will mainly use the term factor rather than component in order to avoid confusion with the chemical components of a mixture. The factors play an intermediary role as regressors in the calibration process. In PCR the factors are obtained as the principal components (PCs) from a principal component analysis (PC A) of the predictor data, i.e. the calibration spectra S (nxp). In Chapters 17 and 31 we saw that any data matrix can be decomposed ( factored ) into a product of (object) score vectors T(nxr) and (variable) loadings P(pxr). The number of columns in T and P is equal to the rank r of the matrix S, usually the smaller of n or p. It is customary and advisable to do this factoring on the data after columncentering. This allows one to write the mean-centered spectra Sq as ... [Pg.358]

We have considered a large number of values (including the molality of each aqueous species, the mole number of each mineral, and the mass of solvent water) to describe the equilibrium state of a geochemical system. In Equations 3.32-3.35, however, this long list has given way to a much smaller number of values that constitute the set of independent variables. Since there is only one independent variable per chemical component, and hence per equation, we have succeeded in reducing the number of unknowns in the equation set to the minimum possible. In addition,... [Pg.44]

Fig. 4.4. Comparison of the computing effort, expressed in thousands of floating point operations (Aflop), required to factor the Jacobian matrix for a 20-component system (Nc = 20) during a Newton-Raphson iteration. For a technique that carries a nonlinear variable for each chemical component and each mineral in the system (top line), the computing effort increases as the number of minerals increases. For the reduced basis method (bottom line), however, less computing effort is required as the number of minerals increases. Fig. 4.4. Comparison of the computing effort, expressed in thousands of floating point operations (Aflop), required to factor the Jacobian matrix for a 20-component system (Nc = 20) during a Newton-Raphson iteration. For a technique that carries a nonlinear variable for each chemical component and each mineral in the system (top line), the computing effort increases as the number of minerals increases. For the reduced basis method (bottom line), however, less computing effort is required as the number of minerals increases.
At this point we can see the advantage of working with the reduced problem. Most published algorithms carry a nonlinear variable for each chemical component plus one for each mineral in the system. The number of nonlinear variables in the method presented here, on the other hand, is the number of components minus the number of minerals. Depending on the size of the problem, the savings in computing effort in evaluating Equation 4.31 can be dramatic (Fig. 4.4). [Pg.63]

A homogeneous open system consists of a single phase and allows mass transfer across its boundaries. The thermodynamic functions depend not only on temperature and pressure but also on the variables necessary to describe the size of the system and its composition. The Gibbs energy of the system is therefore a function of T, p and the number of moles of the chemical components i, tif. [Pg.24]

In the preceding section we have set up the canonical ensemble partition function (independent variables N, V, T). This is a necessary step whether one decides to use the canonical ensemble itself or some other ensemble such as the grand canonical ensemble (p, V, T), the constant pressure canonical ensemble (N, P, T), the generalized ensemble of Hill33 (p, P, T), or some form of constant pressure ensemble like those described by Hill34 in which either a system of the ensemble is open with respect to some but not all of the chemical components or the system is open with respect to all components but the total number of atoms is specified as constant for each system of the ensemble. We now consider briefly the selection of the most convenient formalism for the present problem. [Pg.17]

As discussed in the introduction, the solution of the inverse model equation for the regression vector involves the inversion of R R (see Equation 5 23). In many anal al chemistry experiments, a large number of variables are measured and R R cannot be inverted (i.e., it is singular). One approach to solving this problem is called stepwise MLR where a subset of variables is selected such that R R is not singular. There must be at least as many variables selected as there are chemical components in the system and these variables must represent different sources of variation. Additional variables are required if there are other soairces of variation (chemical or physical) that need to be modeled. It may also be the case that a sufficiently small number of variables are measured so that MIR can be used without variable selection. [Pg.130]

Although gibbsite and kaolinite are important in quantity in some soils and hydrothermal deposits, they have diminishing importance in argillaceous sediments and sedimentary rocks because of their peripheral chemical position. They form the limits of any chemical framework of a clay mineral assemblage and thus rarely become functionally involved in critical clay mineral reactions. This is especially true of systems where most chemical components are inert or extensive variables of the system. More important or characteristic relations will be observed in minerals with more chemical variability which respond readily to minor changes in the thermodynamic parameters of the system in which they are found. However, as the number of chemical components which are intensive variables (perfectly mobile components) increases the aluminous phases become more important because alumina is poorly soluble in aqueous solution, and becomes the inert component and the only extensive variable. [Pg.33]

There are zeolite-bearing rocks in which one mineral is apparently being replaced by another mineral under constant P-T conditions. This indicates a system in which certain chemical components appear to be perfectly mobile a system in which the total number of phases that can coexist at equilibrium is reduced as a function of the number of chemical components which ar e internal variables of the system. Two examples of this type of equilibrium concerning zeolites can be cited saline lakes and analcite-bearing soil profiles (Hay, 1966 Hay and Moiola, 1963 Jones, 1965 and Frankart and Herbillon, 1970). In both cases a montmorillonite-bearing assemblage becomes analcite or zeolite-bearing at the expense of the expandable phyllosilicate. Other phases remain constantly present. [Pg.133]

The use of the "closed system" to describe the assemblages in these closed basins seems justified in that frequently, most always, in fact, the number of clay minerals present in the sediments discussed above is two or more. The omnipresence of amorphous silica or chert raises the total number of phases to three. In an essentially three-component system, Mg-Si-Al or possibly four if H+ is considered, this indicates that the chemical components of the minerals are present in relatively fixed quantities in the chemical system which produces the mineral assemblages. None of the first three components is "mobile", i.e., its activity is independent of its relative mass in the solids or crystals present. However, there are sediments which present a monophase assemblage where only one variable need be fixed. Under these conditions sepiolite can be precipitated from solution and pre-existing solid phases need not be involved. [Pg.151]

Figure 46a. Development of a portion of the system K-Al-Si as an increasing number of the chemical components become intensive variables of a given system. F = feldspar Mi = mica G = gibbsite Kaol = kaolinite Q -quartz Mo = montmorillonite ML = mixed layered mineral. All chemical components are extensive variables. Figure 46a. Development of a portion of the system K-Al-Si as an increasing number of the chemical components become intensive variables of a given system. F = feldspar Mi = mica G = gibbsite Kaol = kaolinite Q -quartz Mo = montmorillonite ML = mixed layered mineral. All chemical components are extensive variables.
It is evident then that the number of phases present can be reduced by restricting the intensive variables to non-unique or inter-dependent conditions and thus a general case" is valid. As the number of chemical components which are perfectly mobile (intensive variables) increases, the number of phases will be decreased. Thus a rock banded in mono-mineralic zones, such as is commonly found in hydrothermal deposits, indicates a system with few inert components. The zonation most commonly represents gradients of the extensive variables temperature and chemical activity of ionic species in aqueous solution. However, it is not always easy to determine which variables are active in a given sequence of mineral bands related to an alteration source. [Pg.176]

Since the state of a crystal in equilibrium is uniquely defined, the kind and number of its SE s are fully determined. It is therefore the aim of crystal thermodynamics, and particularly of point defect thermodynamics, to calculate the kind and number of all SE s as a function of the chosen independent thermodynamic variables. Several questions arise. Since SE s are not equivalent to the chemical components of a crystalline system, is it expedient to introduce virtual chemical potentials, and how are they related to the component potentials If immobile SE s exist (e.g., the oxygen ions in dense packed oxides), can their virtual chemical potentials be defined only on the basis of local equilibration of the other mobile SE s Since mobile SE s can move in a crystal, what are the internal forces that act upon them to make them drift if thermodynamic potential differences are applied externally Can one use the gradients of the virtual chemical potentials of the SE s for this purpose ... [Pg.21]

The results of the discussion on the phenomenological thermodynamics of crystals can be summarized as follows. One can define chemical potentials, /jk, for components k (Eqn. (2.4)), for building units (Eqn. (2.11)), and for structure elements (Eqn, (2.31)). The lattice construction requires the introduction of structural units , which are the vacancies V,. Electroneutrality in a crystal composed of charged SE s requires the introduction of the electrical unit, e. The composition of an n component crystal is fixed by n- 1) independent mole fractions, Nk, of chemical components. (n-1) is also the number of conditions for the definition of the component potentials juk, as seen from Eqn. (2.4). For building units, we have (n — 1) independent composition variables and n-(K- 1) equilibria between sublattices x, so that the number of conditions is n-K-1, as required by the definition of the building element potential uk(Xy For structure elements, the actual number of constraints is larger than the number of constraints required by Eqn. (2.18), which defines nk(x.y This circumstance is responsible for the introduction of the concept of virtual chemical potentials of SE s. [Pg.26]

Consider a material or system that is not at equilibrium. Its extensive state variables (total entropy number of moles of chemical component, i total magnetization volume etc.) will change consistent with the second law of thermodynamics (i.e., with an increase of entropy of all affected systems). At equilibrium, the values of the intensive variables are specified for instance, if a chemical component is free to move from one part of the material to another and there are no barriers to diffusion, the chemical potential, q., for each chemical component, i, must be uniform throughout the entire material.2 So one way that a material can be out of equilibrium is if there are spatial variations in the chemical potential fii(x,y,z). However, a chemical potential of a component is the amount of reversible work needed to add an infinitesimal amount of that component to a system at equilibrium. Can a chemical potential be defined when the system is not at equilibrium This cannot be done rigorously, but based on decades of development of kinetic models for processes, it is useful to extend the concept of the chemical potential to systems close to, but not at, equilibrium. [Pg.8]

In the Monte Carlo approach, there are no inherent limitations on the complexity of the exposure equation, the number of component variables, the probability distributions for the variable components, or the number of iterations. This freedom from limitations is especially useful in simulating the distributions of a LADD for the different exposure scenarios considered here. As its name suggests, a LADD is the average over all the days in an individual s lifetime of the dose of a chemical (e.g., atrazine, simazine, or both) received as a result of his or her exposure from one or more exposure pathways (e.g., water, diet, or herbicide handling). Because the exposure equation can explicitly consider each day individually, the values of the equation s variable components can vary from day to day and have different distributions for different ages and different lifespan projections. [Pg.481]

The number of unknowns and the number of equations relating these unknowns can become very large in a process-design problem. The number of unknowns and independent equations must be equal in order that a unique solution to a problem exists. Therefore, it is necessary to have a systematic method for enumerating them. The total number of independent extensive and intensive variables associated with each stream in a process is C + 2, where C is the number of independent chemical components in the stream. The quantity and the condition of the stream are completely determined by fixing the flow rate of each component in the stream (or, equivalently, the total flow rate and the mole or mass fractions of C — 1 components) and two additional variables, usually the temperature and pressure, although other choices are possible. This number includes situations where physical and chemical equilibrium exist.t... [Pg.128]

The low correlation coefficient observed when combining the data for alcohol ethoxylates from two different manufacturers is probably due to chemical structure variables not incorporated in equation 1. These include different degrees of hydrophobe linearity, different distributions of carbon numbers and ethylene oxide chain lengths around the average values used in the correlation equation, and the presence of differing amounts of other components such as unreacted alcohols in the surfactants. [Pg.190]

The variance of the combined set of equations of the types shown in Equations (3) and (5) is equivalent to the thermodynamic degrees of freedom ( the number of chemical components plus 2 (for P and T) minus the number of minerals ). This set of equations involves only intensive variables, and is important for inferring P-T paths from chemical zoning. [Pg.1495]

H = H T,p, M, M2, Ms,Mjv), where N is the number of chemical components. An extensive quantity can be divided by the mass of the system constituting a new variable defining a specific quantity ([3], p 103). The specific enthalpy (per unit mass) is then expressed as ... [Pg.1140]

The smallest number of independent variables that must be specified in order to describe completely the state of a system is known as the number of degrees of freedom,/. For a fixed mass of a gas, / = 2, because one can vary any two variables (e.g., pressure and temperature) the third variable is fixed by the equation of state (e.g. volume). Hence, only two properties of a fixed mass of gas are independently variable. Stated mathematically, for a system at equilibrium, the number of degrees of freedom, /, equals the difference between the number of chemical components, c, and the number of phases, p (22) ... [Pg.107]

There are two problems leading to the notion of continuous components. At first, the number of species or chemical components may reach a huge number in such areas as oil chemistry, polymerisation or biochemistry. In these cases the number of the variables in the induced kinetic differential equation is so large that this system is difficult to treat it may prove more promising to have a continuous manifold of chemical components. [Pg.78]


See other pages where Chemical component variable number is mentioned: [Pg.894]    [Pg.16]    [Pg.50]    [Pg.245]    [Pg.239]    [Pg.155]    [Pg.446]    [Pg.10]    [Pg.419]    [Pg.161]    [Pg.189]    [Pg.245]    [Pg.208]    [Pg.60]    [Pg.131]    [Pg.168]    [Pg.81]    [Pg.915]    [Pg.966]    [Pg.42]    [Pg.13]    [Pg.111]    [Pg.210]    [Pg.354]    [Pg.441]   
See also in sourсe #XX -- [ Pg.231 ]

See also in sourсe #XX -- [ Pg.231 ]




SEARCH



Chemical variability

Chemical variables

Chemicals components

Components, number

Variables components

© 2024 chempedia.info