Chemical engineering thermodynamics II Topic 1: Solution Thermodynamics Theory - Applications are to gas mixtures and liquid solutions - Determination of thermodynamic properties for the above systems - Also applies to processes that involve composition changes such as mixing, separation, mass transfer operations and chemical reactions Fundamental property relations First law of thermodynamics: ð(ðð) = ðð + ðð (1) But ðð = ðð(ðð) and ðð = −ðð(ðð) This implies that: ð(ðð) = ðð(ðð) − ðð(ðð) (2) By definition: (ððŧ) = (ðð) + ð(ðð) and (ððš) = (ððŧ) − ð(ðð) This means: (ððš) = (ðð) + ð(ðð) − ð(ðð) (3) Differentiating each term in Equation 3, applying product rule, and substituting for ð(ðð), we obtain: ð(ððš) = (ðð)ðð − (ðð)ðð (4) This implies that for a closed system: (ððš) = ð(ð, ð) For an open system with N components: (ððš) = ð(ð, ð, ð1 , ð2 , ð3 , … . , ðð−1 , ðð ) Therefore: ð(ððš) = ( ð(ððš) ðð ) ð(ððš) ð,ð ðð + ( ðð ð(ððš) The last derivative in Equation 5, ( ððð ) ð(ððš) ð,ð ) ðð + ∑ð ( ððð ) ð,ð,ðð ððð (5) is called the chemical potential of species i in the ð,ð,ðð mixture or solution with symbol ðð Comparing Equation 5 with Equation 4 and substituting for ðð , we obtain: ð(ððš) = (ðð)ðð − (ðð)ðð + ∑ð ðð ððð (6) ð(ððš) Where (ðð) = ( ðð ) ð(ððš) ð,ð and (ðð) = − ( ðð ) ð,ð Equation 6 is the fundamental equation for solution thermodynamics which involves phase equilibria and chemical reaction equilibria. Partial properties - In a solution, each species contributes to the overall property of the system - Addition of a differential amount of species i to a finite amount of solution at constant temperature and pressure causes a change to the system properties ð(ðð) Ė ð = ( The partial molar property of species i in the mixture or solution is given by: ð ððð ) ð,ð,ðð Where ð is the solution property eg V, H, S, G Ė ð is the partial property of species i in the mixture or solution ð ðð is the pure-species property of species i at the solution temperature and pressure If ð = ðš, we have ðšĖ ð and this is equal to ðð Chemical potential and phase equilibria - When two or more phases are in equilibrium, temperature and pressure are uniform throughout the whole system - Equation 6 applies to species i in all phases in question Dealing with two phases, ðž and ð―, each extensive property of the whole system is a sum of it in the ðž phase and the ð― phase. This is represented as: (ðð) = (ðð)ðž + (ðð)ð― (7) For the ðž phase: ð(ððš)ðž = (ðð)ðž ðð − (ðð)ðž ðð + ∑ð ðð ðž ððð ðž For the ð― phase: ð(ððš)ð― = (ðð)ð― ðð − (ðð)ð― ðð + ∑ð ðð ð― ððð ð― Finding total differential change in Gibbs Free Energy according to Equation 7: ð(ððš) = (ðð)ðð − (ðð)ðð + ∑ ðð ðž ððð ðž + ∑ ðð ð― ððð ð― ð ð If the system is a closed system: ∑ð ðð ðž ððð ðž + ∑ð ðð ð― ððð ð― = 0 (8) If the system is not yet at equilibrium, there will be mass transfer of species i between the phases and this is represented by ððð ðž and ððð ð― in Equation 8. Therefore, any increase in the amount of species i in the ðž phase is equal to the corresponding decrease in the amount of species i in the ð― phase until equilibrium is established as shown in Figure 1. α phase Exchange of species i due to mass transfer β phase Figure 1 This means: ððð ðž = −ððð ð― ∑ð(ðð ðž − ðð ð― )ððð ðž = 0 This implies that: Since ððð ðž ≠ 0, this implies that ðð ðž − ðð ð― = 0 Therefore: ðð ðž = ðð ð― = âŊ = ðð ð for i = 1, 2, 3, ….., N and there are ð phases Conclusion: multiple phases at the same temperature and pressure are in equilibrium when the chemical potential of each species is the same in all phases. Gibbs/Duhem equation Thermodynamic properties of a homogenous phase: (ðð) = ð(ð, ð, ð1 , ð2 , ð3 , … . , ðð−1 , ðð ) Therefore: ð(ðð) = ( ð(ðð) ðð ) ð,ð ðð + ( ð(ðð) ðð ) ð,ð ð(ðð) ðð + ∑ð ( ððð ) ð,ð,ðð ððð Since the total number of moles in the system, n, is constant: ðð âđ ð(ðð) = ð ( ðð ) ðð ð,ðĨ ðð + ð ( ðð ) ð,ðĨ Ė ð ððð ðð + ∑ð ð (9) But ðð = ððĨð , meaning ððð = ðĨð ðð + ðððĨð . Also ð(ðð) = ððð + ððð Substituting into Equation 9 and simplifying gives: [ðð − ( ðð ðð Ė ð ððĨð ] ð + [ð − ∑ ðĨð ð Ė ð ] ðð = 0 ) ðð − ( ) ðð − ∑ ð ðð ð,ðĨ ðð ð,ðĨ ð ð The size of the system, n, and the differential change, dn, can be arbitrarily chosen and are therefore not equal to zero, it means the terms in the square brackets must be zero. ðð Therefore: ðð = ( ðð ) ðð ð,ðĨ ðð + ( ðð ) ð,ðĨ Ė ð ððĨð ðð + ∑ð ð Ė ð And: ð = ∑ð ðĨð ð (10) (11) Ė ð Multiplying Equation 11 by n gives (ðð) = ∑ð ðð ð Ė ð + ∑ð ð Ė ð ððĨð Differentiating Equation 11: ðð = ∑ð ðĨð ðð Substituting this result into Equation 10 and simplifying gives: ðð ( ðð ) ðð ð,ðĨ ðð + ( ðð ) ð,ðĨ Ė ð = 0 ðð − ∑ð ðĨð ðð (12) Equation 12 is the Gibbs/Duhem equation. At constant temperature and pressure, the Gibbs/Duhem equation reduces to: Ė ð = 0 ∑ð ðĨð ðð (13) Equations 11 and 13 show that each species in a mixture or solution cannot have its own private properties. This is due to molecular interactions within the system. However, partial properties have all the characteristics of individual species properties as they exist in solution. Ė 1 ðĨ1 + ð Ė 2 ðĨ2 Applying Equation 11 to a binary system: ð = ð Ė 1 ððĨ1 + ðĨ1 ðð Ė 1 + ð Ė 2 ððĨ2 + ðĨ2 ðð Ė 2 Differentiating gives: ðð = ð Ė 1 + ðĨ2 ðð Ė 2 = 0 (from Gibbs/Duhem equation) But ðĨ1 ðð Ė 1 ððĨ1 + ð Ė 2 ððĨ2 Therefore: ðð = ð (14) Since ðĨ1 + ðĨ2 = 1, this means ððĨ1 = −ððĨ2 âđ ð ðī Ė ð −ðī Ė ð = ðī ð ðð Ė 1 ðĨ1 + ð Ė 2 ðĨ2 and using ðĨ1 + ðĨ2 = 1 Back to Equation 14: ð = ð Simplifying yields the following equations: Ė 2 = ð − ðĨ1 ðð ð ððĨ (15) Ė 1 = ð + ðĨ2 ðð ð ððĨ (16) 1 1 Ė ð = ðð To calculate pure-species property of species i: lim ð = lim ð ðĨð →1 ðĨð →1 Ė ð = ð Ė ð ∞ To calculate property when species i approaches infinite dilution: lim ð ðĨð →0 Relations among partial properties ððšĖ ð ð ð(ððš) ð ð(ððš) ð ð(ðð) (ðð) = ( ( ) = ( ) = ( )= ) = ðĖ ð ðð ð,ðĨ ðð ððð ð,ð,ð ððð ðð ððð ððð ð,ð,ð ð ð ððšĖ ð ð ð(ððš) ð ð(ððš) ð ð(ðð) (−ðð) = (− ( ) = ( ) = ( )= ) = −ððĖ ðð ð,ðĨ ðð ððð ð,ð,ð ððð ðð ððð ððð ð,ð,ð ð ð Therefore: ððšĖ ð = ðĖ ð ðð − ððĖ ðð Conclusion: the same relations on molar properties apply to partial properties Ideal gas model ðð ðð = ð ð âđ ð ðð = ð ð⁄ð This shows that all ideal gases (be it a single component or mixture) have the same molar volume when at the same temperature and pressure Determining partial molar volume ð(ðð ðð ) ð(ð ð ð⁄ð) ð ð ðð ðð ( ) ðĖ ð = ( ) = ( ) = ððð ðð ð ððð ð ð ð,ð,ð ð,ð,ð ð ðð But ð = ðð + ∑ð ðð ðð This means: ðĖ ð = âđ (ðð ) ð ð ð =1 ðð ð ð ð ðð Therefore: ðĖ ð = ð ðð = ðð ðð (17) Determining partial pressure ðð = ðĶð ð = ðĶð ð ð ðð ðð = ðĶð ð ð ð ðð Conclusion: since the ideal gas model assumes molecules with zero volume (volume of molecules is negligible compared to volume of container), which do not interact (no intermolecular forces, elastic collisions between molecules), other thermodynamic properties other than molar volume of the constituent species or components are independent of one another. Each species has its own set of private properties Gibbs theorem A partial molar property (other than volume) of a constituent species in an ideal gas mixture is equal to the corresponding molar property of the species as a pure ideal gas at the mixture temperature but at a pressure equal to its partial pressure in the mixture. Ė ð ðð (ð, ð) = ðð ðð (ð, ðð ) where ð ≠ ð This means: ð (18) Determining partial molar enthalpy The enthalpy of an ideal gas depends on temperature only. This means it is independent of pressure. Ė ð ðð (ð, ð) = ðŧð ðð (ð, ðð ) = ðŧð ðð (ð, ð) ∴ ðŧ Ė ð ðð = ðŧð ðð âđ ðŧ (19) Determining partial molar entropy From the 1st law: ðð = ðð − ðð Substituting: ðð = ðķð ðð, ðð = −ððð, ð = We obtain: ððð ðð = ðķð ð ðð + ð ð ð ð ð ðð , ððð ðð = ðð ð ðð At constant temperature: ðð = 0 ð 1 Also: ð ðð = ð ln ð = ln ðððððð − ln ðððððĄððð = ln ð ððððð ððððĄððð ðððððð From ideal gas law: ð ððððĄððð 1 = ðððððĄððð ð ðððððð ðððððð −1 = (ð ð ððððĄððð −1 Therefore: ð ðð = ln ð ððððð = ln (ð ððððð ) ððððĄððð ððððĄððð ) ð = − ln ð ððððð = −ð ln ð ððððĄððð This means: ððð ðð = −ð ln ð ð Integrating from Pi to P: ðð ðð (ð, ð) − ðð ðð (ð, ðð ) = −ð ln ðĶ ð = ð ln ðĶð ð âđ ðð ðð (ð, ðð ) = ðð ðð (ð, ð) − ð ln ðĶð ðð Applying Equation 17: ððĖ (ð, ð) = ðð ðð (ð, ðð ) ðð Therefore: ððĖ (ð, ð) = ðð ðð (ð, ð) − ð ln ðĶð ∴ ððĖ ðð = ðð ðð − ð ðð ðĶð (20) Determining partial Gibbs Free Energy ðð Ė ð ðð − ðððĖ ðð Since ðš = ðŧ − ðð, this means: ðšĖ ð = ðŧ ðð âđ ðšĖ ð = ðð ðð = ðŧð ðð − ð(ðð ðð − ð ln ðĶð ) = (ðŧð ðð − ððð ðð ) + ð ð ln ðĶð ∴ ðð ðð = ðšð ðð + ð ð ðð ðĶð Applying Equation 11 to Equations 19, 20 and 21 yields the following results: (21) ðŧ ðð − ∑ð ðĶð ðŧð ðð = 0 (22) ð ðð − ∑ð ðĶð ðð ðð = −ð ∑ð ðĶð ðð ðĶð (23) ðš ðð − ∑ð ðĶð ðšð ðð = ð ð ∑ð ðĶð ðð ðĶð (24) The terms on the LHS of Equations 22, 23 and 24 represent property changes of mixing of ideal gases, as illustrated in the next topic. Alternative expression for chemical potential of species i in an ideal gas mixture ððšð ðð = ðð ðð ðð − ðð ðð ðð ð ð At constant temperature and using ðð ðð = ð ð⁄ð : ððšð ðð = ð ðð = ð ðð ln ð Integrating both sides: ðšð ðð = ðĪð (ð) + ð ð ðð ð (25) where Γð (ð) is an integration constant that depends on the species and is a function of temperature only. Substituting into Equation 21: ðð ðð = ðĪð (ð) + ð ð ðð ð + ð ð ðð ðĶð = ðĪð (ð) + ð ð ðð(ðĶð ð) (26) Fugacity and fugacity coefficient: Pure species Fugacity is defined as the effective partial pressure of a real gas. For a real gas: ðšð = ðĪð (ð) + ð ð ðð ðð (27) Comparing Equation 27 with Equation 25: ð = ðð ðð ð Subtracting Equation 25 from Equation 27: ðšð − ðšð ðð = ðšð ð = ð ð ðð ( ðð ) (28) The term in the brackets in Equation 28 is called the fugacity coefficient of pure species i which is: ðð = ðð ð For an ideal gas: ðð = ð, meaning ðð = 1 and also ðšð ð = 0 Vapour/Liquid equilibrium for pure species For saturated vapour: ðšð ðĢ = Γð (ð) + ð ð ln ðð ðĢ For saturated liquid: ðšð ð = Γð (ð) + ð ð ln ðð ð Change of phase from saturated liquid to saturated vapour or vice versa occurs at T and Pisat At equilibrium: ððš = ðšð ðĢ − ðšð ð = 0 But ðšð ðĢ − ðšð ð = ð ð ln ( ðð ðĢ ðð ð ) This means at equilibrium: ðð ðĢ = ðð ð = ðð ð ððĄ Therefore: ðð ð ððĄ = ðð ð ððĄ ⁄ ð ððĄ ðð Fugacity and fugacity coefficient: Species in solution - Applies when species is in a mixture of real gases or a solution of liquids ðð = ðĪð (ð) + ð ð ðð ðĖð (29) For phase equilibrium the chemical potential of species i in all phase must be equal. This means ðž ð― ð that: ðĖð = ðĖð = âŊ = ðĖð This is the criterion for equilibrium that is used to solve phase equilibrium problems. ðĢ ð For vapour/liquid equilibrium: ðĖð = ðĖð This is the criterion for equilibrium used to solve vapour/liquid equilibrium problems. Ė ð ð Subtracting Equation 26 from Equation 29: ðð − ðð ðð = ðšĖ ð = ð ð ðð (ðĶ ðð) ð (30) The term in the brackets in Equation 30 is called the fugacity coefficient of species i in solution Ė ð which is: ðĖð = ðĶ ðð ð For an ideal gas: ðĖð = ðĶð ð, meaning ðĖð = 1 and also ðšĖ ð ð =0 Ideal solution model The ideal solution model equation is analogous to Equation 21. ðð ðð ðð = ðšĖ ð = ðšð (ð, ð) + ð ð ðð ðĨð (31) Determining partial molar volume for species i in an ideal solution It is given by: ðĖ ð ðð = ( ðð ððšĖ ð ðð ððš ) ð,ðĨ = ( ððð ) = ðð (32) ð Determining partial molar entropy for species i in an ideal solution ððĖ ðð = −( ðð ððšĖ ð ðð ) ð,ðĨ ððš = − ( ððð ) − ð ðð ðĨð = ðð − ð ðð ðĨð (33) ð Determining partial molar enthalpy for species i in an ideal solution Ė ð ðð = ðšĖ ð ðð + ðððĖ ðð ðŧ Ė ð ðð = ðšð + ð ð ðð ðĨð + ððð − ð ð ðð ðĨð = ðŧð âđ ðŧ (34) Applying Equation 11 to Equations 31, 32, 33 and 34 yields the following results: ð ðð = ∑ð ðĨð ðð (35) ðŧ ðð − ∑ð ðĨð ðŧð (36) ð ðð = ∑ð ðĨð ðð − ð ∑ð ðĨð ðð ðĨð (37) ðš ðð = ∑ð ðĨð ðšð + ð ð ∑ð ðĨð ðð ðĨð (38) Lewis/Randall Rule ðĖ Subtracting Equation 27 from Equation 29: ðð = ðšð + ð ð ðð (ðð ) (39) ð Applying Equation 39 to an ideal solution: ðð ðð Comparing Equation 40 with Equation 31: ðĨð = Ė ðð ðð ð = ðšĖ ð = ðšð + ð ð ðð ( ðð ) ð ðð ðĖð ðð ðð âđ ðĖð = ðĨð ðð (40) (41) Equation 41 is called the Lewis/Randall Rule. It applies to each species in an ideal solution at all conditions of temperature, pressure and composition. ðð is determined when pure species i is in the same physical state as the solution and at the same temperature and pressure. The Lewis/Randall Rule is analogous to Raoult’s Law. Dividing Equation 41 by ððĨð : ðð ðĖð ððĨð = ðð ð ðð âđ ðĖð = ðð (42) Excess properties Liquid solution thermodynamic properties are dealt with easily by their departure from ideal solution behaviour. At the same temperature, pressure and composition: ððļ = ð − ððð The same applies for partial properties ðð = ðšĖ ð = ðĪð (ð) + ð ð ðð ðĖð (29) ðð Applying Equation 29 to an ideal solution: ðð ðð = ðšĖ ð = ðĪð (ð) + ð ð ðð ðĨð ðð (43) Ė ðð ðļ ð Subtracting Equation 43 from Equation 29: ðšĖ ð − ðšĖ ð = ðšĖ ð = ð ð ðð (ðĨ ðð ) (44) ð ð ðĖ The term in the brackets in Equation 44 is called activity coefficient: ðūð = ðĨ ðð ð ð The activity coefficient accounts for liquid phase non-idealities. Summability (Equation 11) and Gibbs/Duhem equation (Equation 13) can be applied to Equation 44. Topic 2: Solution Thermodynamics Applications We will focus on property changes of mixing Δð = ð − ∑ ðĨð ðð ð Recall equation for excess properties: ððļ = ð − ððð Excess molar Gibbs Free: ðš ðļ = ðš − ðš ðð = ðš − ∑ð ðĨð Gi − ð ð ∑ð ðĨð ln ðĨð = Δðš − ð ð ∑ð ðĨð ln ðĨð Excess molar entropy: ð ðļ = ð − ð ðð = ð − ∑ð ðĨð Si + ð ∑ð ðĨð ln ðĨð = Δð + ð ∑ð ðĨð ln ðĨð Excess molar enthalpy: ðŧ ðļ = ðŧ − ðŧ ðð = ðŧ − ∑ð ðĨð Hi = Δðŧ Excess molar volume: ð ðļ = ð − ð ðð = ð − ∑ð ðĨð Vi = Δð Heat effects of mixing processes The enthalpy change of mixing is also called heat of mixing. This is analogous to the heat of reaction though it is relatively smaller in magnitude. Heats of solution Heat effect when solids or gases are dissolved in liquids. It is based on 1 mole of solute and has Ė . The enthalpy change of mixing, Δðŧ is the heat effect per mole of solution (mol of the symbol Δðŧ solution = mol of solute + mol of solvent) ððð ðð ð ðððĒðĄð If species 1 is the solute: ðĨ1 = ððð ðð ð ðððĒðĄððð Therefore, ratio of mol of solvent per mol of solute is: ðĖ = Ė = Hence: Δðŧ âðŧ ðĨ1 = 1−ðĨ1 1 âđ ðĨ1 = 1+ðĖ ðĨ1 âðŧ 1+ðĖ Solution processes are largely physical changes and can be represented as: For an anhydrous salt: ðŋððķð(ð ) + 12ðŧ2 ð(ð) → ðŋððķð(12ðŧ2 ð) For a hydrated salt: ðŋððķð â 5ðŧ2 ð(ð ) + 7ðŧ2 ð(ð) → ðŋððķð(12ðŧ2 ð) Ė =âŊ Δðŧ Ė =âŊ Δðŧ Heat of formation of solute in solvent, for example of LiCl in 12 mol of H2O is represented as: 1 ðŋð + ðķð2 + 12ðŧ2 ð(ð) → ðŋððķð(12ðŧ2 ð) 2 This is a result of adding the enthalpy change or heat of formation of LiCl and the heat of solution of 1 mol of LiCl and 12 mol of H2O Heat of formation of a hydrated salt, for example LiCl·5H2O is the sum of the heat of formation of LiCl and the heat of formation of the water of hydration and is given by: 1 5 ðŋð + ðķð2 + 5ðŧ2 + ð2 → ðŋððķð â 5ðŧ2 ð(ð ) 2 2 The heats of solution can be obtained by combination of the above reactions or graphically from a Ė vs ðĖ diagram. Figure 2 shows heats of solution for HCl(g) and LiCl(s). Δðŧ Figure 2 An enthalpy/concentration diagram, H vs x, makes it easier to determine the enthalpy of a solution with a certain concentration at a particular temperature. Figure 3 shows the variation of enthalpies of NaOH solutions of different concentrations with temperature. Figure 3 For adiabatic mixing of two solutions of different concentrations, the final solution lies on a straight line that joins the two initial solutions. Topic 3: Vapour/Liquid Equilibria Phase rule and Duhem theory Phase rule: ðđ = 2 − ð + ð where F is the number of intensive variables that can be independently fixed π is the number of phases in the system N is the number of chemical species or components in the system Duhem theory: for any closed system formed initially from given masses of prescribed chemical species, the equilibrium state is completely determined when any two independent variables are fixed. They may be intensive or extensive. Phase diagrams - P vs x,y phase diagram at constant T - T vs x,y phase diagram at constant P You should know the different regions, curves and important points. • Dew point curve is the P or T vs y curve and represents all saturated vapour mixtures (vapour at boiling point). • Bubble point curve is the P or T vs x curve and represents all saturated liquid mixtures (liquid at boiling point). • Mixtures boil over a range of temperatures hence the bubble point curve is different from the dew point curve. Between the bubble point curve and the dew point curve exist two phases, saturated vapour and saturated liquid that are at equilibrium. The equilibrium compositions are determined by tie line (horizontal lines drawn at a certain temperature or pressure). The overall composition of the mixture does not change in this region and can be determined using the lever rule. For a binary mixture, both curves begin and terminate at saturation points of the pure species. • Superheated vapour mixtures exist at temperatures above dew point temperature of the mixture or at pressures below the dew point pressure of the mixture. • Subcooled liquid mixtures exist at temperatures below bubble point temperature of the mixture or at pressures above the bubble point pressure of the mixture. Simple models for Vapour/Liquid Equilibria (VLE) - Raoult’s law - Henry’s law - Modified Raoult’s law Raoult’s law - Applicable at low to moderate pressures - The major assumptions are: 1) the vapour phase is an ideal gas 2) the liquid phase is an ideal solution. An ideal solution is formed by species that are chemically similar (in terms of magnitude and type of intermolecular forces). Therefore, the species must not be too different in size and must be of the same chemical nature such as adjacent members of a homologous series - Applies to species of known vapour pressure which requires the species to be subcritical, that is at a temperature below critical temperature, Tc The equation for Raoult’s law is: ðð = ðĶð ð = ðĨð ðð ð ððĄ ðð ð ððĄ is the vapour pressure of pure species i at the system temperature This shows that the graph of ðð vs ðĨð is a straight line, where ðð = 0 when ðĨð = 0 and ðð = ðð ð ððĄ when ðĨð = 1. For a binary system: when ðĨ1 = 0, ðĨ2 = 1, ð1 = 0, ð2 = ð2 ð ððĄ and ð = ð2 ð ððĄ When ðĨ1 = 1, ðĨ2 = 0, ð2 = 0, ð1 = ð1 ð ððĄ and ð = ð1 ð ððĄ Therefore the graph of P vs x1: ð = ðĨ1 ð1 ð ððĄ + ðĨ2 ð2 ð ððĄ = ð2 ð ððĄ + ðĨ1 (ð1 ð ððĄ − ð2 ð ððĄ ) is a straight line since ð1 ð ððĄ and ð2 ð ððĄ are constants at a particular temperature. Deviations from Raoult’s law There are two types of deviations from Raoult’s law: - Negative deviation - Positive deviation Negative deviation When the bubble point pressure curve lies below the Raoult’s law line. This means Pactual < Pideal. If the deviations are sufficiently large for a binary mixture of liquids 1 and 2, the 1—2 intermolecular forces will be stronger than those of both pure liquids 1—1 and 2—2 and a minimum point is formed on the bubble point pressure curve. The dew point pressure curve also exhibits a minimum at the same point such that ðĨ1 = ðĶ1 . A boiling liquid (saturated liquid) of this composition produces a saturated vapour of exactly the same composition, behaving like a pure liquid hence no separation by distillation is possible for this constant boiling solution. This solution is called an azeotrope and for a solution that shows negative deviation from Raoult’s law, it is called a minimum pressure azeotrope or a maximum boiling azeotrope. Positive deviation When the bubble point pressure curve lies above the Raoult’s law line. This means Pactual > Pideal. If the deviations are sufficiently large for a binary mixture of liquids 1 and 2, the 1—2 intermolecular forces will be weaker than those of both pure liquids 1—1 and 2—2 and a maximum point is formed on the bubble point pressure curve. The dew point pressure curve also exhibits a maximum at the same point such that ðĨ1 = ðĶ1. This solution is called a maximum pressure azeotrope or a minimum boiling azeotrope. If 1—1 and 2—2 are so strong such that complete miscibility of the liquids is prevented, the system forms two separate liquid phases over a range of compositions, which is the basis of liquid-liquid equilibrium. Solving VLE problems using Raoult’s law ðĩ ð The Antoine equation gives the variation of ðð ð ððĄ with T: ln ðð ð ððĄ (ððð) = ðīð − ð(â)+ðķ ð There are four classes of problems: 1) BUBL P: Calculate yi and P given xi and T 2) DEW P: Calculate xi and P given yi and T 3) BUBL T: Calculate yi and T given xi and P 4) DEW T: Calculate xi and T given yi and P Solving 1st class of problems Step 1: At the given T, calculate Pisat for each species Step 2: Since xi is also given, P is calculated as follows: ðĶð ð = ðĨð ðð ð ððĄ ∑ð ðĶð ð = ð ∑ð ðĶð = ∑ð ðĨð ðð ð ððĄ Summing both sides over all i: But ∑ð ðĶð = 1, therefore: ð = ∑ð ðĨð ðð ð ððĄ (45) Step 3: Calculate yi of the required species using Raoult’s law Solving 2nd class of problems Step 1: At the given T, calculate Pisat for each species Step 2: Since yi is also given, P is calculated as follows: ðĶð ð = ðĨð ðð ð ððĄ ðĶð âđ ðĨð = ð ðð ððĄ ð ðĶð But ∑ð ðĨð = 1, therefore: ð = ðĶ ð ð ∑ð ðĨð = ∑ð ð ððĄ = ð ∑ð ð ð ððĄ ð Summing both sides over all i: ð ð 1 (46) ðĶ ð ∑ð ð ððĄ ð ð Step 3: calculate xi of the required species using Raoult’s law Solving 3rd class of problems Making T the subject of the formula in the Antoine equation: ðð ð ððĄ (â) = For a binary system, Equation 45 can be written as: ð2 ð ððĄ = where ðž = ð1 ð ððĄ ð2 ð ððĄ ðĩð ðīð −ðð ð (ððð) − ðķð ð (47) (48) ðĨ1 ðž+ðĨ2 ðĶ ⁄ðĨ which is the relative volatility using Raoult’s law, that is, ðž12 = ðĶ1⁄ðĨ1 2 2 ðĩ ðĩ 1 2 âđ ðð ðž = ðð ð1 ð ððĄ − ðð ð2 ð ððĄ = ðī1 − ð+ðķ − (ðī2 − ð+ðķ ) 1 2 (49) The iterative steps for T are as follows: Step 1: Calculate T1sat and T2sat using Equation 47 Step 2: Select any temperature between T1sat and T2sat and calculate P1sat and P2sat using the Antoine equation at that temperature Step 3: Calculate α0 using the P1sat and P2sat determined in Step 2 Step 4: Calculate P2sat using α and Equation 48 Step 5: Calculate T2sat using P2sat as P in Equation 47 Step 6: Calculate new α using T2sat as T in Equation 49 Step 7: Repeat steps 4 to 6 until T2sat converges. This is the required temperature, T. Step 8: Calculate P1sat using T in step 7 and the Antoine equation Step 9: Calculate y1 using Raoult’s law Solving 4th class of problems For a binary system, Equation 46 can be written as: ð1 ð ððĄ = ð(ðĶ1 + ðĶ2 ðž) where ðž = (50) ð1 ð ððĄ ð2 ð ððĄ The iterative steps for T are as follows: Step 1: Calculate T1sat and T2sat using Equation 47 Step 2: Select any temperature between T1sat and T2sat and calculate P1sat and P2sat using the Antoine equation at that temperature Step 3: Calculate α0 using the P1sat and P2sat determined in Step 2 Step 4: Calculate P1sat using α and Equation 50 Step 5: Calculate T1sat using P1sat as P in Equation 47 Step 6: Calculate new α using T1sat as T in Equation 49 Step 7: Repeat steps 4 to 6 until T1sat converges. This is the required temperature, T. Step 8: Calculate P1sat using T in step 7 and the Antoine equation Step 9: Calculate x1 using Raoult’s law Henry’s law - Applies to species whose Tc is less than operating temperature (these species do not have vapour pressure at the operating temperature). - Pressure is low enough such that the vapour phase is assumed to be an ideal gas - The species is also present as a very dilute solute in the liquid or approaches infinite dilution Henry’s law is given by: ðĶð ð = ðĨð ðŧð Hi is the Henry’s law constant of species i. The larger the Henry’s law constant of the species, the less soluble it is in the liquid phase. Modified Raoult’s law The liquid phase is not an ideal solution. The law is given by: ðĶð ð = ðĨð ðūð ðð ð ððĄ The activity coefficient, ðūð can be modelled by various correlations from Solution Thermodynamics Applications as a function of composition and one of them is: ln ðū1 = ðīðĨ2 2 and ln ðū2 = ðīðĨ1 2 where ðī = ðī(ð) We will focus on three classes of problems: 1) BUBL P: Calculate yi and P given xi and T 2) DEW P: Calculate xi and P given yi and T 3) Azeotropic composition and pressure at a given T Solving 1st class of problems Step 1: At the given T, calculate Pisat for each species and A if A is given as a function of T. Step 2: At the given xi, calculate ðūð for each species Step 3: Since xi is given, P is calculated as follows: ðĶð ð = ðĨð ðūð ðð ð ððĄ ∑ð ðĶð ð = ð ∑ð ðĶð = ∑ð ðĨð ðūð ðð ð ððĄ Summing both sides over all i: But ∑ð ðĶð = 1, therefore: ð = ∑ð ðĨð ðūð ðð ð ððĄ (51) Step 4: Calculate yi of the required species using the modified Raoult’s law Solving 2nd class of problems Step 1: At the given T, calculate Pisat for each species and A when A is given as a function of T. Step 2: Since xi is not given, it is not possible to calculate the activity coefficient of each species. An iterative procedure is required where the initial values of each activity coefficient are determined by assuming an ideal solution, that is each ðūð = 1 Step 3: Since yi is also given, P is calculated as follows: ðĶð ðĶð ð = ðĨð ðūð ðð ð ððĄ âđ ðĨð = ðū ðð ð ððĄ ð ð ðĶð But ∑ð ðĨð = 1, therefore: ð = ðĶ ∑ð ðĨð = ∑ð ð ð ððĄ = ð ∑ð ðð ððĄ ðūð ðūð Summing both sides over all i: ð ð ð ð 1 ∑ð ðĶð ðūð ðð ð ððĄ Step 4: Calculate xi of each species using the modified Raoult’s law Step 5: Calculate the new activity coefficient of each species using the given correlation. Step 6: Repeat steps 3 to 5 until P converges. From this value of P, determine the required xi (52) Solving 3rd class of problems ðĶ ⁄ðĨ An azeotrope exists when ðĨð = ðĶð . This means: ðž12 = ðĶ1⁄ðĨ1 = 1 2 2 ðū ð ð ððĄ From modified Raoult’s law: ðž12 = ðū1 ð1ð ððĄ 2 2 First of all, a check must be done on whether an azeotrope exists. ðž12 is a continuous function of x1 therefore an azeotrope exists when for 0 ≤ ðĨ1 ≤ 1, ðž12 at one limit is less than 1 and ðž12 at the other limit is greater than 1. This ensure that the graph of ðž12 vs x1 definitely passes through 1. Otherwise, an azeotrope does not exist. ðū ð ð ððĄ Since for an azeotrope ðū1 ð1ð ððĄ = 1, each activity coefficient is given by a correlation as a function 2 2 of composition and each ðð ð ððĄ can be determined from the Antoine equation at the given temperature, the azeotropic composition can be calculated. The azeotropic pressure is then calculated using the modified Raoult’s law at the azeotropic composition. Topic 4: Phase equilibria Our focus will be on osmotic equilibrium and osmotic pressure The equilibrium criterion is always: (ððš ðĄ ) ð,ð = 0 ððĖ For stability: (ððš ðĄ ) ð,ð < 0. This leads to another stability criterion which is: ððĨð > 0 at constant ð T and P. Figure 4 Species 1 is the solute and species 2 is the solvent. The membrane is permeable to species 2 (solvent) only. If ð′ = ð, the stability criterion applied to species 2 is applicable: âđ ðĖ2 (ð, ð′ = ð, ðĨ2 < 1) < ðĖ2 (ð, ð, ðĨ2 = 1) ððĖ2 ððĨ2 >0 but for the RHS, since ðĨ2 = 1, ðĖ2 = ð2 âđ ðĖ2 (ð, ð′ = ð, ðĨ2 < 1) < ð2 (ð, ð) This difference in fugacities represents a driving force for mass transfer, where the solvent diffuses from right to left through the semi-permeable membrane in Figure 4. This is called osmosis. This mass transfer increases pressure on the left side until equilibrium is established at a new pressure P*. Therefore: ðĖ2 (ð, ð′ = ð∗ , ðĨ2 < 1) = ð2 (ð, ð) (equilibrium criterion) The pressure difference ð∗ − ð is called osmotic pressure of the solution with symbol Π. Since T is constant: ðĖ2 (ð + Π, ðĨ2 ) = ð2 (ð) Taking the LHS: ðĖ2 (ð + Π, ðĨ2 ) = ð2 (ð) â Looking at the first ratio: ðĖ2 (ð,ðĨ2 ) ð2 (ð) Looking at the second ratio: ðĖ2 (ð,ðĨ2 ) ðĖ2 (ð+Π,ðĨ2 ) â ðĖ (ð,ðĨ ) ð2 (ð) 2 2 = ðĨ2 ðū2 ðĖ2 (ð+Π,ðĨ2 ) ðĖ2 (ð,ðĨ2 ) (from the definition of ðūð ) is called the Poynting factor Recall: ðð = ðĪð (ð) + ð ð ðð ðĖð (29) Ė ðð Differentiating with respect to P: ( ððð ) ð,ðĨ ð ln ð = ðĖ ð = ð ð ( ðð ð ) ð,ðĨ ðĖ 2 ð ln ðĖ2 = ð ð ðð âđ Separating variables and integrating from ð to ð + Π ð+Π,ðĨ2 âđ |ln ðĖ2 |ð,ðĨ = ∫ 2 ð+Π ð ðĖ 2 ðð ð ð ð+Π Ė ðĖ2 (ð + Π, ðĨ2 ) ð2 âđ = ððĨð ∫ ðð Ė ð ð ð2 (ð, ðĨ2 ) ð But ðĖ2 (ð, ðĨ2 ) = ðĨ2 ðū2 ð2 (ð) from first ratio and ðĖ2 (ð + Π, ðĨ2 ) = ð2 (ð) from equilibrium criterion. Ė 2 ð+Π ð Substituting them results in: ððĨð ∫ð ð ð ðð = Ė 2 ð+Π ð Taking natural logarithm both sides: ∫ð ð ð 1 ðĨ2 ðū2 ðð = − ln(ðĨ2 ðū2 ) Looking at the LHS: If ðĖ 2 is a weak function of P (common in liquids), it means that: Ė 2 ð+Π ð ∫ð âđ ð ð Ė 2 Πð ð ð ðð = Ė 2 ð ð ð (ð + Π − P) = = − ln(ðĨ2 ðū2 ) Ė 2 Πð ð ð ð ð âđ Π = − ðĖ ln(ðĨ2 ðū2 ) 2 If the solution is very dilute in the solute (species 1): ðĨ1 → 0, ðĨ2 → 1, ðĖ 2 ≈ ð2 and ðū2 ≈ 1 âđ ln(ðĨ2 ðū2 ) = ln(ðĨ2 ) = ln(1 − ðĨ1 ) Applying Maclaurin series and neglecting x12 and higher powers of x1: ln(1 − ðĨ1 ) = −ðĨ1 Therefore: ð· = ðð ðđðŧ ð―ð this is called the van’t Hoff equation. The van’t Hoff equation applies where species 1 is a non-electrolyte. If ð′ > ð + Π then: ðĖ2 (ð′ , ðĨ2 ) > ð2 (ð) This difference in fugacities is a driving force where solvent molecules are transferred from left to right in Figure 4. This process is called reverse osmosis, which is a pressure-driven process. Topic 5: Chemical reaction equilibria A chemical reaction is represented as: |ðĢ1 |ðī1 + |ðĢ2 |ðī2 + âŊ → |ðĢ3 |ðī3 + |ðĢ4 |ðī4 + âŊ Where |ðĢð | is the stoichiometric coefficient and ðīð is the chemical formula. ðĢð itself is called the stoichiometric number and is positive for products and negative for reactants. For an inert species (a species that does not take part in the reaction), it is zero. As the chemical reaction progresses, the reactants are consumed whilst products are formed. The changes in the number of moles (positive for products and negative for reactants) of each species are proportional according to the stoichiometry of the reaction. âđ âđ ððð = ðĢð ðð ðð1 ðð2 ðð3 ðð4 ððð = = = = = ðð ðĢ1 ðĢ2 ðĢ3 ðĢ4 ðĢð where ð is called the reaction coordinate which characterizes the extent of reaction or degree of completion ð ð Integrating both sides: ∫ð ð ððð = ðĢð ∫0 ðð âđ ðð = ðð0 + ðĢð ð ð0 Let ð = ∑ð ðð , ð0 = ∑ð ðð0 , ðĢ = ∑ð ðĢð : Therefore: ðð = ðð ð = âđ ð = ð0 + ðĢð ððð +ðð ðš ðð +ððš Equilibrium of chemical reactions The equilibrium criterion: (ððš ðĄ ) ð,ð = 0 Recall: ð(ððš) = (ðð)ðð − (ðð)ðð + ∑ð ðð ððð But ððð = ðĢð ðð, therefore ð(ððš) = (ðð)ðð − (ðð)ðð + ∑ð ðĢð ðð ðð Since ððš or ðš ðĄ is a state function, the RHS is an exact differential expression, meaning: ð(ððš) ððš ðĄ ∑ ðĢð ðð = ( ) = ( ) ðð ð,ð ðð ð,ð ð (6) ððš ðĄ At equilibrium: ( ðð ) ð,ð = ∑ð ðĢð ðð = 0 Recall: ðð = ðĪð (ð) + ð ð ðð ðĖð (29) For pure species i in its standard state (usually P = 1 bar or 1 atm) at the same temperature, Equation 27 can be written as: ðšð ð = ðĪð (ð) + ð ð ðð ðð ð (53) ðĖ Subtracting Equation 53 from Equation 29: ðð = ðšð ð + ð ð ln (ð ðð ) ð ðĖð âđ ∑ ðĢð [ðšð ð + ð ð ln ( ð ) ] = 0 ðð ð ðĢð ðĖð âđ ∑ ðĢð ðšð + ð ð ∑ ln ( ð ) = 0 ðð ð ð ð ðð ∑ð ðð ðŪð ð ðĖð âđ ðĨð§ ∏ ( ð ) = − ðđðŧ ðð ð ðĖ ðĢð But ∏ð (ð ðð ) = ðū which is the equilibrium constant at constant pressure and ∑ð ðĢð ðšð ð = âðš ð ð which is the standard Gibbs Free Energy change of reaction âðŪð âđ ðĨð§ ðē = − ðđðŧ Since pressure is fixed at standard state conditions, âðš ð is a function of T only. ðĖ Fugacity ratios ð ðð give the connection between the equilibrium state of interest and standard state ð of each species. The general formula for standard property changes of reaction is: âðð = ∑ð ðĢð ðð ð which are also functions of T only ð(ðš ⁄ð ð ) From Gibbs Free Energy as a generating function: ( ðð ðŧ ) = − ð ð 2 ð Applying this: ðŧð ð = −ð ð 2 ð(ðšð ð ⁄ð ð) ðð ∑ð ðĢð ðŧð ð = −ð ð 2 Multiplying by ðĢð and summing over all i: âđ ð(âðš ð ⁄ð ð ) ðð âðŧ ð = − ð ð 2 âđ ð ðĨð§ ðē ð ðŧ = ð(∑ð ðĢð ðšð ð ⁄ð ð ) ðð âðŊð ðđðŧð This shows that for an exothermic reaction (âðŧ ð is negative), K decreases as T is increased. The converse is true for an endothermic reaction (âðŧ ð is positive) If âðŧ ð is assumed to be independent of temperature, integrating both sides yields: ðū ∫ðū′ ð ln ðū = âðŧ ð ð ð 1 ðū ∫ð ′ ð 2 ðð âđ ln (ðū′ ) = − 1 âðŧ ð 1 ð 1 (ð − ð ′ ) Therefore a plot of ln ðū vs ð gives a straight line as shown in Figure 5. Figure 5 Relation of equilibrium constant to composition: Gas-phase reactions ðĖ Recall that: ðū = ∏ð (ð ðð ) ðĢð ð The standard state of a gas is the ideal gas state of the pure gas at the standard-state pressure (ðð ) of 1 bar or 1 atm. But ðð ðð = ð, therefore this implies that ðð ð = ðð ðĢð ðĖð âđ ðū = ∏ ( ð) ð ð ðĢð Ė ððĶð But ðĖð = ðĖð ðĶð ð. This implies that: ðū = ∏ð ( ðððð ) ð ðĢð ð ðĢ But ∏ð (ðð ) = (ðð ) where ðĢ = ∑ð ðĢð ð ðĢ ðĢð âđ ðū = ( ð ) ∏(ðĖð ðĶð ) ð ð ð −ðĢ ðĢð Ė âđ ∏(ðð ðĶð ) = ðū ( ð ) ð ð If the equilibrium mixture is assumed to be an ideal solution of gases: ðĖð âđ ∏(ðð ðð )ðð ð âđ ∏(ðð ð = ðð (42) ð· −ð = ðē ( ð) ð· If the equilibrium mixture is assumed to be an ideal gas: ðĖð )ðð ðð ðð = 1 ð· −ð = ðē ( ð) ð· This shows that whilst K is not a function of P, ∏ð (ðĶð )ðĢð is a function of P. - An increase in K at constant pressure increases the product ∏ð(ðĶð )ðĢð which implies a shift of equilibrium to the right (more products are formed) - If ðĢ is negative, an increase in P (equilibrium mixture pressure) at constant temperature increases the product ∏ð (ðĶð )ðĢð but has no effect on K. If ðĢ = 0, P has no effect on the product ∏ð (ðĶð )ðĢð - Addition of inert gases does not affect K (since ðĢðððððĄ = 0). But it does affect the equilibrium composition. Since ðĶð = ðð ð , adding inerts increase n. ðð ðĢð ð −ðĢ âđ ∏ ( ) = ðū ( ð) ð ð ð 1 ðĢð 1 ðĢ ð −ðĢ But ∏ð (ð) = (ð) . This therefore means: ∏ð (ðð )ðĢð = ðū (ððð ) • If ðĢ is negative, an increase in n at constant temperature and pressure decreases the product ∏ð(ðð )ðĢð which implies a shift of equilibrium to the left (more reactants are formed). If ðĢ = 0, addition of inerts has no effect on the product ∏ð(ðð )ðĢð - The factor by which the reaction is multiplied is the power by which K is raised. - The larger the K, the more complete the forward reaction - If the chemical reaction is reversed, new K is the reciprocal of the old K - If the overall reaction is a sum of different individual reactions, K for the overall reaction is equal to the product of the K of each individual reaction