Sans titre

1 downloads 0 Views 3MB Size Report
Abstract. Supercritical fluids (SFC) have unique advantages over gases and liquids. Many processes have been developed all around the world to benefit of ...
STATE OF THE ART OF EQUATIONS OF STATE Christophe Coquelet and Dominique Richon Laboratoire Thermodynamique et Equilibres entre Phases Centre Energétique et Procédés, Ecole Nationale supérieure des Mines de Paris CNRS FRE 2861 35, rue Saint Honoré 77305 Fontainebleau [email protected] fax: 33164694968 [email protected] fax: 33164694968 Abstract Supercritical fluids (SFC) have unique advantages over gases and liquids. Many processes have been developed all around the world to benefit of possible SFC performances near critical conditions. Process optimization must rely on reliable models and unfortunately although thermodynamic properties are conveniently represented or even predicted in large ranges of conditions, the supercritical region caused drastic troubles. A quick review of modelling is done herein showing the various drawbacks pointed out regarding conventional data treatments. Finally, optimistic ways are examined to solve difficulties approaching critical points. INTRODUCTION Supercritical fluids (SF) exhibit very interesting properties, giving them great potential for industrial use. As the supercritical state is an intermediate state between a liquid and a gas, some of physical properties lie in between those of these two states, while the others show unique behavior around the critical point. Supercritical fluids are employed in many industrial applications, such as in the food and pharmaceutical industries, in the field biotechnologies, etc and for development of new materials. Typical examples of applications are: - Extraction of chemicals such as caffeine, aromas, nicotine, active principles for drug making, pesticides from food, … - Treatment of heavy hydrocarbons, polymer fractionation and extraction of monomers, - Regeneration of filters, absorbents and catalysts, - Separation, purification (isomer separation), - Use of supercritical CO as a green solvent or refrigerant, as a substitute of the forbidden CFCS, sterilization of food product with high pressure CO , … 2

2

- Chemical reactions in supercritical conditions - Supercritical fluid chromatography,etc. Supercritical fluid extraction (SFE) is definitely a more and more used technology. The basic steps of SFE processes are the following:

1

1- The material (solid or liquid) containing the substance (or substances) to be extracted is put in contact with a high-pressure supercritical fluid, and the substance (substances) is (are) partially dissolved by the supercritical phase 2-­‐‑   The   substance   is   recovered   after   changing   the   conditions   of   temperature   and   pressure.   Two   phases   are   created:   the   solvent   which   behaves   as   a   gas,   and   the   substance,  which  remains  in  the  dense  phase  that  is  poor  in  solvent.   Supercritical   fluid   extraction   has   several   advantages:   many   substances   are   more   soluble  in  high-­‐‑pressure  supercritical  fluids  than  in  liquids  at  ambient  conditions.  As   diffusion   coefficients   are   higher   in   supercritical   fluids   than   in   liquids   so   are   mass   transfer  rates.  The  viscosity  of  SF  is  also  lower  than  that  of  liquids.  The  most  common   compound  used  as  solvent  in  SFE  is  carbon  dioxide  (CO2)  which  is  a  very  interesting   compound   for   several   reasons:     it   is   cheap,   non-­‐‑toxic,   and   its   critical   temperature   (304.21  K)  is  close  to  ambient  temperature,  allowing  reduced  heating  costs.  Another   advantage   and   not   the   least,   CO2   is,   at   high   pressures,   a   good   sterilizer   for   food   products.     Very accurate thermodynamic models are required for both new designs and improvments of SFE separation processes. These thermodynamic models must be able to accurately predict both the densities and the phase equilibrium properties of systems involving supercritical and near-critical fluids. This issue is particularly important, as solubility of substances and selectivity in supercritical fluids depends greatly on the density of the system, which changes considerably even for very small pressure modifications. The use of high pressures in SFE processes represents unfortunately the largest part of the operating costs. Equations of State (EoS) allow developing the most convenient thermodynamic models to represent phase equilibrium properties when dealing with high pressures. The most widespread EoS in industry are probably the cubic equations of state, such as the Peng-Robinson and Patel-Teja EoS etc…, because these equations have a very simple analytical form, and they provide very accurate predictions of vapor-liquid equilibrium for mixtures of non polar molecules. Their great disadvantage is to provide inaccurate density representations, in the liquid and in the near critical regions. For pure compounds, it is possible to introduce temperature dependent volume translations in cubic EoS. For mixtures the main problem of most of proposed volume translations is their inadequacy to improve PVT properties on the whole range of pressures (improvements can be obtained in a given limited pressure range while deviation are higher in another limited pressure range…) For very non-ideal mixtures, volume translations are mostly unreliable and do not enable better representation of phase equilibria. During the last twenty years, several theoretical equations of state such as the SAFT EoS and its various versions (SAFT-VR, PC-SAFT, soft SAFT …) have been developed thanks to the progress in both statistical mechanics and computer simulation. Although these equations of state are more complex than cubic EoS, they are more and more used thanks to the increase of the computer speed; some of them have been implemented in process simulation softwares (for example, PC-SAFT in ASPEN + software from ASPEN Tech). The main advantages of these equations are their capabilities to treat very complex mixtures involving a large variety of compounds: associating fluids (molecules with hydrogen bonds), electrolytes, polymers, colloids… Such equations of state are usually more accurate than cubic EoS for the predictions of liquid densities, and require less binary parameters to represent phase equilibrium data. However, these EoS are unable to predict the thermodynamic properties in near critical regions, hence they have the same 2

drawback as cubic EoS. This is currently common problem to all classical EoS that do not take into account the large density fluctuations occurring close to the critical point. I – THERMODYNAMIC BACKGROUND I.1. Definition of the phase diagrams I.1.1. Pure compound The phase diagram of a pure compound is clearly defined by two points: the critical point and the triple point. At the triple point, solid, liquid and vapour phase are together coexisting. The critical point can be defined as a limit for the pure component vapour pressure. For temperature and pressure above the critical condition, there is no possibility to have vapour-liquid equilibrium. We can also consider the supercritical state as a “stable state” with no phase separation. For certain temperature and pressure conditions, phase separation appears leading to two phases in equilibrium (liquid and vapour). The critical point can be considered as a limit of stability of the supercritical phase. The first equation of state, which can describe the behaviour of a pure fluid, was developed by van der Waals [1]. Two types of interactions (repulsive and attractive interactions) are considered in his equation.

a ⎞ ⎛ ⎜ P + 2 ⎟(v − b ) = RT v ⎠ ⎝

(1)

The stability conditions can be defined with the following equations:

⎛ ∂ 2 P ⎞ T ⎛ ∂P ⎞ >0 ⎜ ⎟ < 0, ⎜⎜ 2 ⎟⎟ < 0 and cv ⎝ ∂v ⎠ T ⎝ ∂v ⎠ T At the limit, these two previous conditions are equal to zero. ⎛ ∂ 2 P ⎞ T ⎛ ∂P ⎞ =0 ⎜ ⎟ = 0, ⎜⎜ 2 ⎟⎟ = 0 and cv ⎝ ∂v ⎠ T ⎝ ∂v ⎠ T It can be seen that at the critical point, the isochoric heat capacity has infinite value.

3

I.1.2. Mixture To illustrate this part, we just consider a binary mixture. Phase diagram concerning the mixture depend of the behaviour of the species. Van Konynenburg and Scott [2] have classified the mixture into VI types. Figure 1 presents the different types of phase diagrams.

Figure 1: Six types of phase behaviour in binary fluid systems. C: Critical point, L: Liquid, V: Vapor. UCEP : Upper critical end point, LCEP : Lower critical end point. Dashed curve are critical lines and hatching marks heterogeneous regions. Extract from J.M. Prausnitz, R. N. Lichtenthaler, E.G. de Azevedo, Molecular thermodynamics of fluid phase equilibria, 1999, Prentice Hall PTR, Upper Saddle river, New Jersey, USA. Type I phase behaviour It is the simplest type of phase diagram. The mixture critical point line starts at the first pure component critical point and finish at the second pure component critical point. It concerns mixtures where the two components are chemically similar or their critical properties are comparable. We can cite systems with CO and light hydrocarbons, systems with refrigerants HFC (figure 2), and benzene + toluene binary systems. 2

4

P /MPa

6.5 6.0 5.5 5.0 4.5 4.0 3.5 3.0 2.5 2.0 0.0

0.2

0.4

0.6

0.8

1.0

x1 , y1 Figure 2: Critical point line of R32 (1) + propane (2) (Coquelet et al. [3]) binary system calculated with cubic equation of state. Symbols: experimental VLE data at 343.26 K. Type II phase behaviour It is similar to type I but at low temperatures, the two components are not miscible in the liquid mixtures. Consequently a liquid - liquid equilibrium appears. The mixture critical point line for liquid - liquid equilibrium starts from UCEP (upper critical end point). At the UCEP, the two liquid phases merge into one liquid phase. We can cite systems with hydrocarbons and fluorinated fluorocarbons. Type III phase behaviour It concerns mixtures with very large immiscibility gaps. We can cite for example systems with water like hydrocarbons + water systems. A liquid – liquid - vapour curve appear and a first mixture critical point line starts from pure component 1 critical point and ends at the UCEP. The second one starts from the infinite (P-> ∞) and ends at the pure component 2 critical point, generally the solvent i.e. water. The slope of this second curve can be positive, negative or positive and negative. Concerning the positive curve, we have two phases at temperature larger than the critical temperature of the pure component two. Some example: Positive slope: helium + water binary system Negative slope: methane + toluene binary system Positive and negative slope: nitrogen + ammonia or ethane + methanol binary systems. Figures 3a, 3b and 3c present some example from the paper of Heidemann and Khalil [4]. It concerns CO + n-octane (a), CO + n-hexadecane (b) and CH + H S (c). 2

5

2

4

2

Figure 3a: Critical lines for CO + n-octane binary system. 2

Figure 3b: Critical lines for CO + n-hexadecane binary system. 2

Figure 3c: Critical lines for methane + H S binary system. 2

Type IV phase behaviour It is similar to type V. the vapour liquid critical point line starts at the critical point of component 2 and ends at the LCEP (lower critical end point). Vapour-liquidliquid equilibrium (VLLE) exists and is present in two parts. Ethane + n-propanol and CO + nitrobenzene binary systems behave like this. 2

Type V phase behaviour

6

It is a modification of type III phase diagram. There are two vapour-liquid critical point lines. One goes from the pure component critical point 1 and ends at the UCEP. The other one starts at the pure component critical point 2 and ends at the LCEP. Contrary to type IV, below the LCEP the liquids are completely miscible. Ethylene + methanol binary system is classified as a type V system (see figure 4).

Figure 4: Vapour-liquid equilibrium pressure – composition diagram of ethylene + methanol binary system. Extract from J. M. Prausnitz, R. N. Lichtenthaler, E. G. de Azevedo, Molecular thermodynamics of fluid phase equilibria, 1999, Prentice Hall PTR, Upper Saddle river, New Jersey, USA. Type VI phase behaviour There are two critical point curves. The first one is similar to one presented with the type I diagram: a connection between the two pure component critical points. The second one connects the LCEP and thee UCEP. Between these two points, there are VLLE. The system “water + 2-butanol” is one typical example. I.2. Thermodynamic approach There are mainly two different approaches to model phase equilibrium properties. The two approaches are based on the fact that at thermodynamic equilibrium, fugacity values are equal in all phases, at isothermal conditions. For vapor-liquid equilibrium we have :

f iV (T , P, y ) = f i L (T , P, x )

7

(2)

I.2.1. gamma-Phi approach This approach is based on the use of an activity model for the liquid phase and an equation of state for the vapor phase, the γ – Φ approach. The equilibrium equation can be written:

Φ Vi (T , P, y )y i P = γ iL (T , x )xi f i 0 L (T , P )

(3)

because

f iV (T , P, x ) = γ iL (T , x )xi f i 0 L (T , P ) The vapor fugacity is calculated as follows: f iV (T , P, y ) = Φ Vi (T , P, y )y i P

(4)

The fugacity coefficient in the vapor phase is calculated using an equation of state or a virial equation. The fugacity of the pure compound, i, in the liquid phase can be written using the vapor pressure as:

(

⎛ v L P − Pi sat f i 0 L (T , P ) = Pi sat Φ i0 T , Pi sat exp⎜⎜ i RT ⎝

(

)

)⎞⎟ ⎟ ⎠

(5)

The exponential factor is known as the Poynting factor and viL is the liquid molar volume, at saturation, of the component i. But, the utilization of this approach is limited to systems studied at low pressures. And it is obvious that close to the mixture critical point, the calculation will not converge correctly as two different equations are used. The best way is to use similar equations for both vapour and liquid phase. This solution will be also used to describe supercritical states. 1.2.1. Activity coefficient models

γ iL (T , P, x ) is known as the activity coefficient. The activity is defined by the ratio between fugacity in the mixture and in the reference state. Activity coefficient is fi f introduced to characterise the non ideality of the mixture, γ i = ideal = *i . fi f i xi Several models can describe the behaviour of liquid mixtures. We van cite the empirical model based on Redlich Kister [5] equation. In general, models with theoretical origin are preferred for their better reliability when interpolating or extrapolating as necessary for process design. The most important one concerns the concept of local composition introduced by Wilson [6]. Models like NRTL [7], UNIQUAC [8] and predictive UNIFAC [9] model are based on this local composition concept. 1.2.2. Virial type equations These equations are used to accurately represent experimental properties of pure compounds. The extension from pure compounds to fluid mixtures is problematic, as 8

in theory a mixing rule is needed for each parameter. The virial equation of state is one of the most known of the large number of equations, which have been proposed. This equation is a development of the compressibility factor in series expanded in powers of the molar density with density-independent coefficients, B and C:

Z=

Pv B(T ) C (T ) = 1+ + 2 + .... RT v v

(6)

The density-independent coefficients are known as virial coefficients. (B called the second virial coefficient, C the third). Thus, the second virial coefficient depends on the interaction between two molecules, the third between three molecules, etc… Correlation based corresponding state principle are used to determine the different parameters (ref [10]). I.2.2. Phi-Phi approach This approach is more powerful as it can be used for low pressures and also for high pressures. In this approach the same model is used for the liquid and the vapour phase of either pure compounds or mixtures. Consequently, Eq. 2 becomes:

Φ Vi (T , P, y ) = Φ iL (T , P, x )

(7)

For the determination of fugacity coefficient, phi, we have to consider the following relation:

RTd ln( f i (T , P, z )) = vi dP

(8)

leading to : v ⎡ ⎤ RT ⎛ ∂P ⎞ ⎥dnv − RT ln(Z ) ⎟⎟ RT ln(Φ i ) = ∫ ⎢ − ⎜⎜ nv ∂ n ⎢ ⎝ i ⎠ T ,nv,n j ≠ i ⎥⎦ ∞ ⎣

(9)

It is obvious that a relation between all the variables T, P and v must be used. Equations of state correspond to the main way chosen in industry to describe and calculate fluids properties of mixtures.

9

II. Description of models II.1. The van der Waals type model Van der Waals [1] develops the first equation of state which can describe vapor, liquid and supercritical states. He took into account the different interactions between the molecules. Several authors have proposed modifications of this equation and extended them to the representation of more sophisticated fluids (pure component or mixtures). II.1.1. Description of the equation Cubic equations of state can be described using this generalized form:

P=

RT a(T ) − 2 v − b v + ubv + wb 2

(10)

With parameters: a, b, u and w. a is the attractive term. It takes into account the attractive forces between the molecules. It is temperature dependant ( a(T ) = aCα (T ) ,α is called alpha function). b is designed as the molar co-volume. It takes into account the repulsive interactions. Table 1 gives the values of u and w for the most important cubic EoS. At the critical point, we have:

(v − vC )3 = v 3 − 3vC v 2 + 3vC2 v − vC3 = 0

(11)

Using the cubic equation of state at the critical point leads to equation (12).

⎡ RT ⎤ ⎡ RT a ⎤ ⎡ RT ab ⎤ v 3 − ⎢ C − (u − 1)b⎥v 2 − ⎢ C ub − (w − u )b 2 − ⎥v − ⎢ C wb 2 + wb 3 + ⎥ = 0 PC ⎦ ⎣ PC PC ⎦ ⎣ PC ⎦ ⎣ PC Where a = Ωa

(12)

R 2TC2 RTC , b = Ωb PC PC

Using values of u et w presented in Table 1, with c = Ω c

RTC the following system of PC

equations must be solved.

uΩb = 1 + Ωb − 3Z C

(13)

Ω 3b + [(1 − 3Z C ) + (u + w)]Ω b2 + 3Z C2 Ω b − Z C3 = 0

(14)

Ω a = 1 − 3Z C (1 − Z C ) + 3(1 − 2Z C )Ω b + [2 − (u + w)]Ω b2   Equation of State u

(15)

10

w

Van der Waals Redlich and Kwong [11] Peng and Robinson [12] Patel and Teja [13]

0 1 2 1+ c

Harmens and Knapp [14]

c

b

0 0 -1 −c

b -(c-1)

Table 1 – u and w parameters of the main cubic equations of state. c is a third parameter.   II.1.2. The alpha function Alpha functions are used to have the best representation of the pure component vapor pressure. They are empirical equation but their mathematics expressions are very important. α   must   tend   towards   zero   at   high   temperatures   (thermal   agitation   overrides   the   attractive   interactions),   α   must   be   equal   to   1   for   TR   =   1   (critical   point)   and   must   tend   towards   infinite   values   when   the   temperature   tends   towards   0   (the   motionless  molecules  more  strongly  attract  each  other).  All  the  alpha  functions  must   take  a  similar  form.  In  1972,  Soave  [15]  introduced  an  alpha  function  (see  equations   16  and  17)  in  order  to  improve  calculation  of  the  vapor  pressures  (and  of  liquid  and   vapor  volumes).  TR  is  the  reduced  temperature  and  ω  the  acentric  factor:    

α (T ) = [1 + m(1 − TR1 2 )] m = 0.480 + 1.574ω − 0.175ω 2 2

(16) (17)

Other   generalized   alpha   functions   were   developed   according   to   the   cubic   type   of   equation.  They  were  established  starting  from  experimental  data  of  pure  substances   between   the   boiling   point   under   1   atm   and   the   critical   point.   The   generalization   of   the  coefficient  m  according  to  the  acentric  factor  involves:   For RKS : m = 0.47830 + 1.6337ω − 0.3170ω 2 + 0.760ω 3

(18)

For PR : m = 0.374640 + 1.542260ω − 0.26992ω 2

(19)

Mathias  and  Copeman  [16]  proposed  a  function  with  three  parameters  (c0,  c1  and  c2)   adjustable  on  experimental  data:   2 3 1 1 1 ⎡ ⎤ α (T ) = ⎢1 + c1 ⎛⎜1 − TR 2 ⎞⎟ + c2 ⎛⎜1 − TR 2 ⎞⎟ + c3 ⎛⎜1 − TR 2 ⎞⎟ ⎥ ⎝ ⎠ ⎝ ⎠ ⎝ ⎠ ⎦ ⎣

2

(20)

When  TR  >1,  one  uses  only  the  first  term  of  the  alpha  function.   1

α (T ) = ⎡⎢1 + c1 ⎛⎜1 − TR 2 ⎞⎟⎤⎥ ⎣

11

⎝

⎠⎦

2

(21)

It   should   be   noticed   that   this   last   expression   makes   it   possible   to   have   a   representation  much  more  precise  of  vapor  pressures  than  with  the  alpha  function  of   Soave.   Nevertheless   it   is   possible   to   choose   other   generalized   alpha   functions   in   order   to   obtain   a   better   restitution   of   experimental   vapor   pressures   and   also   for   a   better  representation  in  supercritical  conditions  (use  of  temperature  dependent  alpha   functions).   Trebble   and   Bishnoi   (1987)   [17]   proposed   an   expression   different   from   that  proposed  by  Soave:     (22) α (T ) = exp(m(1 − TR )) Lastly,   Twu   et   al.   [18]   developed   an   alpha   function   based   on   a   different   approach.   They  considered  a  linear  function  with  respect  to  the  acentric  factor,  ie:  

α (T ) = α (0 ) (T ) + ω (α (1) (T ) − α (0 ) (T ))

(23)

with

α (T ) = TRN (M −1) [exp( L(1 − TRNM ))]

(24)

In  the  same  style,  Coquelet  et  al.  [19]  proposed  a  generalization  of  the  Mathias  and   Copeman  alpha  function  (equations  25  and  26)  for  the  cubic  equations  of  Redlich  and   Kwong  and  of  Peng  and  Robinson,  based  on  the  use  of  23  reference  compounds.   For the Redlich - Kwong equation: c1 = −0.1094ω 2 + 1.6054ω + 0.5178

c 2 = −0.4291ω + 0.3279

(25)

c 3 = 1.3506ω + 0.4866 For the Peng - Robinson equation:

c1 = −0.1094ω 2 + 1.6054ω + 0.5178 c2 = −0.4291ω + 0.3279

(26)

c3 = 1.3506ω + 0.4866 Coquelet   et   al.   [19]   also   proposed   another   alpha   function   (see   equations   27   and   28)   which  combines  the  advantages  of  the  Mathias  and  Copeman  and  of  the  Trebble  and   Bishnoi   alpha   functions,   and   in   particular   for   temperatures   higher   than   the   concerned  critical  temperature.   2 3 1 1 ⎡ ⎤ α (T ) = exp[c1 (1 − TR )]⎢1 + c2 ⎛⎜1 − TR 2 ⎞⎟ + c3 ⎛⎜1 − TR 2 ⎞⎟ ⎥ ⎝ ⎠ ⎝ ⎠ ⎦ ⎣

For the Peng - Robinson equation:

12

2

(27)

c1 = 0.1441ω 2 + 1.3838ω + 0.387 c2 = −2.5214ω 2 + 0.6939ω + 0.0325

(28)

c3 = 0.6225ω + 0.2236 II.1.2. Presentation of the mixing rules The  mixing  rules  must  be  able  to  take  into  account  the  ideal  and  nonideal  characters   of   the   solutions.   With   the   cubic   equations   containing   two   parameters,   the   objective   consists  in  recalculating  a  and  b  parameters  considering  the  mutual  influence  of  the   various  compounds.  The  first  set  of  mixing  rules  is  that  initially  proposed  by  van  der   Waals,   it   corresponds   to   the   so   called   "ʺclassical   mixing   rules"ʺ.   On   the   basis   of   an   equation  of  state,  carrying  out  a  development  of  the  volume  virial  development  and   by   applying   statistical   thermodynamics   one   leads   to   the   following   classical   mixing   rules:     (29) a = ∑∑ xi x j aij i

j

and b = ∑ xi bi i

(

where aij = ai a j 1 − kij

)

kij  is  called  binary  interaction  parameter.  This  parameter  must  take  into  account  the   fact   that   the   attractive   interactions   between   compounds   i   and   j   are   different   from   those  between  i  and  i  and  j  and  j.   Many   more   advanced   mixing   rules   have   been   developed.   Their   authors   took   into   account   models   based   on   the   calculation   of   the   activity   coefficient   (excess   free   enthalpy)  and  models  involving  equations  of  state.  Indeed  the  first  are  adapted  to  the   treatment   of   the   polar   and/or   non   polar   bodies   at   low   pressures   and   are   in   general   not   dependent   of   the   pressure,   while   the   equations   of   state   give   satisfactory   results   only  for  non-­‐‑polar  bodies  but  without  pressure  limitation.  

gγE (T, P → ∞) = g EEoS (T, P → ∞) and v = b when P→∞. The first mixing rule (equations 30 and 31) of this type was presented by Huron and Vidal [20] in 1979.

⎛ ⎛ a a = b⎜⎜ ∑ xi ⎜⎜ i ⎝ i ⎝ bi b = ∑ xi bi i

13

⎞ ⎞ ⎟⎟ + g PE=∞ × C ⎟ ⎟ ⎠ ⎠

(30) (31)

1 2 2 and C = − ln(2) ln 3 + 2 2 Robinson EoS respectively. With C = −

(

)

for the Redlich-Kwong and the Peng-

Wong   and   Sandler   [21]   (1992)   preserved   the   classical   mixing   rules   of   mixtures   obtained  through  virial  development.  However,  as  for  the  Huron-­‐‑Vidal  mixing  rule,   equality   of   the   free   excess   enthalpies,   according   to   the   approaches   by   equation   of   state  and  activity  coefficient  makes  it  possible  to  obtain  another  relation  between  the   a   parameter   of   attraction   and   molar   co-­‐‑volume   b.   From   there,   one   can   also   write   starting  from  the  Helmholtz  molar  free  energy,  A:   E (T , P → ∞, x) = AγE (T , P → ∞, x) AEoS

(32)

On   the   other   hand,   the   authors   considered   that   the   free   energy   is   less   pressure   dependent  than  the  free  enthalpy.  Thus,  they  wrote:  

AγE (T , P → ∞, x) = AγE (T , P = 1 bar, x ) Indeed,  the  fundamental  relation  of  thermodynamics  is  written:  

g E (T , P, x) = A E (T , P, x ) + Pv E They   considered   that   at   low   pressures   the   PvE   product   is   negligible   with   respect   to   value  of  AE.  Consequently,  they  obtained:   E (T , P → ∞, x) = AγE (T , P → ∞, x) = AγE (T , P = 1 Bar, x) = gγE (T , P = 1 bar, x) AEoS Thus,   after   expressing   the   free   energy   of   mixture,   we   obtain   the   Wong-­‐‑Sandler   mixing  rule.  

i

b=

b−

⎛

a ⎞

∑∑ x x ⎜⎝ b − RT ⎟⎠ i

j

j

ij

a ⎛ ⎞ ⎜ ∑ xi i ⎟ E bi g γ (T , P, x ) ⎟ i ⎜ 1− + ⎜ RT ⎟ CRT ⎜ ⎟ ⎝ ⎠

(33)

a a ⎞ a ⎞ 1 ⎡⎛ a ⎞ ⎛ a ⎞ ⎤ ⎛ ⎛ = ∑∑ xi x j ⎜ b − ⎟ with ⎜ b − ⎟ = ⎢⎜ b − ⎟ + ⎜ b − ⎟ ⎥(1 − kij ) (34) RT RT ⎠ ij RT ⎠ij 2 ⎢⎣⎝ RT ⎠i ⎝ RT ⎠ j ⎥⎦ ⎝ ⎝ i j

Michelsen [22] (1990) used the concept of the Huron-Vidal mixing rule, but he chose to calculate the a parameter by determining the free excess enthalpy at null pressure. For that, he applied it to the RKS cubic equation of state. Thus, it determined the molar excess free enthalpy of in the following form:

14

gE ⎛ b ⎞ = ∑ xi Ln⎜ i ⎟ + q(α ) − ∑ xi q(α i ) RT ⎝ b ⎠ i i With α =

(35)

a bRT

We considered q(α) as q(α ) = q0 + q1α The corresponding modification leads to MHV1 mixing rules:

b = ∑ xi bi i

With

⎡ a RT a = b ⎢∑ xi i − bi q1 ⎣⎢ i

E ⎛ bi ⎞ gγ ⎤ ∑i xi Ln⎜⎝ b ⎟⎠ + q ⎥⎥ 1 ⎦

(36)

The   numerical   value   of   q1   is   -­‐‑0.593   for   RKS   EoS   and   -­‐‑0.53   for   the   PR   EoS.   Another   version   of   MHV1   was   proposed   by   Dahl   and   Michelsen   [23]   (1990),   it   consists   in   choosing  a  second  order  q  (α)  function.     (37) q(α ) = q0 + q1α + q2α 2 This leads to the second order equation which is the MHV2 mixing rule. (equation 38).

⎛ ⎞ ⎛ 2 GE ⎛ b ⎞ 2 ⎞ q1 ⎜ α − ∑ xiα i ⎟ + q2 ⎜ α − ∑ xiα i ⎟ = − ∑ xi Ln⎜ i ⎟ ⎝ b ⎠ i i i ⎝ ⎠ ⎝ ⎠ RT

(38)

a  while  for  co-­‐‑volume  we  still  have   bRT b = ∑ xi bi .   The   numerical   values   of   q1   and   q2   are   respectively:   -­‐‑0.478   and   -­‐‑0.0047   for  

Its  resolution  makes  it  possible  to  obtain   α = i

RKS  EoS  and  -­‐‑0.4347  and  -­‐‑0.003654  for  the  PR  EoS.       In   1991,   Holderbaum   and   Gmehling   [24]   developed   a   mixing   rule   by   keeping   the   principle  of  MHV1  rule,  but  by  taking  the  atmospheric  pressure  as  reference  pressure   (Predictive   Soave   Redlich   and   Kwong,   PSRK).   The   modification   that   they   brought   relates   to   the   calculation   of   the   excess   free   enthalpy.   They   use   UNIFAC   model   to   calculate   it.   On   the   other   hand,   the   change   of   reference   pressure   involves   a   light   modification   of   the   MHV1   q1   parameter   (-­‐‑0.64663).   The   values   of   the   UNIFAC   parameters  are  different  and  directly  adjusted  on  experimental  data.  This  is  why,  we   can  speak  about  the  PSRK  model  in  the  place  of  the  PSRK  mixing  rule.   II.1.3. one example

15

With cubic equations of state it is possible to try to represent vapor-liquid equilibrium properties for any systems. Figure 5 presents phase envelops obtained with the RK equation of state using the MHV1 mixing rules and the NRTL model. This model appears as quite convenient in this case to represent data at temperatures where we still have the presence of an azeotrope. However at higher temperatures close to the mixture critical point (see T = 343.26 K), this model is really not able to determine correctly VLE data. 6 5

P /MPa

4 3 2 1 0 0.0

0.2

0.4

x1 , y1

0.6

0.8

1.0

Figure 5: R32 + propane binary system. Symbols: experimental data from Coquelet et al. [3], temperature from 278.10 K to 343.26 K. Lines: calculated with RK EoS involving the MHV1 mixing rules and NRTL G model. E

Figure 2, shown previously for its critical point line determination, concerns the same binary system but the model is changed: PR EoS with the Wong Sandler mixing rules instead of MHV1 mixing rules. II.2. The GC EoS The group contribution equation of state (GC EoS) is based on generalized van der Waals partition function Q. By definition, the free Helmholtz energy can be determined through equation (39).

A = −kT ln(Q)

(39)

⎛ E ⎞ Where, Q = ∑ exp⎜ − i ⎟ ⎝ kT ⎠ i The residual Helmholtz function (at constant volume and constant temperature) becomes fv

att

⎛ A R ⎞ ⎛ A R ⎞ ⎛ A R ⎞ ⎜⎜ ⎟⎟ ⎟⎟ ⎟⎟ = ⎜⎜ + ⎜⎜ ⎝ RT ⎠T ,V ,n ⎝ RT ⎠T ,V ,n ⎝ RT ⎠T ,V ,n

16

(40)

The repulsive term (free volume) corresponds to the expression given by Carnahan and Starling, equation for hard spheres). For mixtures, the free volume contribution is given by expression developed by Mansoori and Leland (1972) [25]. fv

⎛ A R ⎞ ⎛ λ3 ⎞ ⎜⎜ ⎟⎟ = 3(λ1λ2 / λ3 )(Y − 1) + ⎜ 2 2 ⎟ − Y + Y 2 − ln Y + n ln Y ⎝ λ3 ⎠ ⎝ RT ⎠T ,V ,n

(

)

(41)

With

Y = ⎛⎜1 − ⎝ And

πλ3

⎞ 6V ⎟⎠

−1

(42)

NC

λk = ∑ n j d kj

(43)

j

d is the hard sphere diameter for one mole of species j. j

The attractive part of the residual Helmoltz function is written as follow: att

⎛ A R ⎞ a ⎜⎜ ⎟⎟ =− V ⎝ RT ⎠T ,V ,n

(44)

For pure components, the parameter a has the following expression:

a=

z 2 gq 2

(45)

Where, g is the characteristic attractive energy parameter per segment and q is the surface segment area per mole as defined by the UNIFAC method or the number of surface segments assigned to molecule. z is the coordination number (set equal to 10). For mixtures, a two-fluid model is considered with local surface fractions (not local mole fractions). The following expression is similar to the NRTL model for the excess Helmholtz function.

R

⎛ A ⎜⎜ ⎝ RT

att

⎞ ⎟⎟ ⎠T ,V ,n

⎛ g kj q~τ kj θ k ⎜⎜ ⎛ z ⎞ NC NG i NG ⎝ RTV = −⎜ ⎟∑ ni ∑ν j q j ∑ NG ⎝ 2 ⎠ i j k θτ

∑ l

With

17

l lj

⎞ ⎟⎟ ⎠

(46)

q

θ j = ~j q

NC

∑ν

i j

qj

i

NC

NG

i

j

q~ = ∑ ni ∑ν ij q j

(47)

Δg ji = g ji − g ii

ν ij is the number of groups of type j in molecule i, θ j is the surface fraction of group j and q j is the number of surface segments assigned to group j.

⎛ α ji Δg ji q~ ⎞ ⎟⎟ τ ji = exp⎜⎜ ⎝ RTV ⎠

(48)

Skjold-Jorgensen [26] gives more details in his paper. More sophisticated equations of state where developed recently. They take into account associating interactions between molecules. II.3. SAFT approach Several version of SAFT (Statistical Association fluid theory) are currently available. The first one was developed by Huang and Radosz [27] in 1990. It is based on theory of associating fluids developed by Wertheim [28-31]. It expresses the residual Helmholtz energy as a sum of segment chains, and association terms. The general expression for the Helmholtz energy is given by:

a = a seg + a chain + a asso

(49)

In this paper, we just consider the PC-SAFT (Perturbation Chain) equation based on perturbation theory of Barker and Anderson [32]. Gross and Sadowski [33] considered a second order theory for the determination of dispersive term. Each   molecule   is   regarded   as   a   chain   of   spherical   segments   not   being   identified   inevitably  with  an  atom.  The  interaction  potential,  u(r),  between  segments  of  a  chain   corresponds   to   a   modification   of   the   square   well   potential   (proposed   by   Chen   and   Kreglewski  [34]).   ⎧ ∞ r < (σ − s1 ) ⎪3ε (σ − s ) ≤ r < σ ⎪ 1 (50) u (r )⎨ − ε σ ≤ r < λσ ⎪ ⎪⎩ 0 r ≥ λσ Where  r  is  the  distance  between  two  segments,  σ    the  diameter  of  the  segment,  ε  the   depth  of  potential  well  and  λ  the  reduced  width  of  the  well  (s1=0.12σ).     The  compressibility  factor  is  the  sum  of  three  terms:  

Z =1 + Z seg + Z chain + Z asso

18

(51)

The  first  term  takes  into  account  repulsive  and  attractive  interactions.  The  repulsive   interactions   are   considered   using   a   model   of   hard   sphere.   Expressions   of   Boublick   [35]  and  Mansoori  et  al.  [36]  are  used  (Eq.  53).  

Z seg = mZ hc + Z disp

(52)

ξ3 3ξ 23 − ξ 3ξ 23 3ξ1ξ 2 + + 1 − ξ 3 ξ 0 (1 − ξ 3 )2 ξ 0 (1 − ξ 3 )3 π where ξ n = ρ ∑ xi mi d in 6 i Z hc =

(53)

d  is  the  diameter  of  collision  corresponding  to  a  segment  of  chain,  m  is  the  number  of   segments   per   chain.   The   chain   term   is   calculated   by   considering   the   following   relation:  

∂ ln g iihc (54) ∂ρ i g iihc is  the  radial  distribution  function  for  the  segments  of  compound  i  in  a  system  of   hard  spheres.   Z chain = −∑ xi (mi − 1)ρ

Concerning  the  dispersive  part,  the  theory  of  Barker  and  Anderson  is  used  regarding   the  chain  of  segments  of  hard  spheres  as  a  reference.  The  dispersive  interactions  are   only   one   perturbation   of   the   reference   state.   They   are   applied   to   molecules   comprising  several  chains  of  segments.  It  is  what  makes  it  possible  PC-­‐‑SAFT  model   to  be  able  to  be  used  for  the  study  of  polymers.     One  has:  

A A A disp = 1 + 2 NkT NkT NkT

(55)

where   A1   and   A2   are   the   first   and   second   order   contributions,   and   k   the   Boltzmann   constant.   A1   and   A2   are   obtained   through   the   following   relations   which   can   be   applied  to  any  potential  of  interaction.       ∞ A1 σ ⎞ ⎛ 2 ⎛ ε ⎞ 3 ~ = −2πρm ⎜ ⎟σ ∫ u (x )g hc ⎜ m; x ⎟ x 2 dx NkT d ⎠ ⎝ kT ⎠ 1 ⎝

⎡∞ ~ ⎤ σ ⎞ ⎛ ∂ ⎢ ∫ u (x )g hc ⎜ m; x ⎟ x 2 dx ⎥ 2 hc −1 d ⎠ ⎛ ⎝ A2 ∂Z ⎞ ⎛ ε ⎞ ⎦ ⎟⎟ m 2 ⎜ ⎟ σ 3 ⎣ 1 = −πρm⎜⎜1 + Z hc + ρ NkT ∂ρ ⎠ ∂ρ ⎝ kT ⎠ ⎝ with x =

19

r ~ u (x ) and , u (x ) =

σ

ε

(56)

(57)



σ ⎞ ⎛ I 1 = ∫ u~ (x )g hc ⎜ m; x ⎟ x 2 dx d ⎠ ⎝ 1 ∞ ∂ ⎡ ~ σ ⎞ 2 ⎤ hc ⎛ I2 = ⎢∫ u (x )g ⎜ m; x ⎟ x dx ⎥ ∂ρ ⎣ 1 d ⎠ ⎝ ⎦

(58) (59)

ghc  is  the  distribution  function  making  it  possible  to  know  the  number  of  molecules  in   a  certain  element  of  volume.     In   PC-­‐‑SAFT   theory,   I1   and   I2   can   be   estimated   through   weighted   sums.   The   associative   term   is   deduced   directly   from   the   expressions   of   Wertheim.   If   one   considers   two   spherical   segments   provided   with   a   site   of   association   A,   the   associative  connection  can  be  formed  only  when  the  distance  and  the  orientation  are   favourable.  Association  is  modelled  by  a  square  well  potential  of  interaction  centered   on   site   A.   Two   parameters   are   necessary:   the   parameter   ε

asso

which   corresponds   to  

asso

the   depth   of   the   well   and   the   parameter   κ which   characterizes   the   volume   of   association   (related   to   the   range   of   the   interaction).   These   parameters   make   it   possible  to  calculate  XA,  the  molar  fraction  of  molecules  not  associated  with  site  A.  

Z

asso

⎡ ⎡⎛ ∂X A j ⎢ = ∑ xi ∑ ρ j ∑ ⎢⎜⎜ ⎢ j i A j ⎢⎝ ∂ρ i ⎣ ⎣

⎤ ⎤ ⎞ 1 ⎡ ⎤ ⎟ ⎥ ⎥ ⎢⎣ X A j − 0.5⎥⎦ ⎥ ⎥ ⎟ ⎠T , ρk ≠i ⎦ ⎦

(60)

This model makes it possible to calculate for any type of associative molecules the thermodynamic properties. Binary interaction parameters, which allow the calculation of m 2εσ 3 and m 2ε 2σ 3 (see equations 61 and 62), can be adjusted directly on experimental data.

⎛ ε ij ⎞ m 2εσ 3 = ∑∑ xi x j mi m j ⎜⎜ ⎟⎟σ ij3 i j ⎝ kT ⎠

(61)

2

⎛ ε ij ⎞ m ε σ = ∑∑ xi x j mi m j ⎜⎜ ⎟⎟ σ ij3 i j ⎝ kT ⎠ 2

2

3

(62)

II.4. Limitation of these models Cubic or SAFT equations of state are limited to represent correctly the properties close to the critical point. The critical point is a particular point where the densities of the vapor and the liquid become identical. To describe correctly thermodynamic properties very close to critical points another theory must be used, based on statistical physics. III. The crossover EoS III.1. General presentation Both original cubic and non cubic EoS are unable to represent the densities and other properties over the full fluid range, because they do not satisfy the asymptotic laws 20

that are observed near the critical point. For instance, it is experimentally observed that the saturated liquid density ρ scales as ρ L − ρ C ∝ (TC − T )β where ρ and T are the critical density and temperature, and β one of the critical exponents. The experimental value of β is universal and is about 0.326, while any classical equation of state predicts that β = 0.5. This is why all cubic EoS underestimate the liquid densities. Classical EOS is unable to represent the divergence of the isochoric heat capacity and isothermal compressibility at the critical point. Other expressions can be written (for example): L

P − PC ρ − ρC ∝ PC ρC

C

C

δ

for the pressure along the coexistence VLE curve

T − TC TC The exponents δ, α and Δ are called critical exponents. It exist other critical exponents: β , γ , ν and η .   There are relations between them that were determined from the renormalization group theory: for instance,   α + 2β + γ = 2.   The experimental and classical values of these exponents are reported in table 2.

(

)

−α Δ CV ∝ kα 0 + kα1 (Δτ ) 1 + kα1 (Δτ ) 1 + ... for ρ ρ with Δτ = =

C

1

Critical exponents α β γ δ ν η Δ 1

Best estimate values 0.110 0.3255 1.239 4.800 0.630 0.033 0.510

Values from a classical EoS 0 0.5 1 3 0.5 0 0

Table 2: Critical exponents Asymptotic scaled equations of state have been developed from the renormalization group to take density fluctuations into account. Such EoS satisfies the universal scaling laws observed experimentally, but are inapplicable far away from the critical point. It is possible to combine such an asymptotic and non-analytical model to a classical EoS that is valid far away from the critical point, to obtain a global EoS applicable over the entire fluid range. Such approaches are called crossover equations of state (see the good review by Sengers in “Equations of State for Fluids and fluid mixtures”, Elsevier, Amsterdam, 2000]). In this paper, we will limit our discussion to pure fluids. III.2. Crossover Equation for pure fluids A simple method for incorporating scaling laws into an analytical equation of state has been developed by Kiselev [37]. This approach is based on renormalization group results in the critical region. First, we start with the Landau expansion for the pure fluid at the critical point gives the value of the thermodynamic potential.

φ (P, T ,η ) = φ0 (P, T ) + Aʹ′τη 2 + bη 4 − ηhV

21

(63)

With h, the intensity of the external field in Pa ( h = P − PC ), τ =

T v − 1 and η = − 1. TC vC

For constant T, P and η, the Landau expansion reduces to:

dφ = ΔA(τ ,η ) = Aʹ′τη 2 + bη 4

(64)

Considering the Helmholtz energy, the Taylor series expansion at the critical point, v where v r = = 1, gives (fourth order development): vC

⎛ 4 ⎞ ⎛ ∂A ⎞ ⎟⎟ (vr − 1) + ⎜⎜ ∂ A4 ⎟⎟ (vr − 1)4 + ... A(Tr , vr ) = A(Tr ,1) + ⎜⎜ ⎝ ∂vr ⎠Tr ,vr =1 ⎝ ∂vr ⎠Tr ,vr =1

(65)

The Landau expansion is inserted to the Taylor expansion:

⎛ ∂A ⎞ ⎟⎟ (vr − 1) + ΔA(τ ,η ) A(Tr , vr ) = A(Tr ,1) + ⎜⎜ ⎝ ∂vr ⎠ Tr ,vr =1

(66)

The development of this equation leads to

A (Tr , vr ) =

A(Tr , vr ) v = ΔA (τ ,η ) − P 0 +µ0 RT vC

(67)

III.3. Crossover function In order to satisfy the correct critical behavior in the vicinity of the critical point, i.e. satisfy the deviation of the thermodynamic properties like heat capacity with the critical exponents, the Landau expansion had to be corrected. The renormalization group theory is applied to renormalize the dimensionless variable τ and η (Nicoll et al. [38-40], Chen et al. [41-42]). It has been developed to deal with the effect of critical fluctuations. Variables are renormalized into:

τ = τ ×Y

−α 2Δ1

(68)

γ −2 β

η =η ×Y

4 Δ1

(69) Chen et al.[41-42] added a new term which is called the Kern term in order to provide correct critical isochoric specific heat capacity.

( )

ΔA(τ ,η ) = Aʹ′τ η 2 + bη 4 − K τ 2

( )

Where K τ

2

−α ⎞ 1 2 ⎛⎜ Δ1 = cτ Y − 1⎟ . ⎜ ⎟ 2 ⎝ ⎠

The crossover function Y can be written in the parametric form: 22

(70)

Δ1

⎡ q 2 ⎤ Y (q ) = ⎢ ⎥ ⎣ R(q )⎦ q is the measure of the distance from the critical point. Kiselev [37] uses the following crossover function. ⎡ q 2 ⎤ R(q ) = ⎢ + 1⎥ ⎣1 + q ⎦

(71)

2

(72)

It can be seen that close to the critical point (q=0), R(q) = 1 and Y (q ) = q 2Δ1 , far from the critical point, q → ∞ R(q ) = q 2 and Y (q ) = 1. The distance q is given by:

⎛ η + d1τ + d 2τ 2 ⎞ ⎟⎟Y q = + b ⎜⎜ β G G ⎝ ⎠ 2

τ

2 lm

1− 2 β Δ1

(73)

With blm2 = 1.361 and G, the Ginsburg number which was introduced to find the 1 distance where the fluctuations become negligible. Moreover, q = gr where g ∝ G and r is the distance from critical point. Finally, the resulting expression for the pressure (pure fluid) is:

P(T , v ) =

RT v0C

⎡ v0C ⎢− ⎢⎣ vC

⎛ ∂ΔA ⎞ v ⎜⎜ ⎟⎟ + P 0 + 0C vC ⎝ ∂η ⎠T

⎛ ∂K ⎞ ⎤ ⎜⎜ ⎟⎟ ⎥ ⎝ ∂η ⎠T ⎥⎦

(74)

Figure 6 presents the results obtained with CO for temperature higher than the critical temperature of CO . It can be seen that in modeling the renormalization theory can be used with cubic equation of state in order to improve the representation of the PVT properties in the critical region. Unfortunately, this model has too many parameters and so it is very difficult to easily apply this equation in industry and also to adjust parameters from experimental data. 2

2

23

60

Pressure  /MPa

50 40 30 20 10 0 0

5000

10000

15000

20000

25000

30000

-­3

Molar  density/mol.m

Figure 5: Comparison of crossover Patel-Teja equation of state (in red) and classical Patel-Teja equation of state (in black) calculation of density for CO at 313 K. (Δ): Calculated with the Span and Wagner equation for CO2 [43]. 2

CONCLUSION Industrial applications need various models depending on the operating conditions. Thus for long decades, modelling of thermodynamic properties has been approached through several ways. At low pressures dissymmetric methods with equation of state for vapour phase and GE models for liquid phase are preferred and more again when dealing with very non ideal mixtures. At medium and high pressures, symmetric methods implying one equation of state for all phases are preferred. The most used equations in industrial world are the cubic equation of state even if PVT properties are not well represented. If VLE and PVT properties are of interest volume translations in cubic EoS can give satisfactory results but only on small pressure ranges. For non ideal mixtures we have seen that G models can be used with success, even in the symmetric approach, through the mixing rules. Near the critical point crossover corrections give very good results. Convenient tools implying combinations of means are now available for full representation of thermodynamic properties and for accurate development of SFC processes. E

REFERENCES: [1] VAN DER WAALS, J. D., Over de Continuiteit van den Gas- en Vloestoftoestand. (Über die Kontinuittät des Gas- und Flüssigkeitszustands) , 1873, Dissertation, Universität Leiden, Niederlande, deutsche Übersetzung, Leipzig, (1899) [2] VAN KONYNENBURG, P.H., SCOTT, R.L., Philos. Trans. R. Soc., vol. 298, 1980, p. 495 [3] COQUELET, C., CHARETON, A., VALTZ, A., BABA-AHMED, A., RICHON, D., J. Chem. Eng. Data, Vol. 48, 2003, p. 317 24

[4] HEIDEMANN, R.A., KHALIL, A.M., AIChE J., Vol. 26, 1980, p. 769 [5] REDLICH, O., KISTER, A.T., Ind. Eng. Chem., vol. 40, 1948, p. 345 [6] WILSON, G.M., J. Am. Chem. Soc., Vol. 86, 1964, p. 127 [7] RENON, H., PRAUSNITZ, J.M., AIChE J., Vol. 14, 1968, p. 135 [8] ABRAMS, D.S., PRAUSNITZ, J.M., AIChE J., Vol. 21, 1975, p. 116 [9] FREDENSLUND, Aa., JONES, R.L., PRAUSNITZ, J.M., AIChE J., Vol. 21, 1975, p. 1086 [10] TSONOPOULOS, C., AIChE J., Vol. 20, 1974, p. 263 [11] REDLICH, O., KWONG, J. N. S., Chem. Rev, Vol. 44, 1949, p. 233 [12] PENG, D. Y., ROBINSON, D. B., Ind. Eng. Chem. Fundam. Vol. 15, 1976, p. 59 [13] PATEL, N. C., TEJA, A. S., Chem. Eng. Sci., Vol. 37, 1982, p. 463 [14] HARMENS, A., KNAPP, H., Ind. Eng. Chem. Fundam., Vol. 19, 1980, p. 291 [15] SOAVE, G., Chem. Eng. Sci., Vol. 4, 1972, p. 1197 [16] MATHIAS, P. M., COPEMAN, T. W., Fluid Phase Equilib., Vol. 13, 1983, p. 91 [17]  TREBBLE, M. A., BISHNOI, P. R., Fluid Phase Equilib., Vol. 35, 1987, p. 1 [18] TWU, C. H., COON, J. E., CUNNINGHAM, J. R., Fluid Phase Equilib., Vol. 105, 1995, 105, p. 9 [19] COQUELET,  C.,  CHAPOY,  A.,  RICHON,  D., Int. J. Thermophysics, Vol. 25, 2004, p. 133 [20] HURON, M.J., VIDAL, J., Fluid Phase Equilib., Vol. 3, 1979, p. 255 [21] WONG, D.S.H., SANDLER, S.I., AIChE., Vol. 38, 1992, p. 671 [22] MICHELSEN, M.L., Fluid Phase Equilib., Vol. 60, 1990, p. 213 [23] DAHL, S., MICHELSEN, M. L., AIChE J., Vol. 36, 1990, p. 1829   [24] HOLDERBAUM, T., GMEHLING, J., Fluid Phase Equilib., Vol. 70, 1991, p. 251 [25] MANSOORI, G.A., LELAND, T.W., J. Chem. Soc. Faraday Trans., Vol. 68, 1972, p. 320 [26] SKJOLD-JORGENSEN, S., Fluid Phase Equilib., Vol. 16, 1984, p. 317 25

[27] HUANG, S.H., RADOSZ, M., Ind. Eng. Chem. Res., Vol. 30, 1990, p. 1994 [28] WERTHEIM, M.S., J. Stat. Phys., Vol. 35, 1984, p. 19 [29] WERTHEIM, M.S., J. Stat. Phys., Vol. 35, 1984, p. 35 [30] WERTHEIM, M.S., J. Stat. Phys., Vol. 42, 1986, p. 459 [31] WERTHEIM, M.S., J. Stat. Phys., Vol. 42, 1986, p. 477 [32] BARKER, J.A., HENDERSON, D., J. Chem. Phys., Vol. 47, 1967, p. 2856 [33] GROSS, J., SADOWSKI, J., Ind. Eng. Chem. Res., Vol. 40, 2001, p.1244 [34] CHEN, S.S., KREGLEWSKI, A., Ber. Bunsen-Ges., Vol. 81, 1977, p. 1048. [35] BOUBLIK, T, J. Chem. Phys., Vol. 53, 1970, p. 471 [36] MANSOORI, G.A., CARNAHAN, N. F., STARLING, K.E., LELAND, T.W., J. Chem. Phys., Vol. 54, 1971, p. 1523 [37] KISELEV, S. B., Fluid Phase Equilib., Vol. 147, 1998, p. 7 [38] NICOLL, J.F., BHATTACHAFJEE, J.K., Phys. Rev. B, Vol. 23, 1981, p. 389 [39] NICOLL, J.F., ALBRIGHT, P.C., Phys. Rev. B, Vol. 31, 1985, p. 4576 [40] NICOLL, J.F., ALBRIGHT, P.C., Phys. Rev. B, Vol. 34, 1986, p. 1991 [41] CHEN, Z.Y., ALBRIGHT, P.C., SENGERS, J.V., Phys. Rev. A, Vol. 41, 1990, p. 3161 [42] CHEN, Z.Y., ABBACI, A., TANG, S., SENGERS, J.V., Phys. Rev. A, Vol. 42, 1990, p. 4470 [43] SPAN, R., WAGNER, W., J. Phys. Chem. Ref. Data, Vol. 25, 1996, p. 1509

26