Flory-Huggins theory: Difference between revisions

From SklogWiki
Jump to navigation Jump to search
No edit summary
No edit summary
 
(27 intermediate revisions by 5 users not shown)
Line 1: Line 1:
The '''Flory-Huggins theory''' for [[solutions]] of [[polymers]] was developed by [[Maurice L. Huggins ]] (Ref. 1) and [[Paul J. Flory]] (Ref. 2). The Flory-Huggins theory defines the volume of a [[polymers |polymer system]] as a lattice which is divided into small subspaces (called sites) of the same volume. In the case of polymer solutions, the solvent is assumed to occupy single sites, while each polymer chain occupies <math>n</math> sites. The repulsive forces in the system are modelled by requiring each lattice site to be occupied by only a single segment. Attractive interactions between non-bonded sites are included at the lattice neighbour level. Assuming random and ideal mixing of two polymers, i.e. mixing volume <math>\Delta V_m = 0</math>, it is possible to obtain the well-known expression for the combinatorial [[entropy]] of mixing <math>\Delta S_m</math> per site of the Flory Huggins theory,
The '''Flory-Huggins theory''' (although chronologically speaking it should be known as the Huggins-Flory theory <ref>Paul J. Flory in [http://garfield.library.upenn.edu/classics1985/A1985AFW2600001.pdf  Citation Classic '''18''' p. 18 May 6 (1985)]</ref>) for [[solutions]] of [[polymers]] was developed by [[Maurice L. Huggins ]] <ref>[http://dx.doi.org/10.1063/1.1750930 Maurice L. Huggins "Solutions of Long Chain Compounds", Journal of Chemical Physics '''9''' p. 440 (1941)]</ref> and [[Paul J. Flory]] <ref>[http://dx.doi.org/10.1063/1.1750971 Paul J. Flory "Thermodynamics of High Polymer Solutions", Journal of Chemical Physics '''9''' pp. 660-661 (1941)]</ref><ref>[http://dx.doi.org/10.1063/1.1723621 Paul J. Flory "Thermodynamics of High Polymer Solutions", Journal of Chemical Physics '''10''' pp. 51-61 (1942)]</ref>, following the work by Kurt H. Meyer <ref>[http://dx.doi.org/10.1002/hlca.194002301130 Kurt H. Meyer "Propriétés de polymères en solution XVI. Interprétation statistique des propriétés thermodynamiques de systèmes binaires liquides", Helvetica Chimica Acta '''23''' pp. pp. 1063-1070 (1940)]</ref>. The description can be easily generalized to the case of polymer mixtures. The Flory-Huggins theory defines the volume of a [[polymers |polymer system]] as a lattice which is divided into microscopic subspaces (called sites) of the same volume. In the case of polymer solutions, the solvent molecules are assumed to occupy single sites, while a polymer chain of a given type, <math>i</math>, occupies <math>n_i</math> sites. The repulsive forces in the system are modelled by requiring each lattice site to be occupied by only a single segment. Attractive interactions between non-bonded sites are included at the lattice neighbour level. Assuming random and ideal mixing, i.e. mixing volume <math>\Delta V_m = 0</math>, it is possible to obtain the well-known expression for the combinatorial [[entropy]] of mixing <math>\Delta S_m</math> per site of the Flory-Huggins theory for the general case of a mixture of two components, A (polymer or solvent) and polymer B (Eq. 10.1 of Ref. 3):


:<math>\Delta S_m = -R \left[ \frac{\phi_A}{n_A} \ln \phi_A + \frac{\phi_B}{n_B} \ln \phi_B \right]</math>
:<math>\Delta S_m = -R \left[ \frac{\phi_A}{n_A} \ln \phi_A + \frac{\phi_B}{n_B} \ln \phi_B \right]</math>


where <math>\phi_i</math> is the [[volume fraction]] of the component <math>i</math> and <math>n_i</math> is the number of segments in each type of polymer chain, and <math>R</math> is the [[molar gas constant]]. Applying the concept of regular solutions and assuming all pair interactions in the framework of a mean-field theory yields for the [[enthalpy]] of mixing  
where <math>\phi_i</math> is the [[volume fraction]] of the component <math>i</math>, and <math>R</math> is the [[molar gas constant]]. Applying the concept of regular solutions and assuming all pair interactions in the framework of a mean-field theory yields for the [[enthalpy]] of mixing (Eq. 20 of Chapter XII in <ref>Paul J. Flory "Principles of Polymer Chemistry" Cornell University Press (1953) ISBN 0801401348</ref>):


:<math>\Delta H_m = RT \chi \phi_A \phi_B</math>
:<math>\Delta H_m = RT \chi \phi_A \phi_B</math>


where <math>T</math> is the absolute [[temperature]].
where <math>T</math> is the absolute [[temperature]].
According to the preceding equations, the Flory-Huggins equation can be expressed in terms of the [[Gibbs energy function]] of mixing for a binary system
According to the preceding equations, the Flory-Huggins equation can be expressed in terms of the [[Gibbs energy function]] of mixing for a binary system <ref>[http://dx.doi.org/10.1063/1.1724018 Robert L. Scott and Michael Magat "The Thermodynamics of High‐Polymer Solutions: I. The Free Energy of Mixing of Solvents and Polymers of Heterogeneous Distribution", Journal of Chemical Physics '''13''' pp.  172-177 (1945)]</ref>:


:<math>\Delta G_m = -RT \left[ \frac{\phi_A}{n_A} \ln \phi_A + \frac{\phi_B}{n_B} \ln \phi_B  + \chi \phi_A \phi_B\right]</math>
:<math>\Delta G_m = RT \left[ \frac{\phi_A}{n_A} \ln \phi_A + \frac{\phi_B}{n_B} \ln \phi_B  + \chi \phi_A \phi_B\right]</math>


where <math>\chi</math> is the Flory-Huggins binary interaction parameter, defined as:
where <math>\chi</math> is the dimensionless Flory-Huggins binary interaction parameter (similar to the [[Johannis Jacobus van Laar |van Laar]] [[heat of mixing]] <ref>[http://www.dwc.knaw.nl/DL/publications/PU00013947.pdf J. J. van Laar "On the latent heat of mixing for associating solvents", Proceedings of the Koninklijke Nederlandse Akademie van Wetenschappen '''7''' pp. 174-177 (1905)]</ref>), which can be expressed as (Eq. 21 of Chapter XII):


:<math>\chi \approx \epsilon_{ij} - \frac{(\epsilon_{ii}-\epsilon_{jj})}{2} </math>
:<math>\chi = \frac{z\Delta w_{AB}}{RT}</math>


where <math>\epsilon_{ij}</math> is the net energy associated with two neighbouring lattice sites of the different
where <math>z</math> is the coordination number and (Eq. 17 of Chapter XII)
 
:<math>\Delta w_{AB} = w_{AB} - \frac{(w_{AA}+w_{BB})}{2} </math>
 
where <math>w_{AB}</math> is the net energy associated with two neighbouring lattice sites of the different
polymer segments for the same type or for the different types of polymer chains. Although the theory considers <math>\chi</math> as a fixed parameter, experimental data reveal that actually <math>\chi</math> depends on such quantities as temperature, concentration, [[pressure]], molar mass, molar mass distribution. From a theoretical point of view it may also depend on model parameters as the coordination number of the lattice and segment length.
polymer segments for the same type or for the different types of polymer chains. Although the theory considers <math>\chi</math> as a fixed parameter, experimental data reveal that actually <math>\chi</math> depends on such quantities as temperature, concentration, [[pressure]], molar mass, molar mass distribution. From a theoretical point of view it may also depend on model parameters as the coordination number of the lattice and segment length.


Line 24: Line 28:


Therefore:
Therefore:
*Positive values of <math>\chi</math> necessarily lead to incompatibility for polymers of high molecular weight.
*Positive values of <math>\chi</math> necessarily lead to immiscibility for polymer mixtures of high molecular weight.
*Mixing always take place if the <math>\chi</math> parameter is negative. Compatible mixtures with negative <math>\chi</math> are due to specific interactions between given polymer segments. Compatibility can be induced by introducing opposite charges or by the presence of a copolymer containing A and B segements.
*Polymer mixing always take place if the <math>\chi</math> parameter is negative. Miscible polymer mixtures with negative <math>\chi</math> exist due to specific interactions between given polymer segments. Miscibility or compatibility can be induced by several methods. For instance, introducing opposite charges in the different polymers or adding a copolymer containing A and B segments.
*For a polymer solution, the critical Flory-Huggins parameter is close to <math>1/2</math>.
*For polymer solutions (whose sites have the volume of a solvent molecule, <math>n_A</math>=1), the critical Flory-Huggins parameter is close to <math>1/2</math>. The temperature corresponding to this value <math>\chi</math>=<math>1/2</math> would be the critical temperature if the polymer is infinitely long and defines the [[theta solvent | theta temperature]] of the polymer-solvent system. Good solvent systems show significantly smaller positive values of <math>\chi</math>, e.g. 0.2.
*For polymer mixtures, <math>\chi</math> should be referred to the arbitrarily chosen microscopic volume defined as a site, e.g. 100 Angstroms. <math>\chi</math> values can be positive or negative and they are usually very small in absolute value for compatible or near to compatible blends <ref>[ N.  P. Balsara "Thermodynamics of Polymer Blends", in J.  E. Mark, editor, “Physical Properties of Polymers Handbook” AIP Press, pp. 257-268, (1996) ISBN 1563962950] </ref>


The <math>\chi</math> parameter is somewhat similar to a [[second virial coefficient]] expressing binary interactions between molecules and, therefore, it usually shows a linear dependence of <math>1/T</math>
The <math>\chi</math> parameter is somewhat similar to a [[second virial coefficient]] expressing binary interactions between molecules and, therefore, it usually shows a linear dependence of <math>1/T</math>


:<math>\chi(T) = A + \frac{B}{T}</math>
:<math>\chi(T) = a + \frac{b}{T}</math>


where <math>A</math> and <math>B</math> are assumed to be constants, but can actually depend on density,
where <math>a</math> and <math>b</math> are assumed to be constants, but can actually depend on density,
concentration, molecular weight, etc. A usual interpretation is that <math>A</math> represents an [[enthalpy |enthalpic]] quantity and <math>B</math> an [[entropy | entropic]] contribution, although both <math>A</math> and <math>B</math> are actually empirical parameters. According to this description, the systems should show an upper critical temperature. Many polymer systems and solutions, however, show an increase of <math>\chi</math> for increasing temperatures (negative entropic contribution) what implies the existence of a lower critical temperature.
concentration, molecular weight, etc. A usual interpretation is that <math>a</math> represents an [[enthalpy |enthalpic]] quantity and <math>b</math> an [[entropy | entropic]] contribution, although both <math>a</math> and <math>b</math> are actually empirical parameters. According to this description, the systems should show an upper critical temperature. Many polymer mixtures and solutions, however, show an increase of <math>\chi</math> for increasing temperatures (negative entropic contribution) what implies the existence of a lower critical temperature.




==References==
==References==
#[http://dx.doi.org/10.1063/1.1750930 Maurice L. Huggins "Solutions of Long Chain Compounds", Journal of Chemical Physics '''9''' p. 440 (1941)]
<references/>
#[http://dx.doi.org/10.1063/1.1723621 Paul J. Flory "Thermodynamics of High Polymer Solutions", Journal of Chemical Physics '''10''' pp. 51-61 (1942)]
;Related reading
*[http://dx.doi.org/10.1146/annurev.pc.02.100151.002123 P. J. Flory, and W. R. Krigbaum "Thermodynamics of High Polymer Solutions", Annual Review of Physical Chemistry '''2''' pp. 383-402 (1951)]
[[Category: Polymers]]
[[Category: Polymers]]

Latest revision as of 11:34, 29 August 2012

The Flory-Huggins theory (although chronologically speaking it should be known as the Huggins-Flory theory [1]) for solutions of polymers was developed by Maurice L. Huggins [2] and Paul J. Flory [3][4], following the work by Kurt H. Meyer [5]. The description can be easily generalized to the case of polymer mixtures. The Flory-Huggins theory defines the volume of a polymer system as a lattice which is divided into microscopic subspaces (called sites) of the same volume. In the case of polymer solutions, the solvent molecules are assumed to occupy single sites, while a polymer chain of a given type, , occupies sites. The repulsive forces in the system are modelled by requiring each lattice site to be occupied by only a single segment. Attractive interactions between non-bonded sites are included at the lattice neighbour level. Assuming random and ideal mixing, i.e. mixing volume , it is possible to obtain the well-known expression for the combinatorial entropy of mixing per site of the Flory-Huggins theory for the general case of a mixture of two components, A (polymer or solvent) and polymer B (Eq. 10.1 of Ref. 3):

where is the volume fraction of the component , and is the molar gas constant. Applying the concept of regular solutions and assuming all pair interactions in the framework of a mean-field theory yields for the enthalpy of mixing (Eq. 20 of Chapter XII in [6]):

where is the absolute temperature. According to the preceding equations, the Flory-Huggins equation can be expressed in terms of the Gibbs energy function of mixing for a binary system [7]:

where is the dimensionless Flory-Huggins binary interaction parameter (similar to the van Laar heat of mixing [8]), which can be expressed as (Eq. 21 of Chapter XII):

where is the coordination number and (Eq. 17 of Chapter XII)

where is the net energy associated with two neighbouring lattice sites of the different polymer segments for the same type or for the different types of polymer chains. Although the theory considers as a fixed parameter, experimental data reveal that actually depends on such quantities as temperature, concentration, pressure, molar mass, molar mass distribution. From a theoretical point of view it may also depend on model parameters as the coordination number of the lattice and segment length.

For polymers of high molecular weight (i.e. ) the entropic contribution is very small and the miscibility or immiscibility of the system mainly depends on the value of the enthalpy of mixing. In this case, miscibility can only be achieved when is negative. Assuming a temperature-dependent parameter, T vs. phase-separation diagrams can be constructed. For long polymers, miscibility can only be achieved when . The parameter at the critical point can be obtained from the definition of the critical point and the Flory-Huggins expression for the free-energy of mixing. The result is

Therefore:

  • Positive values of necessarily lead to immiscibility for polymer mixtures of high molecular weight.
  • Polymer mixing always take place if the parameter is negative. Miscible polymer mixtures with negative exist due to specific interactions between given polymer segments. Miscibility or compatibility can be induced by several methods. For instance, introducing opposite charges in the different polymers or adding a copolymer containing A and B segments.
  • For polymer solutions (whose sites have the volume of a solvent molecule, =1), the critical Flory-Huggins parameter is close to . The temperature corresponding to this value = would be the critical temperature if the polymer is infinitely long and defines the theta temperature of the polymer-solvent system. Good solvent systems show significantly smaller positive values of , e.g. 0.2.
  • For polymer mixtures, should be referred to the arbitrarily chosen microscopic volume defined as a site, e.g. 100 Angstroms. values can be positive or negative and they are usually very small in absolute value for compatible or near to compatible blends [9]

The parameter is somewhat similar to a second virial coefficient expressing binary interactions between molecules and, therefore, it usually shows a linear dependence of

where and are assumed to be constants, but can actually depend on density, concentration, molecular weight, etc. A usual interpretation is that represents an enthalpic quantity and an entropic contribution, although both and are actually empirical parameters. According to this description, the systems should show an upper critical temperature. Many polymer mixtures and solutions, however, show an increase of for increasing temperatures (negative entropic contribution) what implies the existence of a lower critical temperature.


References[edit]

Related reading