Sunteți pe pagina 1din 19

LIMITS EQUATIONS IN THE STUDY OF CHEMICAL SYSTEMS; THE SCIENCES ROUTINE C.

Contreras-Ortega*1 , Nelson Bustamante 1 , Juan Luis Guevara 2 , Victor Kesternich1 and Carlos Portillo3 1 Universidad Catlica del Norte, Facultad de Ciencias. Casilla 1280 Antofagasta, Chile. *e-mail: ccontrer@ucn.cl 2 Universidad de Chile, Facultad de Ciencias, Casilla 653, Santiago, Chile 3 Universidad de Chile, Facultad de Ciencias Fsicas y Matemticas, Casilla 487-3, Santiago, Chile

ABSTRACT
The scope of limit equations is analyzed and their usefulness as starting points to deduce general equations, as well. A general strategy is proposed to go systematically from limit equations to general ones. Emphasis is made on the necessity of making clear to students the points above. On the base of the proposed strategy, the most common general equations in chemistry, like the equation for the chemical equilibrium constant, the Nernst equation for an electrochemical reaction, and the Henrys and Raoults equations for the vapor pressure of solutions, are deduced and their scope analyzed. Key words: ideal gases, ideal solutions, real solutions, real gases, limit equations

Introduction General Chemistry and introductory Physical Chemistry courses teach mostly equations that are strictly valid for system under limiting physical - conditions only (1 - 6). These conditions are normally defined in terms of physical or chemical parameters at their limiting values, as for example, infinitely dilute concentrations, infinitely high temperatures or low pressures, etc. Examples of these equations are the well known concentrations-based equations for the chemical equilibrium constant, for the Nernsts equation for electrochemical cells, and for the Raoults and Henrys laws for the solution constituents; or, the widely used equation PV = nRT for gases. They are named limit equations. Regarding above several questions arise: i) why do limit equations appear?; ii) what part of Nature do they explain?; iii) how are explained situations others than the limiting ones?. Generally, the answers to these questions are not given at first year chemistry courses neither students are warned about the real scope of these limiting equations, exception made of the equations for gaseous systems. If students were aware about the scopes of limit equations concerning, for example, liquid solution systems they would ask themselves what about the situations far away from the limiting behaviors of nature. To answer above questions concepts like activity and activity coefficients are needed, concepts that are introduced only in upper chemistry courses. However, before receiving those answers students in these courses spend maybe too much time studying ideal solution systems and real solution system at their limiting borderline conditions, where limit equations apply, that they get convinced that these systems are the most common ones to be encountered in nature. Then when real systems far from their limiting borderline are introduced, they look like

the exception to the most common cases of interest and because of that not much attention is paid to them. Furthermore, the success of most of the limit equations in many practical applications seems to rule out the need of more sophisticated equations. The aim of the present article is to answer above questions in a general way in order to get a better understanding of the real scope of the limit equations and of their usefulness to derive the equations for real systems away from limiting physical borderline conditions. A subject closely related is the one concerning the roles of mathematical tendencies and limits in equations of chemical systems at limiting physical conditions, a subject that we analyzed in a previous article in this Journal (7). It is also a purpose of the present work to illustrate the way science infers general laws from particular ones (the scientific inductive method). We believe that this methodology is not consciencetly perceived by students (and may be not by some teachers either) in the long way of going from the limit laws (limit equation) of the general chemistry courses to the general ones learned in their last physical chemistry courses. And, generally, textbooks and teachers do not make explicit to students this role of the science either. The Science Routine Limit equations are the mathematical formulations of laws describing the regular behaviors of systems under limiting physical borderline conditions. As such, those laws are known also as limit laws. These regularities have been possible to observe in the laboratory by just claiming the parameters that are normally measured: pressures, volumes, temperature, and concentrations. They are observed by the experimenter at physical limiting conditions (low pressures, high temperatures, large volumes, highly dilute solutions, etc.) because under these conditions mater interactions became simpler (and weaker) as theories explain. And the simplest (and the weaker) the interactions are the simplest the explanations for them are. And the simplest explanations just claim to the parameters easier to observe, i.e., the ones that are normally measured. Nevertheless, science must look for general equations, capable to explain the whole behavior of nature. In such regard, limiting physical situations must also be explained by these general equations. Then, limits equations, the equations for the limiting physical situations, are nothing but particular forms of these general equations. The way science normally searches above general equations is as follows. In a first step, scientists simplify the reality they try to understand by progressively making simpler the experiment that represents it, until they start to observe regularities; they start to understand that reality. The regularities are then expressed as laws and these ones coded in mathematical formula. This is the case of all limit laws and equations, being the limit systems the simplified experimental systems. In a second step, these limiting real situations are taken as models (experimental models) to imaging idealized experimental situations, which weather eliminate the borderline restrictions of the limiting real systems or restrict reality to the borderlines where it can be understood. Because of that, these systems are considered as ideal ones. Thus, limit laws and equations are raised to the category of general laws; they are now capable to describe the nature in its all range of existence, but of an idealized nature. These ideal systems, and their laws and equations, are then submitted to the formalism of general equations and new equations are deduced for them. The final step is to modify these fictional general equations to obtain real general equations, capable of giving account of the complete reality. We name all this procedure as the science routine.

I.

The Limit Equations for Limiting Real Systems: the reality we can understand. Let analyze the most common examples of limit laws or equations:

The concentrationsbased chemical equilibrium constant. Let represent a general chemical equilibrium reaction as follows: aA + bB cC + dD Then, the expression for its equilibrium constant is: K(C i) = CCc CDd / CAa CBb ;
Ci

[1]

[2]

In eqn. [2], Ci represents the concentration of each one of the participants in the reaction, at the equilibrium, in a given concentration scale and the exponents are the respective reaction stoichiometry coefficients. As it is normally said, this equation is valid at dilute to highly dilute concentrations of the participants of the chemical reaction, which is indicated in eqn. [2] as Ci 0. From a practical point of view, this means that from tenth of a molar to lower concentrations the equilibrium constant calculated with eqn. [2] differs from the value believed to be real from 10% percent to better agreements in most of the cases. The concentrations-based Nernsts equation for electrochemical cells. Let assume that reaction [1] is the electrochemical reaction responsible of the electromotive force, emf, of a cell and that it takes place at constant temperature and pressure. Then, concentrations-based Nernst equation predicts for that cell:

= - (RT/nF) ln { C Cc CDd / C Aa CBb } ;

Ci

[3]

In above equation, and are the real and the standard emf of the cell, respectively, T is the reaction temperature, R is the gases constant, F is the Faraday constant, n is the net number of electron involved in the cell reaction, and the Cis are the actual concentrations, in a given concentration scale, at their dilutions limiting values, as indicated by Ci 0. This means, as before, that at higher concentrations the emf of the cell calculated with eqn. [3] differs from the real one (the measured value) in a magnitude that might not be not acceptable for the experimenter. Generally, the agreement between these values is much better than the agreement between the calculated and the actual equilibrium constants for the same involved chemical reaction. This is because the concentrations ratio in the emf equation, [3], is within a logarithmic expression, which diminishes the errors in that calculation. Thus, while at a tenth of a molar concentration the equilibrium constant calculated with eqn. [2] may differ from the value believed to be real in a 10% percent, the corresponding emf value may differ from the value believed to be real in about 0.1% percent when calculated with eqn. [3]. The vapor pressures of infinitely dilute solutions. For a general solution formed by a solvent A and a solute B, the vapor pressures for the solvent and the solute are respectively given by:

Raoults law: Henrys law:

PA = P*Ax A ; PB = k Bx B ;

x A 1 xB 0

[4] [5]

In above equations P*A is the vapor pressure of the pure solvent at the temperature of the experiment, k B is a proportionality constant, named Henrys constant, and x A and x B are the molar fractions of the solvent and the solute, respectively. The Henrys constant has pressure units. (It represents the vapor pressure of a solute in a hypothetical solution of concentration x B = 1; not a real pure solute, but a solution containing only solute. However, within the logic of this presentation it must be considered only as a proportional constant). A solution where Raoults law applies to the solvent and simultaneously the Henrys law applies to the solute is named an infinitely dilute solution. The adjective infinitely is given to emphasize the fact that above equations accomplish only for very dilute solutions, which is indicated in above equations as x A 1 and x B 0. As an example, for x A from 0.98 up to 1.00 and for x B from 0.05 down to 0.00, the real vapor pressures differs from the calculated with equations [4] and [5] in only a few percent, let say 3% percent, to a non significant difference. The colligative properties of infinitely dilute solutions. Similar limit equations, and closely related to the above ones, are the equations describing the vapor pressure decreases, the boiling point increases and the melting point decreases of a solvent when a non-volatile solute is added to it to form and infinitely dilute solution. Also, the limit equation describing the osmotic pressure of an infinitely dilute solution in contact with a pure solvent trough a semi-permeable membrane. All these properties depend on the amount of solute but not on its nature. Because of that they are collected together and named as colligative properties. For the sake of briefness will not go on a further discussion about them. The state equation of gases under limiting physical conditions. The equation relating the volume ( the pressure (P), the temperature (T) and the number of mole (n) of a gas is named V), the equation of state of that gas. All gases, pure or as a mixture, obey the following equation of state at certain temperature and pressure values: PV = nRT ; P 0; T [6]

Where R is a constant and equal to 0.082056 atm liter/(mole Kelvin) for any gas. It is a general experience that the agreement between a measured parameter value, let say pressure, and the corresponding one calculated with equation [6], improves in the extent that pressure goes down and temperature goes up. The agreement should be absolute, within the experimental errors, in the limiting values of both pressure and temperature. This fact is indicated in eqn. [6] as P 0 and T Many gases obey that state equation to within a few . percent (0.1 3%) at temperatures between 20 and 25C and at all pressure up to one atmosphere. The pressurebased chemical equilibrium constant. Let assume that in the chemical equilibrium reaction [1] all participants are gases. Then, the expression for its equilibrium constant is written as: K(P i) = PCc PDd / PAa PBb ; Pi 0, T [7]

In eqn. [7], Pi represents the partial equilibrium pressure of each one of the participants in the reaction mixture and the exponents are the respective reaction stoichiometry coefficients. It is normally said that above equation is valid at low pressures and high temperatures, which is indicated in equation [7] as Pi 0 and T From a practical point of view, this means that . equation [7] is valid (is good for the experimenters purposes) in the extent eqn. [6] is valid (it is acceptable for the experimenter).

II.

The Ideal Equations for Fictional (Model) Systems: generalizing the limiting situations

All limit equations we have mentioned so far, are equations for systems at their thermodynamics equilibrium. To make them general we claim to some of the fundamental equations of the thermodynamics and to the basic equations defining those systems. Phase equilibrium. At the phase equilibrium the chemical potential of a given component is the same in all phases of the system. This is true for closed systems and where only P-V work is possible. This law is expressed as follows: i= i [8]

In above equation, and are two phases of the systems at the equilibrium and i is a component present in both phases. Chemical equilibrium. At the chemical equilibrium the sum of chemical potentials of the reactants equals the sum of the chemical potential of the reactants. This is true for closed systems and where only P-V work is possible. This law is expressed as follows: RRR = RPP [9]

In equation [9] R is for the r eactants and P is for the products of the chemical reaction and the s are their respective stoichiometry coefficients. The s are the chemical potentials at the respective equilibrium concentrations. Chemical change at mechanical and thermal equilibrium. For a chemical reaction occurring at constant temperature and pressure, the corresponding Gibbs energy change, G, is given by: G = PPP - RRR [10]

Here, the Rs are the chemical potentials of the reactants at their initial concentrations and Ps are the chemical potentials of the products at their final concentrations.

The Gas Phase i) The system definition: the Ideal Gases Case 1: Pure gases. An ideal gas is defined as one that obeys eqn. [6] at all pressures and temperatures. So, restrictions to that equation are ruled out. Thus, equation [6] becomes a general equation, though of an ideal system. Consequently, eqn. [6] under this n condition is ew the Equation of State of Ideal Gases or (the mathematical expression of) the Law of Ideal Gases. We write it in the following way: PV = nRT ; any T and P [11]

All the physical properties of ideal gases are derived on the base of it. Case 2: Mixture of gases. An ideal gas mixture is defined as one having the two following properties: 1) The below sate equation is accomplished at all temperatures, pressures and compositions: PV = (i ni) RT = ntot R; any T and P [12]

In above equation, ni is the number of mole of i in the mixture, and ntot is the total number of mole 2) The equilibrium partial pressure of gas i in the mixture is equal to the pressure it would have if it were alone under the same conditions All the physical properties of ideal gas mixtures are derived upon above definition, which is a general definition (no restrictions to the temperature and pressure values exist). It introduces the assumptions that in pure ideal gases and ideal gas mixtures, molecular volumes are negligible small and that there are not molecular interactions, which is not true in real gases. Therefore, an ideal gas mixture is an idealized system. Nevertheless, at extreme low pressures and extreme high temperatures all real gases approach their behavior to the one given by the definition of ideal gas mixtures. (A complete definition of an ideal gas mixture additionally requires that if the gas mixture is formed by isothermally mixing the pure components, the heat of mixture is zero, assuming that no chemical reactions occur in the mixture. Nevertheless, this request is not necessary in this presentation). ii) The consequence: the ideal-pressure based chemical potential. On the base of the definition given above for an ideal gas mixture and of eqn. [8], applied to the equilibrium between an ideal pure gas i and an ideal gas mixture containing it, the fundamental thermodynamics equation for a mixture of ideal gases is obtained. It is expressed as follows: i = i (T) + RT ln (P i / P) ; any T and Pi [13]

In above equation, i is the chemical potential of the component i in the gas mixture at the temperature T, Pi is the partial pressure of i, and i is the chemical potential of the pure gas i at

the temperature T and the pressure of 1 atmosphere. P is the standard pressure of 1 atmosphere. Equation [13] is a general equation of an ideal system. It is applicable to real systems only if they are at low pressures and high temperatures. Equation [13] can also be written for a pure gas and all its thermodynamics properties deduced from it. In such a case, the sub index i is deleted in that equation and then i represents the chemical potential of the pure gas at the temperature T and the pressure P. i is the same as before.. iii) The results: the ideal equations. All thermodynamics equations of a mixture of ideal gases can be derived from eqn. [13]. For example, using this equation in the equation for the chemical equilibrium condition, [9], and making use of a basic thermodynamics equation [G = iii(T), a constant value at a fixed T] the equilibrium constant of an all ideal-gas phase reactions, as represented by reaction [1], is obtained: K(P i) = PCc PDd / PAa PBb ; any Pi and T [14]

Any pressure value in eqn. [14] is actually a ratio of the type (Pi / P)i (i is the reaction stoichiometry coefficient of the reactant i). However, the term P has been omitted for the sake of simplicity. Equation [14] is eqn. [7] but with no restrictions to the Pi and T values. Thus, eqn. [14] is a general equation, but of an idealized nature. Equation [14] applies to real systems only if they are at low pressures and high temperatures. In such a case, eqn. [14] becomes eq. [7]. The Liquid Phase Alternative I. The Ideal Solution i) The system definition: the ideal solution. An ideal solution is defined as one that obeys Raoults law, eqn. [4], for any of its components and at all values of their molar fractions. So, the restriction to that equation, x A 1, is ruled out. Thus, equation [4] becomes a general equation, though of an idealized system. Consequently, eqn. [4] under this new condition is the Equation of the Vapor Pressure of Ideal Solutions. We write it in the following way: Pi = P*ix i ; any i and x i [15]

From a molecular point of view, an ideal solution (solid or liquid) is one where all molecular interactions and volumes are the same and no volume neither heating change are expected when the solution components are combined. An example of this type of solution is the mixture of benzene and toluene. ii) The consequence: the concentrationbased chemical potential. On the base of eqn. [8], applied to the equilibrium between an ideal solution and its vapor, and of eqns. [13] and [15], the fundamental thermodynamics equation for an ideal solution is obtained. It is expressed as follows:

i = i + RT ln x i; for any component and any x i value i (T,P)= *i

[16]

In eqn. [16], i is the chemical potential of the component i of the solution at the molar fraction x i, and i is the chemical potential of pure i, indicated with the *, at T and P of the solution. To obtain eqn. [16] two approximations have been made: first, that the vapor in equilibrium with the solution is an ideal gas mixture; and, second, that *i is no sensitive to pressure changes. (An alternative way of introducing eqn. [16] is by just postulating it as the definition for the ideal solution. Roults law is then deduced, as must occur, since the definition for the chemical potential is oriented to reproduce it. However, that presentation is not within the logic of the present article). Equation [16] is a general equation, but of an idealized nature. It is applicable to real systems that accomplish Raoults law (x i 1). iii) The results: the ideal equations. For the sake of briefness, we will rather concentrate in the results obtained following Alternative two, below. Alternative II. The Ideal Dilute Solution i) The system definition: the ideal dilute solution. As we mentioned above, ideal solutions exist when the molecules of the several components are very similar to each other. However, quite few examples of this type of solutions are found in nature. But an ideal solution also exits when the molar fraction of the solvent approximates to 1, such that solute concentrations are very low. In these solutions, the molecules of the solute interact with the molecules of the solvent only, due to the high dilution of the solute. This is the case of most of the solutions. Therefore, we will develop an ideal model based on this fact. From above, our model must consider the existence of solutes and of a solvent. Hence, a concentration range needs to be established. We will take the one of the infinitely dilute solution and we will restrict the existence of solutions to that concentration interval only. In conclusion, our ideal model will be the infinitely dilute solution. Therefore, we define and ideal dilute solution as one where the solvent, A, obeys the Raoults law, eqn. [4], and each solute, i, obeys the Henrys law, eqn. [5]. We can rewrite them in a more formal way, as made elsewhere (4): Henrys law: Raoults law: Pi = k ix i PA = P*Ax A 1 - x A 1 and 0 1 for i A [17] [18] [19]

Equation [19] defines the molar-fraction concentration interval for the solvent, where x A differs from 1 in a very small amount, . Because in our model we have restricted the existence of solutions to that concentration interval only, eqns. [17] and [18] are general equations, but of an idealized system.

ii) The consequence: the concentrationbased chemical potentials. To obtain the fundamental thermodynamics equations for the dilute ideal solution, eqn. [8] is applied to the equilibrium between an ideal dilute solution and its vapor. When applied to the solute eqn. [17] is used while for the solvent eqn. [18] is used, both within the limits established by eqn. [19]. Then, and with the help of eqn. [13], the following expressions are obtained: i = i + RT ln x i , i i (T,P, x i= 1) A = A + RT ln x A, , A *A (T, P) for any solute (iA) [20]

for the solvent (A)

[21]

In eqn. [20], i is the chemical potential of the solute at the molar fraction x i, and i is the chemical potential of the solute standard state. The standard state is a hypothetical solution with x i = 1 and at the temperature, T, and pressure, P, of the actual solution. From a formal point of view the solute standard state is a hypothetical state because it exists beyond the limit imposed by eqn. [19], where eqns. [20] and [21] are strictly valid; it is beyond the reality we have accepted for our model. From a physical point of view, it is a hypothetical state because it has the properties of an infinitely dilute solution and not the properties of the pure solute (x i =1). In eqn. [21], A is the chemical potential of the solvent at the molar fraction x A, and A is the chemical potential of the solvent standard state. This state is the pure solvent at the temperature and pressure of the solution. To obtain eqns. [20] and [21], the approximation that the vapor in equilibrium with the solution is an ideal gas mixture has been made. (An alternative way of introducing eqn. [20] is by just postulating it as the definition for the ideal solution. Equation [21] is then deduced as a consequence of eqn. [20]. Henrys law is also deduced from eqn. [20], as must occur, since the definition for the solute chemical potential is oriented to reproduce it. Finally, Raoults law is deduced as a consequence of eqn. [21]. However, that presentation is not within the logic of the present article). Because of all the discussion above, eqn. [20] and [21] are general equations of an idealized system. They apply to real systems only if they are infinitely dilute solutions, where Henrys law and Raoults law are experimentally valid. iii) The results: the ideal equations. All thermodynamics equations of ideal and ideal dilute solutions are derived from eqn. [16] and from eqns. [20] and [21], respectively. The lasts are the more useful ones because they represent most of the real cases in a better way. Thus, replacing eqns. [20] and [13] in eqn. [8] for the equilibrium between an ideal dilute solution and its ideal gas vapor, Henrys laws is deduced. While doing the same but with eqn. [21] instead of eqn. [20], Raoults law is deduced. Replacing eqn. [20] in eqn. [9] for the chemical equilibrium occurring in an ideal dilute solution, and making use of an appropriate thermodynamics equation [G = iii(T), a constant value at a fixed T], the following equilibrium constant is obtained: K (x i) = x Cc x Dd / x Aa x Bb 1 - (1 x i) 1 and 0 1 [22]

10

For a dilute solution it is a ccomplished that x i Ci (103 MA / dA), where Ci is the molar concentration of the specie i, MA and dA are the molar mass and the density of the solvent, respectively. If that expression is replaced in eqn. [22], it is found that: K(x i) = (C Cc CDd / CAa CBb ) (103 MA / dA)(c + d a b) [23]

The concentration ratio in above equation must also be a constant and equal to K (C i) of eqn. [2], the concentrationsbased chemical equilibrium constant. In that equation, the notation Ci 0 is equivalent to the concentration borderline condition of eqn. [22]. Replacing eqn. [20] in eqn. [10] for the chemical reaction [1] occurring in an electrochemical cell, and making use of some basic thermodynamics equations [G = -nF and G = iii(T)], Nernst equation is deduced. It has the same mathematical shape as eqn. [3] but with x is instead of Cis. It is valid in the same concentration range than K (x i) above, eqn. [22]. If using the relationship between the molar fraction scale and the concentration scale, the Nernst equation as given by eqn. [3] is deduced (in this deduction = -G /nF - (c + d a b) (RT/nF) ln(103 MA / dA)]. Because all the equations mentioned in this section are derived from general equations of ideal systems, all of them are also general equations, though of ideal systems. They apply to real systems at the infinitely-dilute-concentration borderline condition only.

III. The General Equations for General Real Systems: back to reality From fiction to reality: the function A f. All equations above are general equations because they cover the idealized reality they have been invented for, with no restrictions to their applicability. Therefore all those equations are general ones but for ideal systems. What we need now is to see how we use those general equations to explain (or attempt to) the reality as a whole and exactly as it is. The strategy for above purpose is as follows. First, we recognize the reality exactly as it is, which means that there are no other restrictions to its existence than the natural ones. Second, we modify the general ideal equation by a factor-function containing all the significant information about the reality that it is ignored by the ideal system. From a practical viewpoint, the more complete we can construct that factor-function the most it approaches the ideal equation to the real equation. We will name that function as the approximation-to-reality factor,A f . The connection of the real equation, R, to the ideal equation, I , can be written in a general form as follows: R [24] Or R=I+Af [25] = A
f

11

The function A f is commonly referred as a measure of the deviation from ideality. However, in our context it seems to us as more appropriate to refer to it as a measure of the approximation to reality, because that is our purpose when we introduce it. The Gas Phase i) The system definition: the Real Gases. Real gases can be defined by modifying weather the state equations or the chemical potential equation of the ideal gas, eqns. [11] and [12] and eqn. [13], repsectively. As we mentioned, the last is the fundamental thermodynamics equation for pure ideal gases and for mixtures of ideal gases. We define the real gases by mean of the following equations: Alternative I. The Equation of State of Real Gases Pure gases: Mixture of gases: PV = A f* (PV)ideal = A f* nRT ; any T and P PV = A f (PtotV)ideal = A f ntotRT ; any T and P [26] [27]

In above equations: A f* = A f* (P, T) and A f = A f (P, T, xi). The reality factors recognize the influence of the temperature and of the pressure in the mathematical shape of the equations of state, and the importance of molecular interaction and gas volume, as well. This is explicitly indicated in eqn. [27] with the term x i (mixture composition). Equation [6] represents a limiting but real situation. Therefore, under our scheme it must be considered as a particular solution of the general eqns. [26] or [27]. Hence, it must occur that: limit A f* = 1 and limit A f = 1
Pi 0, T P 0, T

[28]

[29]

Appropriate approximation functions are those named virial equations of states. For pure gases, they are as follows: A f* = [1 + A2 (T)P + A3 (T)P 2 + A4 (T)P 3 + ] A f* = [1 + B2 (P)T-1 + B3 (P)T-2 + B4 (P)T-3 + ] [30] [31]

Both expressions satisfy eqn. [28], as it must be. For gas mixtures identical expressions are written. In this case the constants also depend on the mixture composition. Though not very much used, a strategy different to the ones given by eqns.[26] and [27] consist in modifying each term of the ideal equation. A well-known example is the famous van deer Waals equation: _ _ (P + a/V 2 )(V b) = RT [32] _

12

In above equation V = V/n, and a and b are constants that are different for different gases. The term a/V 2 corrects the effect of the intramolecular attractive forces over the gas pressure, while the term b corrects the effect of the molecular volumes over the volume available to gas motion, provided by the gas container. This equation is much better than the ideal gas equation but it is not satisfactory at high pressures. Constants in above equations are experimentally determined as explained elsewhere (4-6). All theoretical work concentrates in reproducing these constants to test different models for gases. Alternative II. The Chemical Potential of Real Gases Pure gases: = ideal + A f * = (T) + RT ln (P / P) + A f * ; any T and P ; any T and Pi [33] [34]

Mixture of gases: i = i, ideal + A f = i (T) + RT ln (P i / P) + A f

In eqn. [33] P is the pressure of the pure gas, while in eqn. [34] P is the total pressure of the gas mixture. Here, Pi is the contribution of component i to that pressure. Terms in both equations above have the same meaning than in eqn. [13]. Thus, the standard state for a real gas, pure or as a component of a mixture, is its state of pure gas at the pressure of 1 atmosphere, at the temperature of the experiment, and behaving as an ideal gas. Therefore, this state is a hypothetical one. But this is not a problem because the standard state is a reference state and as such it is arbitrarily chosen, and, in addition, its chemical potential value can be evaluated from experiments as indicated elsewhere (4-6). Let us concentrate in gas mixtures rather than in pure gases. Equations we will derive can be written for pure gases in exact ways, except that the sub index i is not written and that A f becomes A f *. For a reason that we will explain later on in this section it is convenient to take the reality factor as: A [35] In above expression, i = i (P, T, xi). As before, the reality factor recognizes the influence of temperature, pressure and composition in the shape of the equations of state. i is known as the fugacity coefficient of the gas or the fugacity coefficient of the component i in the gas mixture. It is a dimensionless parameter. i is in turn defined as follows: i = f i /Pi [36]
f

RT

ln

In above equation, f i = fi (P, T, xi). fi is named the fugacity of the gas or the fugacity of the component i in the gas mixture. It has pressure units. Equation [13] ( i for an ideal system) is applicable to a real system at a low pressure and high a temperature. In such a case, eqn. [13] must be considered as a particular solution of general equation [34] and therefore it must occur that: limit A f = 0 [37]

13
Pi 0, T

Which implies that: limit i = limit (f i / Pi) = 1


Pi 0, T Pi 0, T

[38]

And: limit f i = Pi
Pi 0, T

[39]

Equation [39] tells that at low pressures and high temperatures, the fugacity becomes the pressure. Observe that for ideal gases, pure or as gas mixture, i = 1 ii) The consequence: the fugacity-based chemical potential. If eqn. [36] is replaced in eqn. [35], and then this is replaced in eqn. [34], the following result is obtained: i = i (T) + RT ln (f i / P) ; any T and P [40]

Where f i =iPi. As eqn. [13] for an ideal gas mixture, eqn. [40] is the fundamental thermodynamics equation for a real gas mixture. Equation [40] can also be written for a pure gas and all its thermodynamics properties deduced from it. In such a case, the sub index i is deleted in that equation. Equation [39] tells that at low pressures and high temperatures eqn. [40] becomes eqn. [13], which is valid for real systems under that condition. All equations above would be senseless if i values were not available. Fortunately, i values can be evaluated from measurable properties of the gases (4-6). iii) The results: the real equations. The reality factor was defined in the way given by eqn. [35] to have an expression for the fundamental thermodynamics equation of real gases, eqn. [40], of identical mathematical shape of eqn. [13], the fundamental thermodynamics equation of ideal gases. In this way, we can profit of all the results obtained for ideal gases by just changing in those expressions the pressure terms by the corresponding fugacities. Therefore, the fugacity and fugacity coefficients concepts arise not from physical arguments but from a mathematical convenience. Thus, for an all gas-phase real chemical system the chemical equilibrium constant can be directly obtained from eqn. [14]: K(f i) = f Cc f Dd / f Aa f Bb ; at any T and P Or K(f i) = (P Cc PDd / PAa PBb ) ((Cc Dd / Aa Bb ) = K(P i) K(i) limit K(i) = 1
Pi 0, T

[41] [42] [43]

And

According to eqns. [43] and [42], the equilibrium constant for an all-gas-phase real chemical reaction under the low pressure and high temperature condition, is given by K(f i) = K(P i), as it is experimentally found, eqn. [7].

14

The Liquid Phase We have two alternative ways to obtain the equations for real liquid solutions. They arise because we have two solution models for our real systems. Obviously, we will choose the one that more approximates to the real system of our interest. Undoubtedly, the properties of real systems must be independent of any model but not the equations that evaluate them. Effectively, equations provide values for these properties, which for comparison purposes they just need to be relative to some common reference values. These reference values are arbitrarily chosen, including the possibility that values corresponding to hypothetical situations be used. For any alternative, our reasoning strategy will be exactly the same we developed in Alternative II for real gases Alternative I. From Ideal Solutions to Real Solutions. This alternative is used when the molar fraction of any of the components of the solution can varies in a wide concentration range. Most common cases are the solutions made of two or more liquids. The system definition: the Real Solutions. Any real solution (or any component of a real solution) must obey the following equation: i = i, ideal +A f = i + RT ln x i + A f ; for any component and x i values [44]

In above equation, A f = A f (P, T, xi). i, ideal is the expression of the chemical potential for the ideal solution components, its terms having the same meaning than in eqn. [16]. Thus, in the context of the ideal solution model, the standard state of any component of a real solution is its pure state at the temperature and pressure of the solution. We recall here that x i can have any value between 0 and 1, with no other restriction than ii = 1.We take the reality factor as: A [45] In above expression, i = i (P, T, xi). As before, the reality factor recognizes the influence of temperature, pressure and composition in the shape of the equations of state. i is known as the activity coefficient of the component i in the solution. It is a dimensionless parameter. i is in turn defined as follows: i = ai / x i [46]
f

RT

ln

(P,

T,

x i)

In above equation, ai = ai (P, T, xi). ai is named the activity of the component i in the solution. It is a dimensionless parameter. Equation [16] ( i for an ideal solution) is applicable to a real system that accomplishes Raoults law (x i 1). In such a case, eqn. [16] must be considered as a particular solution of general equation [44] and therefore it must occur that:

15

limit A f = 0
x i 1

[47]

Which implies that: limit i = limit (ai / x i ) = 1


xi 1

[48]

And:
xi 1

limit ai = x i

[49]

Equation [49] tells that when in a real solution xi 1, the corresponding activity becomes the molar fraction concentration. Observe that for ideal solutions i = 1 for any component and therefore that in those systems ai = xi always. Observe also that for the standard state i = 1 (for the standard state i = i and x i = 1; then, eqn. [44], i = i + RT ln 1 + RT ln i). ii) The consequence: the activity-based chemical potential. If eqn. [46] is replaced in eqn. [45], and then this is replaced in eqn. [44], the following result is obtained: i = i (T) + RT ln ai ; for any component and x i values [50]

Where ai =i x i. As eqn. [16] for ideal solutions, eqn. [50] is the fundamental thermodynamics equation for a real solution if the standard state definition for any of its components is the one defined for the ideal solution model. Equation [49] tells that at very low concentration values eqn. [50] becomes eqn. [16], which is valid for real systems under that condition. i values can be experimentally evaluated from measurable properties of the solutions in many cases (4-6). As before, all theoretical work concentrates in reproducing the activity coefficient values in order to test different models for solutions. iii) The results: the real equations. For the sake of briefness, we will rather concentrate in the results obtained following Alternative two, below. Alternative II. From Ideal Dilute Solutions to Real Solutions. This alternative is used when one of the component of the solution, the solvent, has properties that make it very different from the other components, the solutes. And these differences determine for this component a much higher concentration than the ones of the other components. Most common cases are the solutions where the solutes are gases or solids. The system definition: the Real Solutions. Any real solution (or any component of a real solution) must obey the following equation: i = i, ideal +A f = i + RT ln x i + A f ; for any component and x i values [51]

In above equation, A f = A f (P, T, xi). i, ideal is the expression of the chemical potential for any component of the ideal dilute solution, i. e., eqns. [20] and [21] in a unified equation, its terms

16

having the same meaning than in those equations. Thus when eqn. [51] is applied to any solute, i = i and i is the chemical potential of the solute standard state, while when applied to the solvent, i = A and i = A = A is the chemical potential of the solvent standard state. The standard states are defined as for eqns. [20] and [21]. Therefore, in the context of the idealdilute solution model, the solvent standard state is the pure solvent while the solute standard state is a hypothetical solution with x i = 1, both at the temperature and pressure of the actual solution. As we mentioned before, this solute standard state is a hypothetical state because it has the properties of an infinitely dilute solution and not the properties of the pure solute. Nevertheless, its chemical potential value can be evaluated from experiments as indicated elsewhere (4-6). Because equation [51] is a general equation for a real system, it must be valid for any component and in the whole range of composition. Therefore the restrictions imposed by eqn. [19] to eqns. [20] and [21] are ruled out in eqn. [51]. We take the reality factor as in Alternative I: A f = RT ln i (P, T, x i) [52]

In above expression, i = i (P, T, xi ). As before, the reality factor recognizes the influence of temperature, pressure and composition in the shape of the equations of state. i is the activity coefficient of the component i in the solution. It is a dimensionless parameter. i is defined as we did it before: i = ai / x i [53]

In above equation, ai = ai (P, T, xi). ai is the activity of the component i in the solution. It is also a dimensionless parameter. As mentioned before, eqns. [20] and [21] apply to infinitely dilute real solutions. On another side, eqn. [51] must be satisfied by any species in the real solution, either recognized as solute or as solvent. Therefore, eqns. [20] and [21] in the infinitely dilute concentration range must be particular solutions of eqn. [51]. Hence, it must occur that: limit A f = 0
x i = A 1, x i 0, i A

[54]

From now on, x i = A = x A. Equation [54] in turns implies that: limit i = limit (ai / x i ) = 1
xi 0, i A xA 1 xi 0, i A xA 1

[55] [56]

limit A = limit (aA / x A ) = 1

And: limit ai = x i
xi 0, i A

[57] [58]

limit aA = x A
xA 1

17

Equations [57] and {58] tell that in very dilute real solutions, the activities become the molar fraction concentrations. Observe from eqns. [51] and [52] that for ideal dilute solutions i = 1 and, therefore, ai = x i for any component (solutes and solvent) in those systems, always. Observe also that for the solvent standard state A = 1 (for the solvent standard state A = A and x A = 1; then, eqn. [51], A = A + RT ln 1 + RT ln A) and that for the solute standard state i = 1 (for the solute standard state i = i and x i = 1; then, eqn. [51], i = i + RT ln 1 + RT ln i). ii) The consequence: the activity-based chemical potential. If eqn. [53] is replaced in eqn. [52], and then this is replaced in eqn. [51], the following result is obtained: i = i (T) + RT ln ai ; for any component and x i values [59]

Where ai =i x i. As eqns. [20] and [21] for ideal dilute solutions, eqn. [59] is the fundamental thermodynamics equation for a real solution if the standard state definitions for any of its components are the ones defined for the ideal-dilute solution model. Equations [57] and [58] tell that for very dilute solutions eqn. [59] becomes eqns. [20] and [21] which are valid for real systems under that condition. i values can experimentally evaluated from measurable properties of the solutions in many cases (4-6). As before, all theoretical work concentrates in reproducing the activity coefficient values in order to test different models for solutions. iii) The results: the real equations. As for all cases before, the reality factor was defined, eqn. [52], to have an expression for the fundamental thermodynamics equation of real solutions, eqn. [59], of identical mathematical form of the fundamental thermodynamics equations developed for the ideal-dilute solution model, eqns. [20] and [21]. In this way, we can profit of all the results obtained for ideal dilute solutions by just changing in those expressions the concentration terms by the corresponding activities. Therefore, as the fugacity and fugacity coefficients concepts, the concepts of activity and activity coefficients arise just because of a mathematical convenience and not for physical reasons. Thus, for a solution-phase real chemical system the chemical equilibrium constant can be directly obtained from eqn. [22]: K(ai) = aCc aDd / aAa aBb ; Or at any x i values [60] [61] [62]

K(ai) = (x Cc x Dd / x Aa x Bb ) (Cc Dd / Aa Bb ) = K(ai) K(i) limit K(i) = 1


xi 0

And

According to eqns. [61] and [62], the equilibrium constant for a solution-phase real chemical reaction occurring under the dilute concentration condition, is given by K(ai) = K(x i), as it is experimentally found, eqn. [23] and, consequently, eqn. [2]. Following above procedure, Henrys and Raoults general equations for real solutions, and Nernst general equation for real electrochemical reactions are deduced. They have the same mathematical shapes than the equations obtained for ideal systems, except than concentration expressions are replaced by activities.

18

Need to mention here that standard chemical potential and activity coefficient values depend upon the concentration scale used. In that regard, their values must be combined with concentration values of the same concentration scale. Obviously, actual chemical potential values do not depend upon the concentration scale used neither on the election of the standard states for the solution components. Finally, we have not made any distinction between nonelectrolyte and electrolyte solutes, but to put above equations in an operational form extra complexity is introduced when dealing with electrolytes. This point, however, is beyond the scope of the present article. Conclusions Limits laws and equations give account of our difficulties to know and understand the whole reality and exactly as it is in a direct way. Or, in other words, they give account of the reality we can understand by direct experimentation, in most cases. Before that point, systems regularities have such a complexity that they are not directly observable or not comprehensible to the experimenter. Limits equations play a crucial role as the starting point for the understanding of the reality as a whole. They orient the design of simplified experimental models. These ideal systems incorporate the properties of the real limit systems, which we understand, and assume that they are the same in the whole range of existence of the idealized system. Applications of general equations from physics and chemistry to these fictional systems, provide with general equations for them. These ideal general equations are then modified by appropriate factors that approximate to us to the understanding of the whole reality and exactly as we think it is. These factors, which we have named approximation-to-reality factors, can be evaluated from measurable properties of the real systems. In this way, general equations for real systems are obtained. Above procedure, which we have named the sciences routine, is obviously implicit in the normal scientific work and in the textbooks. However, we feel it does not become evident to students in their long transit from the limits equations, they are introduced to at first year chemistry courses, to the general real equations, they know at upper courses. Neither it becomes evident from the reading of the hundreds of pages into which the method is implicit in the textbooks, which in no case is made in just one stroke. We believe that students should be introduced to the knowledge of the science routine (the science inductive method) at the very beginning of their learning of chemistry. First, to be aware of what part of reality they are knowing, and, second, to get familiar with the methods of science. Both aspects should help them to understand in a better way the different steps of their science learning. Finally, students in the upper course should be encouraged to a comprehensive discussion on the subject, as the one presented in this article. Or be given as a topic in a formal upper course. Literature Cited 1. Charles Mortimer, Qumica, Editorial Iberoamericana, Bogot, 5 ed.,1983 2. Raymond Chang, Qumica, McGraw-Hill Interamericana, Mxico, 6 ed., 1998 3. Ralph Petrucci, Williams Harwood, Qumica General: Principios y Aplicaciones Modernas, Prentice Hall, Madrid, 7 ed., 1999 4. Ira N. Levine, Fisicoqumica, Mcgraw-Hill Latinoamericana, Mxico, 1 ed., 1978 5. P. W. Atkins, Fisicoqumica, Fondo Educativo Interamericano, Mxico, 2 wd., 1985

19

6. Gilbert W. Castellan, Fisicoqumica, Addison Wesley Longman, Mxico, 2 ed., 1998 7. C. Contreras-Ortega, Nelson Bustamante, Juan Luis Guevara, Victor Kesternich1 and Carlos Portillo. This Journal, XXX, xxx, 2001 ______________________________ * Whom correspondence should be address to

S-ar putea să vă placă și